TH2010 Campbell Paul
TH2010 Campbell Paul
TH2010 Campbell Paul
THESE
Présentée
du DIPLOME DE DOCTORAT
(arrêté du 7 août 2006)
tel-00708600, version 1 - 15 Jun 2012
par
PAUL CAMPBELL
Directeurs de thèse :
Mme. Catherine SANTINI & M. Yves CHAUVIN
Jury: Rapporteurs:
Mme. Anja-Verena MUDRING M. Christopher HARDACRE
M. Agilio PADUA M. Laurent DOUCE
M. Stéphane DANIELE
M. Bruno CHAUDRET (Président)
tel-00708600, version 1 - 15 Jun 2012
2
UNIVERSITE CLAUDE BERNARD - LYON 1
COMPOSANTES SANTE
Faculté de Médecine Lyon Sud – Charles Mérieux Directeur : M. le Professeur F-N. Gilly
UFR d’Odontologie Directeur : M. le Professeur D. Bourgeois
Institut des Sciences Pharmaceutiques et Biologiques Directeur : M. le Professeur F. Locher
Institut des Sciences et Techniques de Réadaptation Directeur : M. le Professeur Y. Matillon
Département de Biologie Humaine Directeur : M. le Professeur P. Farge
4
Les travaux exposés dans ce mémoire ont été réalisés entre novembre 2007 et octobre
2010 au Laboratoire de Chimie, Catalyse, Polymérisation et Procédés dans l’équipe de Chimie
OrganoMétallique de Surface, unité mixte CNRS-CPE Lyon. Je suis reconnaissant envers
Monsieur Gérard Pignault, directeur de CPE Lyon, pour m’avoir accueilli dans ses locaux.
Mes remerciements vont tout d’abord à Madame Bernadette Charleux, qui a bien voulu
m’accueillir au sein de son laboratoire et Monsieur Jean-Marie Basset, qui m’a
chaleureusement accueilli au sein de son équipe.
J’adresse mes plus vifs remerciements à Monsieur Yves Chauvin, Directeur de Recherche
honoraire à l’Institut Français du Pétrole, agissant en qualité de co-directeur. En dépit de son
éloignement, il a fait preuve d’une grande disponibilité et a été une source constante de conseils
et de réflexions. Je lui suis très reconnaissant de sa contribution au développement de ce projet
et de m’avoir fait profiter de ses connaissances et de sa grande expérience en chimie.
Je tiens à remercier Madame Catherine Santini, Directeur de Recherche au CNRS, qui a
tel-00708600, version 1 - 15 Jun 2012
encadré mes travaux de thèse. Je la remercie tout particulièrement pour l’aide et le soutien
qu’elle m’a apportés durant ces trois années.
Que Madame Anja-Verena Mudring, Professeur à la Ruhr Universität Bochum, et
Messieurs Christopher Hardacre, Directeur de l’école de chimie à la Queen’s University Belfast,
Laurent Douce, Directeur de Recherche au CNRS, Agílio Pádua, Professeur à l’Université
Blaise Pascale Clermont Ferrand, Bruno Chaudret, Directeur de Recherche au CNRS et
Stéphane Daniele, Professeur à l’Université Lyon 1, soient vivement remerciés de l’honneur
qu’ils m’ont fait en acceptant de juger ce mémoire.
Ce travail a été finance par la Ministère de l’enseignement supérieur et de la recherche
et par l’Agence Nationale de la Recherche, projet CALIST (ANR-07-CP2D-02-03). Je remercie
tout particulièrement Ajda Podgoršek, Gorka Salas, Margarida Costa Gomes et Karine
Philippot, collaborateurs sur ce projet, pour l'aide et les nouvelles idées qu'ils m'ont apportés. Je
tiens également à remercier Anne Baudouin, François Bayard, Bernard Fenet, Denis Bouchu,
Vincent Collière, et Jacinto Sá pour leurs analyses et leur aide.
Je souhaite également remercier tous les étudiants en stage avec lesquels j’ai eu le grand
plaisir de travailler. Je remercie particulièrement Georgina Fraser et Magali Pillot, étudiantes
en stage de Master, qui ont contribué de façon significative à l’aboutissement de ces travaux.
Que tous les membres du laboratoire trouvent dans ces quelques lignes l’expression de
ma plus profonde gratitude. Je pense en particulier à Jeff, Etienne, Marco, Raphael, Nicolas,
Layane et Vickie avec lesquels j’ai vécu ces trois années dans une ambiance qui fut toujours
conviviale.
5
The work presented in this thesis was carried out between November 2007 and October
2010 in the “Laboratoire de Chimie, Catalyse, Polymères et Procédés” in the team “Chimie
OrganoMétallique de Surface”, CNRS-CPE Lyon. I am grateful Gérard Pignault, director of
CPE Lyon, for welcoming me in his institution.
My warm thanks go firstly to Bernadette Charleux, for welcoming me into her laboratory
and to Jean-Marie Basset, for his willingness to accept me as part of his team.
I am indebted to Mr. Yves Chauvin, honorary research director at the “Institut Français
du Pétrole,” acting as my co-supervisor. Despite his distance, he was always available and has
been a constant source of thought and advice. I am grateful to him for his contribution in the
development of this project and for giving me the opportunity to learn from his great knowledge
and experience in chemistry.
It is difficult to overstate my gratitude towards my Ph.D. supervisor, Catherine Santini,
CNRS Research Director, without whom, this thesis would not have been possible. With her
tel-00708600, version 1 - 15 Jun 2012
enthusiasm, her inspiration, and her great efforts to encourage and give sound ideas and advice,
these three years have been made tremendously enjoyable.
I also extend my sincere gratitude to Anja-Verena Mudring, Professor at Ruhr
Universität Bochum, Christopher Hardacre, Head of the School of Chemistry, Queen’s
University Belfast, Laurent Douce, CNRS Research Director, Agílio Pádua, Professor at
Université Blaise Pascale Clermont Ferrand, Bruno Chaudret, CNRS Research Director and
Stéphane Daniele, Professor at Université Lyon 1, for doing me the honour of reviewing this
thesis.
This work was funded by the “Ministère de l’enseignement supérieur et de la recherché”
and by the “Agence Nationale de la Recherche” under the project CALIST (ANR-07-CP2D-02-
03). I would particularly like to thank Ajda Podgoršek, Gorka Salas, Margarida Costa Gomes
and Karine Philippot, collaborators on this project, for the help and ideas and fruitful discussion
that they provided. I also wish to thank Anne Baudouin, François Bayard, Bernard Fenet, Denis
Bouchu, Vincent Collière, and Jacinto Sá for their analyses and help in interpretation.
It has been a great pleasure to work alongside many undergraduate students during these
three years and I would especially like to give a mention to Georgina Fraser and Magali Pillot,
Master’s students, who both made significant contributions to the outcome of this work.
Last but not least, I would like to express my deepest gratitude to my many colleagues for
providing a stimulating and fun environment in which to learn and grow. I am above all grateful
to Jeff, Etienne, Marco, Raphael, Nicolas, Layane and Vickie for all the emotional support,
camaraderie and particularly for the entertainment that they provided.
6
Abbreviations and acronyms
Units
h: hour min: minute mL: millilitre
g: gram mg: milligram mol: mole
°C: degree Celsius Hz: Hertz K: degree Kelvin
ppm: part per million
Techniques
ESI: ElectroSpray Ionisation Mass Spectrometry
GC: Gas Chromatography
FID: Flame Ionisation Detector
SAXS: Small Angle X-ray Scattering
XRD: X-Ray diffraction
XPS: X-ray photoelectron spectroscopy
NMR: Nuclear Magnetic Resonance
į: chemical shift J: coupling constant
tel-00708600, version 1 - 15 Jun 2012
Chemicals
IL: Ionic Liquid
Im: imidazolium
C1: methyl C2: ethyl C4: butyl C6: hexyl C8: octyl C10: decyl
C4//: but-3-enyl Bz: benzyl
[C1C4Im]: 1-butyl-3-methylimidazolium [C1C1C4Im]: 1-butyl-2,3-dimethylimidazolium
[OTf]: trifluoromethylsulfonate [NTf2]: bis-(trifluoromethylsulphonyl)imide
[BF4]: tetrafluoroborate
Ru(COD)(COT): (1,5-cycooctadiene)(1,3,5-cyclooctatriene)ruthenium
Ni(COD)2: bis(1,-5(-cyclooctadiene)nickel
NP: nanoparticle
7
tel-00708600, version 1 - 15 Jun 2012
8
Abstract – Résumé
their surface. These have both been evidenced through isotopic labelling experiments analysed by NMR
and mass spectrometry. Another advantage of ILs is that they provide an interesting medium for catalytic
reactions. Studies on the influence of the IL on the catalytic performance in homogeneous catalysis have
highlighted the crucial importance of the physical-chemical parameters of ILs, in particular the viscosity,
for which a term must be included in the kinetic rate law. Using these findings, a thorough investigation
of the effect of the NP size on catalytic activity and selectivity in hydrogenation in ILs was undertaken,
confirming the importance of controlling NP size for catalytic applications.
9
tel-00708600, version 1 - 15 Jun 2012
10
Table of Contents
Table of Contents
Introduction 13
11
tel-00708600, version 1 - 15 Jun 2012
12
tel-00708600, version 1 - 15 Jun 2012
Introduction
13
Intro
tel-00708600, version 1 - 15 Jun 2012
14
Introduction
Introduction
Progress in the comprehension of metallic nanoparticles (NPs) is central due to their great
potential in the development of new and innovative materials for applications in areas such as
catalysis or microelectronics. Their small sizes, generally reported between 1 and 100 nm, lead
to unique physical-chemical properties between the bulk and molecular states, which vary
greatly with small changes in NP size. For example, the catalytic properties of NPs are largely
determined by the energy of the surface atoms, in turn controlled by the number of neighbouring
atoms, dictated by their size, as well as the presence and nature of ligands or supports.1-3
Particles less than 10 nm in diameter are particularly interesting in catalysis, due to their
high surface to volume ratio, and are also greatly influenced by the aforementioned size-effect.
tel-00708600, version 1 - 15 Jun 2012
Their preparation by most traditional methods often yields catalysts with broad size distributions.
More recent advances in the field have seen the synthesis of NPs with well controlled size
(narrow distribution), shapes and composition.4-6
NPs are only kinetically stable, the thermodynamically stable state being the bulk metal.
Consequently, NPs dissolved freely in solution tend to agglomerate and coalesce. To prevent
aggregation, NPs must be stabilised by the use of stabilising agents such as water-soluble
polymers, quaternary ammonium salts, surfactants or polyoxoanions, providing electronic and/or
steric protection.4-6
In this context, ionic liquids (ILs), defined as low temperature molten salts, have emerged
as one of the most important and most investigated classes of stabilising agent for the synthesis
and stabilisation of metal NPs. My Ph.D. thesis focuses on the overlap of these two prevailing
fields of research, studying the synthesis small size controlled NPs (< 5 nm) for use in catalysis
in situ in ILs.7
We can tune the IL moieties and reaction conditions to obtain monodisperse and
catalytically active NPs of controlled size.4, 7-9 This is because imidazolium based ILs exhibit a
3-D organisation in the liquid state due to an extended hydrogen-bond network of ionic channels,
coexisting with non-polar domains created by the grouping of lipophilic alkyl chains.
Consequently, ILs present specific solvation properties.10 Polar substrates are preferentially
dissolved in polar domains and non-polar compounds in non-polar ones.11 The non-polar
organometallic complex, Ru(COD)(COT), is expected to be concentrated in the non-polar
15
Introduction
domains of ILs. Therefore the phenomenon of crystal growth is controlled by the local
concentration of Ru(COD)(COT) and consequently limited to the size of the non-polar domains.
These play the role of nanoreactors in which the size of ruthenium nanoparticles generated in situ
can be controlled.9, 12, 13 As a result, ILs may be used for predictive size-control of RuNPs.
Another major advantage of using ILs is that stabilising additives such as ligands,
polymers and supports are not required, meaning that a maximum amount of NP surface is
available for the coordination of substrates. However, the reasons behind the stability of NPs
dispersed in ILs remain under debate.14
After a brief literature survey of the synthesis and catalytic applications of RuNPs in
various media (Chapter 1), we go on to investigate the exact stabilisation mechanism of RuNP/IL
solutions (Chapter 2). Also, we attempt to generalise the size-controlled NP synthesis of
tel-00708600, version 1 - 15 Jun 2012
transition metal nanoparticles, by investigating the synthesis of nickel and tantalum NPs in the
same media (Chapter 2).
In order to study the catalytic reactivity in ionic liquids, we must have an understanding
of how the physical-chemical properties of these media themselves may influence the outcome
of a given reaction. For this reason, we carry out a kinetic study of a simple homogeneous
hydrogenation reaction (1,3-cyclohexadiene over an Osborn-type cationic Rh catalyst) in
different IL media, simultaneously probing the thermophysical properties (viscosity, density,
diffusion) and intermolecular interactions of the IL-substrate mixtures. These findings are used
to rationalise the differences in reactivity and thus gauge the important parameters to consider
when carrying out catalysis in IL media. (Chapter 3)
Finally, using the IL media to synthesise tailor-made RuNPs of distinct sizes, we may
investigate the size-effect on catalytic performance (activity, selectivity) in the hydrogenation of
several substrates including 1,3-cyclohexadiene, taking into account our findings from Chapter 3
to ensure otherwise identical reaction conditions. (Chapter 4)
16
Introduction
References
1. M. Valden, X. Lai and D. W. Goodman, Science, 1998, 281, 1647.
2. A. T. Bell, Science, 2003, 299, 1688 –1691.
3. G. Schmid, Nanoparticles: From Theory to Application, Wiley-VCH, Weinheim, 2004.
4. H. Boennemann, K. S. Nagabhushana and in Metal Nanoclusters in Catalysis and
Materials Science: The Issue of Size Control ed. B. Corain, Schmid, G., Toshima, N.,
Elsevier B.V, Amsterdam, 2008, pp. 21-48.
5. D. Astruc, Nanoparticles and Catalysis, Wiley-VCH, Weinheim, 2008.
6. D. Astruc, F. Lu and J. R. Aranzaes, Angew. Chem. Int. Ed., 2005, 44, 7852.
7. J. Dupont and J. D. Scholten, Chem. Soc. Rev., 2010, 39, 1780-1804.
8. J. Dupont, in Nanoparticles and catalysis, Wiley-VCH, Weinheim, 2008.
tel-00708600, version 1 - 15 Jun 2012
17
tel-00708600, version 1 - 15 Jun 2012
18
I
tel-00708600, version 1 - 15 Jun 2012
Chapter 1
Ruthenium Nanoparticles
19
tel-00708600, version 1 - 15 Jun 2012
20
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
Chapter 1
Ruthenium Nanoparticles
Synthesis, Stabilisation and Applications
21
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
22
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
1.1 Introduction
chemistry. Besides the polymers and oxides that used to be employed as standard, innovative
stabilisers, media, and supports have appeared, such as dendrimers, specific ligands, ionic
liquids, surfactants, membranes, carbon nanotubes, and a variety of oxides.9-14
As one of the catalytically active noble metals, ruthenium has been widely studied in both
homogeneous catalysis, the most well known example being olefin metathesis,15, 16 and in
heterogeneous, for example the partial hydrogenation of benzene to cyclohexene (Asahi
process),17 phenol hydrogenation,18 or in the synthesis of ammonia from N2 (Haber-Bosch
process).19, 20
Over the past couple of decades, much interest has also been devoted towards
nanoparticles of ruthenium, even though it is claimed by Galetti et al. that the preparation of Ru
nanoparticles is more difficult and therefore less-investigated than other noble metals such as Pt
or Pd.18 Their various synthetic routes and catalytic properties are summarised in this chapter.
23
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
1.2 Syntheses
such as ruthenium, after thermal treatment, these surface complexes, e.g. Ru3(CO)12, tend to
decompose and agglomerate, forming small oxide-supported RuNPs (1-2 nm on SiO2, measured
by TEM).22-25 Nonetheless, these supported NPs have been found to display interesting catalytic
behaviour, for example in the hydrogenolysis and homologation of olefins.26, 27
The use of the organometallic precursor ruthenium(η4-1,5-cyclooctadiene)(η6-1,3,5-
cyclooctatriene), commonly denoted Ru(COD)(COT), for the formation of RuNPs supported on
SiO2 was first reported Kitajima et al. This method rendered very small NPs ~ 1 nm, (measured
by H2 adsorption) smaller than those produced from RuCl3 via the same method.28 This is
thought to be due to the absence of coordinating Cl-, which would be present in the latter case.
Bare RuNPs prepared in this way have been found to coordinate 2 hydride ligands per surface
site.29
Similarly, RuNPs have been supported on mesoporous silica, firstly by impregnating the
support with Ru(COD)(COT) solution in THF and then proceeding to decompose the complex.
The size of the NPs obtained varies with the pore size and the impregnation method.30
Alumina supports (Al2O3) have also been studied and give similar results. For example,
impregnation of a precursor (either RuCl3 or Ru3(CO)12) on a K-promoted alumina (added by
prior impregnation of the alumina with KOH) followed by reduction at 400 °C under a flow of
H2 produced supported RuNP catalysts. The size distribution and dispersion were found to be
superior in the case where Ru3(CO)12 was used as the precursor.31
24
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
This technique has also been extended to innovative supports such as mesoporous
polymers to create new catalytic materials as depicted in Scheme 1. This can be used as a highly
recyclable catalyst in the aqueous hydrogenation of olefins and aromatics.32
tel-00708600, version 1 - 15 Jun 2012
Although the direct synthesis of supported NPs in this fashion renders catalytically active
supports, due to the absence of ligands and availability of the NP surface, the resulting size and
size distribution, a factor which dictates catalytic performance, is difficult to control. For this
reason the deposition of ready-synthesised RuNPs onto supports has also received a great deal of
interest.
1.2.2 Solvothermal synthesis
The versatile and facile polyol method for the synthesis of nanoparticles involves firstly
suspending the metal precursor in a polyol such as ethylene glycol, before bringing the resulting
mixture to reflux and awaiting the precipitation of the metallic moieties. The reduction occurs
via the simultaneous oxidation of alcohols to aldehydes as shown in Equation (1). The resulting
metal NPs can be filtered and dried in air. Using RuCl3 as the metal precursor results in RuNPs
with an average diameter of 5 nm.33
25
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
The polyol method has been used in the presence of polyvinylpyrrolidone (PVP) to
obtain polymer-supported RuNPs by microwave-assisted syntheses.34-36 The size of the resulting
RuNPs may be controlled by the concentration of the precursor, as shown by Somorjai’s group
using Ru(acac)2 (acac = acetylacetonate) in a microwave-assisted synthesis. TEM images of the
size-controlled RuNPs produced are depicted in Figure 1. The PVP-stabilised RuNPs could then
be supported on silicon wafers for gas phase catalysis.37
tel-00708600, version 1 - 15 Jun 2012
Figure 1. TEM images of Ru NPs: (a) 2.1, (b) 2.8, (c) 3.1, (d) 3.8, (e) 5.0, and (f) 6.0 nm. TEM images were taken
using a Philips FEI Tecnai 12 machine, operated at 100 kV. 37
RuNPs produced via the polyol method, subsequently coated in dodecane thiol and
dispersed in toluene, have been found to form supramolecular self-assemblies of anisotropic
disc-shaped NPs, the structure varying with the thiol concentration, Figure 2.
26
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
tel-00708600, version 1 - 15 Jun 2012
Figure 2. Supramolecular self assemblies of hexagonal RuNP prepared via the polyol method and stabilised with
dodecane thiol.38
This method has been widely used for subsequent deposition of the NPs onto oxide
supports. For example Miyakazi et al. prepared RuNPs supported on γ-alumina via the polyol
method for use in heterogeneous catalysis. It was found that the best results for size-control were
obtained when the γ-alumina was present during the polyol reduction, as this prevented NP
agglomeration during the cooling of the polyol mixture. The RuNPs are supported on the
alumina by electrostatic forces.20 Similarly, Xu et al. used the polyol method in ethylene glycol
to stabilise RuNPs produced in situ onto MgO substrates.39
Similar to the polyol route, another solvothermal route has been used to produce RuNPs
directly from RuCl3 by heating with an anhydrous alchohol, such as methanol. Upon heating
Ru(III) may oxidise the methanol to formaldehyde, reduced itself to Ru(0) as shown in Equation
(1).40 The RuNPs produced using this method were found to be between 2 and 10 nm in size and
X-ray diffraction analysis indicated the presence of a Ru hcp phase. XPS experiments, however,
indicated a slightly oxidised surface, possibly due to coordination of residual chloride.40
1.2.3 Colloidal synthesis in organic solvents
Zero-valent organometallic precursors are generally easily decomposed under H2
atmosphere rapidly producing NPs. The organic by-products can be easily removed without
further interaction with the surface of the resulting NPs. For this reason, Chaudret’s group has
27
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
thoroughly developed the synthesis of RuNPs in organic solvents from the decomposition of the
organometallic precursor Ru(COD)(COT), Scheme 2.12, 41-43
tel-00708600, version 1 - 15 Jun 2012
Scheme 2. The synthesis of RuNPs from the decomposition of Ru(COD)(COT) in organic solvents.12
Scheme 3. Hypothesised growth and stabilisation mechanism for RuNPs in THF methanol mixtures.43
In a mixture of MeOH/THF 10:90, the size of NPs generated was found to vary inversely
with the temperature used during decomposition, Figure 3b. This could be explained by
28
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
considering the effect of the temperature on the kinetics of decomposition. At higher T, the
decomposition is accelerated and therefore nucleation is rapid leading to a large number of small
particles. At low temperature, decomposition is slow and the kinetics of coalescence dominate.46
tel-00708600, version 1 - 15 Jun 2012
Figure 3. a) variation of RuNP size with MeOH composition (left), b) variation of RuNP size with decomposition
temperature (right).44-46
Through tests of their reactivity towards O2 and CO, the resulting NPs are shown to be
“bare” and therefore have a great surface availability for catalysis. Furthermore, RuNPs
produced in such a manner have been supported on porous alumina membranes, Figure 4, and
used in the gas phase hydrogenation of butadiene.44
Figure 4. RuNPs produced in MeOH/THF mixtures from the decomposition under H 2 of Ru(COD)(COT),
supported in a porous alumina membrane.44
This synthesis has also been carried out in the presence of polymeric materials. For
example with nitrocellulose in THF, polymer-supported NPs were produced. Here, varying the
29
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
ratio of Ru/nitrocellulose was found to control the size of the resulting NPs (2 % - 1.5 nm, 10 % -
2.5 nm).47
In the presence of the polymers polyvinylpyrrolidone (PVP) or cellulose acetate (CAC),
polymer supported RuNPs of 1.1 nm or 1.7 nm result. High resolution electron microscopy
(HREM) images show that these NPs adopt a relaxed hcp structure – a = 2.66 Å, c = 4.36 Å
compared to a = 2.7058 Å and c = 4.2811 Å in bulk ruthenium, Figure 5.48
tel-00708600, version 1 - 15 Jun 2012
Colloidal solutions of RuNPs may also be prepared by using ligands such as amines in
THF. Varying the length of the alkyl chain of the amines used, as well as the quantity of amine
has been found to produce RuNPs of different sizes and shapes. Elongated RuNPs produced
using 0.5 equivalents of hexadecylamine with respect to ruthenium are shown in Figure 6.48
RuNPs prepared in THF and stabilised with hexadecylamine have been proven by H-D exchange
experiments to be covered by surface hydrides, which may provide an additional explanation for
their stability.49
Figure 6. Elongated RuNPs formed in a mixture of THF and hexadecylamine (0.5 equiv.) 48
30
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
A range of thiols has also been studied as potential ligands for the stabilisation of RuNPs
in organic solvent. However, agglomeration was noted in this case, possibly related to the
dimerisation of thiols to disulphides on the RuNP surface, Equation (2), observed by NMR
spectroscopy.48
(2)
Silanes may also be used to stabilise RuNPs in organic solvent. The grafting of the silane
to the NP surface has been observed by solid state NMR spectroscopy.50
From an environmentally-friendly point of view, aqueous catalytic processes are a highly
interesting concept. Indeed, it has been found that the careful selection of a hydrophilic
tel-00708600, version 1 - 15 Jun 2012
stabilising ligand in NP synthesis leads to RuNPs which may be redispersed in water for
catalysis.51, 52 An example depicted in Scheme 4 uses 1,3,5-triaza-7-phosphaadamantane, PTA to
stabilise water soluble RuNPs.51
Of course, synthesis of water-soluble RuNPs may also take place in situ starting from
aqueous solutions of water soluble Ru precursors.
1.2.4 Aqueous Synthesis
For synthesis of NPs in water, non-polar organometallic precursors such as
Ru(COD)(COT) are no longer suitable. Instead, metal salts are used, in the case of RuNPs,
RuCl3.
Roucoux’s group has undertaken much work on the aqueous synthesis of RuNPs from
reduction of RuCl3 with NaBH4 in the presence of cyclodextrins (Scheme 5), which stabilise the
31
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
resulting NPs by formation of inclusion complexes. The cyclodextrins, which are water-soluble,
may also modify the surface of the RuNPs for specific reactivity.53-55
Scheme 5. Schematic representation of α-, β- and γ-cyclodextrins containing 6, 7 and 8 glucopyranose units
tel-00708600, version 1 - 15 Jun 2012
respectively.54
Water-soluble RuNPs have also been achieved using similar experimental conditions but
with less complicated ligands such as ethylenediamine,56, 57 sodium acetate,58 or even polymers
such as poly(4-vinylpyridine).59 These aqueous media may be used in biphasic catalytic systems
for reactions such as hydrogenations, presenting the advantage of easy catalyst recyclability and
product separation.53-55
The possibility of producing RuNPs that can be dispersed in a variety of media by using
amphiphilic ligands has also been explored.60
1.2.5 Synthesis in ionic liquids
Due to their specific solvation properties and 3-D structural organisation in the liquid
state, ionic liquids are “supramolecular fluids” that can be used as ‘‘entropic drivers’’ for the
spontaneous, well-defined and extended ordering of nanoscale structures. This includes the
synthesis of NPs, which can take place with size-control and in the absence of stabilising
additives. The number of publications in this field is increasing exponentially, and these have
recently been comprehensively reviewed.14
RuNPs were first generated “by accident” in ionic liquids. During an investigation of the
catalytic hydrogenation behaviour of RuO2 in ionic liquids, Hg and CS2 poisoning tests
confirmed that the catalytically active species was in fact colloidal Ru(0). The presence of NPs
32
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
was confirmed by TEM, meaning that under mild hydrogenation conditions, RuO2 was reduced
forming a stable suspension of small NPs in the ILs.61, 62
Organometallic precursors such as bis(2-methylallyl)(η4-1,5-cyclooctadiene)ruthenium63
and Ru(COT)(COD) have also been used as a RuNP precursor in ILs.62 Ru(COD)(COT),
particularly in imidazolium ILs based on the NTf2 ion, renders interesting results.64
Firstly, work undertaken in our laboratory uncovered that, unlike in traditional solvents,46
decreasing the temperature of decomposition to 0 °C (hence decreasing rate of nucleation) leads
to smaller NPs. This could only be explained by considering the organisation of the ionic liquid,
which consists of rigid ionic channels and segregated lipophilic domains. The organometallic
precursor dissolves preferentially within these lipophilic domains. The size of the resulting
RuNP is dictated by the number of available nuclides in the non polar domain after
tel-00708600, version 1 - 15 Jun 2012
decomposition. At low temperature, diffusion of the precursor and nuclides between these
domains is limited, hence the growth of NPs is restricted.
This can be confirmed by observing the effect of stirring on the size distribution. Stirring
disrupts the organisation through mechanical forces and increases the diffusion of the precursor
and nuclides in solution. The resulting size distributions are therefore broader and the resulting
NPs are agglomerated.64
Further control of the size of NPs generated in these ionic liquids can be achieved by
controlling the size of the non-polar domains by altering the length of the alkyl chains.65
Ionic liquids have also been used in combination with supports to synthesise and stabilise
RuNPs of a small size. The ionic liquid 1,1,3,3-tetramethylguanadinium (TMG) lactate has been
used to immobilise NPs in mesoporous silica (SBA-15). The resulting material was found to
exhibit synergistic effects between the RuNPs, the support and the TMG+, resulting in an active
catalyst in arene hydrogenation.66
33
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
Table 1. Non-exhaustive glossary of RuNP syntheses reported in the literature, and sizes obtained.
Precursor Conditions Stabiliser/Solvent Size obtained (nm) Reference
Ru(COD)(COT) 100-400 °C, H2 SiO2 0.9-1.0 (depending on 28
(270 mbar) decomposition temperature
and loading)
RuCl3 1.5-5.3 (depending on
loading)
Ru(COD)(COT) 300 °C, H2 SiO2 2.0 ± 0.3 29
(666 mbar)
Ru3(CO)12 220 °C, H2 (1 bar) SiO2 1 25
Ru3(CO)12 120 °C, vacuum SiO2 1.4 22
RuCl3 400 °C, H2 (1 bar) Al2O3- KOH 10 31
Ru3(CO)12 400 °C, H2 (1 bar) 2
RuCl3 reflux Ethylene glycol 5 20, 33
RuCl3 150 °C, NaOAc 1,3-propanediol 4 38, 67
RuCl3 reflux Methanol 2 - 10 40
Ethanol
tel-00708600, version 1 - 15 Jun 2012
34
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
of diphosphite)
Ru(COD)(COT) 45 °C, H2 (1 bar) Polyorganophosphazanes/ 1.55 ± 0.5 74
THF
Ru(COD)(COT) 70 °C, H2 (3 bar) PTAc/THF 1.3 ± 0.2 51
Ru(COD)(COT) 25 °C, H2 (3 bar) 3-APTSd/THF 1.7 ± 0.4 75
RuCl3 r.t., NaBH4 (aq) Dodecylamine/ 4.1 76
THFe
RuCl3 r.t., NaBH4 methylated cyclodextrins/ 1 – 2.5 53-55
H2O
RuCl3 r.t., NaBH4 Methylenediamine/ 2.1 56
H2O
RuCl3 r.t., NaBH4 NaAcO/ 2.2 ± 1.0 58, 77
H2O
RuCl3 r.t., NaBH4 Poly(4-vinylpyridine)/ 1-2 59
H2O
Ru(COD)(COT) 75 °C, H2 (4 bar) C1C4Im BF4 2.6 ± 0:4 (in 57 nm 62
C1C4Im PF6 superstructures)
C1C4Im OTf
Ru(COD)(COT) 0 °C, H2 (4 bar) C1C4Im NTf2 0.9 ± 0.4 (agglomerated in 2- 64
stirred 3 nm clusters )
tel-00708600, version 1 - 15 Jun 2012
35
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
As mentioned previously, the major application for metallic NPs is in catalysis. In this
section will be discussed several literature examples of catalytic studies performed using RuNPs
produced in using different methods.
1.3.1 Nitrogen activation
Nitrogen activation for the synthesis of ammonia is an important catalytic challenge for
which the use of RuNPs is reported in the literature.39, 79 In particular, RuNPs supported on γ-
alumina are active in ammonia synthesis.31 Interestingly, it was observed that those prepared
from the impregnation and reduction of Ru3(CO)12 present a better activity than those from
RuCl3 as shown in Figure 7, despite using the same preparation conditions. This is claimed to be
tel-00708600, version 1 - 15 Jun 2012
due to the greater surface available when produced from Ru3(CO)12, emphasising the importance
of the precursor complex chosen.31 Using the polyol method to produce supported RuNPs
resulted in an even higher activity, possibly due to the weaker interactions between the NPs and
the alumina support.20
Figure 7. Dihydrogen conversion as a function of reaction temperature (N2/H2 = 1:3): (×) equilibrium values; (▲)
4.2% Ru [from Ru3(CO)12, activated under helium up to 673 K]/Al 2O3-KOH (25.4% K); (□) 4.2% Ru [from
Ru3(CO)12, activated under hydrogen up to 673 K]/Al2O3-KOH (25.4% K); (○) 5.5% Ru (from RuC13 activated
31
under hydrogen up to 673 K)/A1203-KOH (25.4% K).
36
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
The activity of Ru in hydrogenation is well known and has been used to titrate the
number of surface hydrides on RuNPs, both supported,29 and in colloidal solutions.82 Scheme 6
illustrates the reactivity of these Ru-H towards ethylene. As well as hydrogenation to give
ethane, many different alkyl species can be formed on the metal surface through hydrogenolysis
and homologation type processes. Therefore, in order to effectively titrate all surface hydrides, it
is important to quantify both the species evolved under ethylene atmosphere and those evolved
during subsequent heating under H2, to account for the surface alkyl species.
1.3.3 Arene hydrogenation
RuNPs are known to be active catalysts in aromatic hydrogenation, and many studies are
reported in the literature. Indeed, it is now commonly believed that in the hydrogenation of
aromatics, the catalytic species cannot be mono-nuclear, as a metal surface is required in order to
activate the π-conjugated system.83
A study by Su et al. compared the reactivity of RuNPs prepared by thermal reduction and
reduction under hydrogen of RuCl3 on various supports (silica and/or carbon). It was consistently
found that the RuNPs prepared through thermal reduction were more active, despite often
exhibiting larger size distributions. These findings were attributed to the more intensive contact
37
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
between the Ru and the carbon of the support, inducing enhanced hydrogen spill-over effects.
The nature of the support was found to play a crucial role.84
Supported RuNPs have also been found to be effective in the hydrogenation of N-
heterocycles, but not S–heterocycles, unlike tethered Ru(II) catalysts which catalyse effectively
both reactions.25 When supported on polyorganophophazanes, a recyclable catalyst system for
the hydrogenation of various olefins, carbonyls and aromatics results. The material may be
dissolved in environmentally friendly solvents such as water or alcohols, or used in a
heterogeneous manner.74
Colloidal solutions of RuNPs in organic solvent have also been tested in the
hydrogenation of aromatic substrates such as benzene and quinoline. In the case of benzene, a
conversion of 82 % to cyclohexane was obtained after 14 h at 80 °C under 20 bar H2
tel-00708600, version 1 - 15 Jun 2012
38
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
39
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
tel-00708600, version 1 - 15 Jun 2012
Figure 8. The synthesis of RuNPs in THF with chrial N-donor ligand, L*, the TEM image of the NPs produced and
the subsequent performance in the asymmetric hydrogenation of methylanisoles. 69
The same catalyst systems were found to be active in hydrogen transfer between iso-
propanol and acetophenone, Scheme 8. The rate of reaction was found to vary with the nature of
the chiral N-donor ligand, and a moderate enantiomeric excess (10 %) was observed in only few
cases.69
69
Scheme 8. Hydrogen transfer between iso-propanol and acetophenone.
40
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
In another example in the literature, diphosphite ligands have been successfully used to
stabilise small RuNPs in organic solvent and thus perform diastereoselective hydrogenations of
methylanisole,72, 73 while the use of P(m-C6H4SO3Na)3 in combination with a mixture ionic
liquids and water has been shown to produce a recyclable and enantio-selective system for
hydrogenation.88 Furthermore, trioctylamine-stabilised RuNPs in a mixture of alcohol and water
show a certain stereoselectivity in the hydrogenation of various substituted arenes.89
1.3.7 Oxidation
Catalytic oxidation processes, which are of particular interest in industry, may benefit
from progress in the field of nanoclusters, but only very few examples of oxidation reactions
have been reported in the literature.
One significant result has demonstrated the ability of RuNPs to oxidise cycloalkanes
tel-00708600, version 1 - 15 Jun 2012
selectively to the corresponding ketones, e.g. cyclooctane to cyclooctanone, Scheme 9. This has
been achieved with tert-butylhydroperoxide (tBHP) in a water/cyclooctane mixture, with very
high selectivity.90, 91
In another catalytic oxidation, RuNPs grafted onto hydroxyapatite have been found to be
active in the cis-dihydroxylation and oxidative cleavage of alkenes, presenting excellent
recyclability.67
The oxidation of CO to CO2 has been achieved quantitatively using RuNPs supported in a
porous alumina membrane. A slight difference was noted in activity with particle size, smaller
particles being slightly more active.92 However, this may have been related to the porous
material, as when supported on a silicon wafer and stabilised by PVP, the inverse trend in
activity with respect to NP size has been observed in the same reaction by Somorjai’s group,
Figure 9.37 This last example demonstrates the importance of the size of the NPs in catalysis, as
well as how altering the nature of the reaction medium or stabilising support may change entirely
the outcome of the reaction under investigation.
41
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
tel-00708600, version 1 - 15 Jun 2012
Figure 9. The variation of TOF and activiation energy with RuNP size in the catalytic oxidation of CO, studied by
Somorjai’s group.37
1.3.8 Miscellaneous
1.3.8.1 Cellubiose hydrolysis
Water-dispersed RuNPs have been shown to be active in a one-step hydrolysis of
cellobiose to C6 alcohols, an important reaction in the production of bio-fuels, Scheme 10.93
42
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
(3)
Very recently, RuNPs coordinated by silane ligands have been found to be highly active
and reusable as catalysts for the dehydrogenation of dimethylamine–borane, which could be an
interesting prospect in solid hydrogen storage materials.75
1.3.8.3 Synthesis of novel functional materials
RuNPs have been found to be active and efficient catalysts for growth of single-walled
tel-00708600, version 1 - 15 Jun 2012
43
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
1.4 Conclusion
As we can see, there has already been a great deal of work reported in the literature
regarding the synthesis and catalytic applications of RuNPs. The catalytic properties of these
NPs are highly dependent on their surface properties and their size. The ability to predict and
control the size of the NPs produced is therefore of the utmost importance. Also, producing NPs
free of ligands and other surface contaminants can only be advantageous in catalysis. In this
respect, the ionic liquid route, which has been shown to produce RuNPs of a controlled size
without necessitating stabilising ligands or supports, seems promising. However, in this case
little is known about the stabilisation mechanism – is the surface really ligand and contaminant
free? This is the first question answered in this thesis, in Chapter 2. Following this we look at the
tel-00708600, version 1 - 15 Jun 2012
possibility of extending the potential of ionic liquids to the controlled growth and stabilisation of
other catalytically active metals, namely nickel and tantalum, also in Chapter 2.
Ionic liquids are also reported as interesting media for catalysis, so performing catalysis
with NPs produced in situ is a very intriguing prospect. However, first we must discern how the
properties ionic liquid itself effect the outcome of the catalytic process. This question is
answered in Chapter 3 by performing a model homogeneous catalytic reaction, namely the
hydrogenation of 1,3-cyclohexadiene using an ionic Osborn-type rhodium catalyst, in different
IL media.
Finally, in Chapter 4, we perform the hydrogenation of several substrates using RuNPs
generated in situ, with precise size-control, and confirm the substantial influence that
nanoparticle size can have on the catalytic performance.
44
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
1.5 References
45
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
28. N. Kitajima, A. Kono, W. Ueda, Y. Moro-oka and T. Ikawa, J. Chem. Soc., Chem.
Commun., 1986, 674-675.
29. R. Berthoud, P. Delichere, D. Gajan, W. Lukens, K. Pelzer, J. M. Basset, J. P. Candy and
C. Coperet, J. Catal., 2008, 260, 387-391.
30. V. Hulea, D. Brunel, A. Galarneau, K. Philippot, B. Chaudret, P. J. Kooyman and F.
Fajula, Microporous Mesoporous Mater., 2005, 79, 185-194.
31. P. Moggi, G. Albanesi, G. Predieri and G. Spoto, Appl. Catal., A, 1995, 123, 145-159.
32. L. Y. Song, X. H. Li, H. N. Wang, H. H. Wu and P. Wu, Catal. Lett., 2009, 133, 63-69.
33. L. K. Kurihara, G. M. Chow and P. E. Schoen, Nanostruc. Mater., 1995, 5, 607-613.
34. R. Harpeness, Z. Peng, X. S. Liu, V. G. Pol, Y. Koltypin and A. Gedanken, J. Colloid
Interface Sci., 2005, 287, 678-684.
35. M. Zawadzki and J. Okal, Mater. Res. Bull., 2008, 43, 3111-3121.
36. X. P. Yan, H. F. Liu and K. Y. Liew, J. Mater. Chem., 2001, 11, 3387-3391.
37. S. H. Joo, J. Y. Park, J. R. Renzas, D. R. Butcher, W. Huang and G. A. Somorjai, Nano
Lett., 2010, 10, 2709-2713.
38. G. Viau, R. Brayner, L. Poul, N. Chakroune, E. Lacaze, F. Fievet-Vincent and F. Fievet,
Chem. Mater., 2003, 15, 486-494.
39. Q. C. Xu, J. D. Lin, X. Z. Fu and D. W. Liao, Catal. Commun., 2008, 9, 1214-1218.
40. S. Gao, J. Zhang, Y. F. Zhu and C. M. Che, New J. Chem., 2000, 24, 739-740.
41. B. Chaudret, C.R. Phys., 2005, 6, 117-131.
46
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
47
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
48
Chapter 1 Ruthenium Nanoparticles – Synthesis, Stabilisation and Applications
Buntkowsky and H. H. Limbach, Angew. Chem. Int. Ed., 2008, 47, 2074-2078.
83. C. M. Hagen, G. Vieille-Petit, G. Laurenczy, G. Suss-Fink and R. G. Finke,
Organometallics, 2005, 24, 1819-1831.
84. F. Su, L. Lv, F. Y. Lee, T. Liu, A. I. Cooper and X. S. Zhao, J. Am. Chem. Soc., 2007,
129, 14213-14223.
85. W. Yu, H. Liu, M. Liu and Z. Liu, J. Colloid Interface Sci., 1998, 208, 439.
86. M. Liu, W. Yu and H. Liu, J. Mol. Catal., 1999, 243, 120.
87. M. Liu, W. Yu, H. Liu and J. Zheng, J. Colloid Interface Sci., 1999, 214, 231.
88. J. Wang, J. Feng, R. Qin, H. Fu, M. Yuan, H. Chen and X. Li, Tetrahedron: Asymmetry,
2007, 18, 1643-1647.
89. F. Fache, S. Lehuede and M. Lemaire, Tetrahedron Lett., 1995, 36, 885.
90. F. Launay and H. Patin, New J. Chem., 1997, 21, 247.
91. F. Launay, A. Roucoux and H. Patin, Tetrahedron Lett., 1998, 39, 1353.
92. H. P. Kormann, G. Schmid, K. Pelzer, K. Philippot and B. Chaudret, Z. Anorg. Allg.
Chem., 2004, 630, 1913-1918.
93. N. Yan, C. Zhao, C. Luo, P. J. Dyson, H. Liu and Y. Kou, J. Am. Chem. Soc., 2006, 128,
8714-8715.
94. V. Matsura, Y. Guari, C. Reye, R. J. P. Corriu, M. Tristany, S. Jansat, K. Philippot, A.
Maisonnat and B. Chaudret, Adv. Funct. Mater., 2009, 19, 3781-3787.
49
tel-00708600, version 1 - 15 Jun 2012
50
II
tel-00708600, version 1 - 15 Jun 2012
Chapter 2
51
tel-00708600, version 1 - 15 Jun 2012
52
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Chapter 2
Generation and Stabilisation of Transition Metal
Nanoparticles in Imidazolium-Based Ionic Liquids
2.2.3 Evidence for close proximity between the RuNP surface and the alkyl chains......................... 64
2.2.4 Role of the continuous 3-D network of ionic channels in the isolation of RuNPs .................... 68
53
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
54
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
2.1 Introduction
During the last decade, the use of ionic liquids (ILs) has become a very popular route for
the generation of transition metal nanoparticles (NPs).1 The main advantage over traditional
solvents is that IL media are able to stabilise NPs in the absence of further additives such as
ligands, inhibiting metal coalescence to the bulk.2 For catalytic applications this is good news, as
the surface of the resultant NPs is ligand-free and therefore completely available for the
coordination of substrates. However, the question of precisely how ILs stabilise transition metal
NPs remains under debate.
Very little data is available in the literature to explain this phenomenon, and much is
speculative.3 Many describe “electrosteric stabilisation” whereby an electrostatic double layer
of anions and cations surround the electropositive NP surface.4, 5
tel-00708600, version 1 - 15 Jun 2012
Surface enhanced Raman spectroscopy (SERS) has been used to probe the interactions
between an imidazolium based IL and the surface of AuNPs generated in situ, providing
evidence for a planar interaction between the imidazolium ring and the NP surface, but very little
interaction between the anion and the NP.11 This is clearly contradictory to the much discussed
DLVO model. Furthermore, the coordination of the IL moieties to the NP surface could be
55
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
envisaged. Indeed, this has been evidenced for IrNPs synthesised by reduction of an ionic
organometallic complex in the ionic liquid [C1C4Im][BF4]. Subsequent treatment of the mixture
under deuterium, followed by 2H NMR analysis showed that hydrogen-deuterium exchange had
occurred on the imidazolium ring, Figure 1. This was used to evidence the formation of labile N-
heterocyclic (NHC) ligands, namely imidazolylidene, however given the acidic nature of the C 2-
H of the imidazolium ring this exchange would be facile and therefore the conclusion seems
speculative.12, 13
tel-00708600, version 1 - 15 Jun 2012
Figure 1. 2H NMR spectrum of [C1C4Im]+ after treatment under D2 in the presence of IrNPs.12, 13
If these literary observations are believed, it seems that the stabilisation of NPs is a
complex issue, dependent on the nature of both the metal and the IL in question, not to mention
the synthetic route. To the best of our knowledge, no stabilisation mechanism has been proposed
for RuNPs in ILs to date. Consequently, in this chapter we investigate and discuss the
stabilisation mechanism for RuNPs in imidazolium ILs. Following this, the possibility of
generalising our findings to other transition metals is explored.
56
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
H O
O
CF
tel-00708600, version 1 - 15 Jun 2012
S 3
S N
N N F3C O
O
H
H H
CF3 H
O N
S
O N
O N
H CF3
N H
S
O H S
F3C O O
N
N
H S
O O
CF3
Scheme 2. Representation of intermolecular organisation in [C1C4Im][NTf2] in the liquid state: ionic channel linked
through H-bonds and consequent aggregation of alkyl chains.
This structural organisation of ILs can be used as an entropic driver for the spontaneous
self-assembly of nanoscale structures with long-range order.18 For this purpose, ILs have already
been used as media for the synthesis of some zeolite-related, microporous aluminophosphates or
ordered mesoporous materials in which ILs have served both as solvents and as structure-
directing agents.19
This segregation of polar and non-polar domains in imidazolium-based ILs dictates their
interactions with given solutes. Polar or ionic species, e.g. water, will preferentially dissolve in
57
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
the polar regions, whereas non-polar lipophilic compounds, e.g. hexane, will be concentrated in
the non-polar pockets.20
The synthesis of RuNPs by direct reduction RuO2 is reported, however this leads to the
presence of an oxide surface and/or strongly coordinating water ligands, which are difficult to
remove.21 The decomposition of ( 4-1,5-cyclooctadiene)( 6-1,3,5-cyclooctatriene)ruthenium(0),
Ru(COD)(COT), by hydrogen is a well known route to obtaining RuNPs, in both organic,22, 23
and ionic liquid media, Scheme 3.24-26 The main advantage of this halogen-free synthesis is that
the only side product, cyclooctane, is easily removed under vacuum. This organometallic
complex is non-polar and therefore when dissolved in an ionic liquid is concentrated in the
aforementioned non-polar pockets. This principle has already been exploited in our laboratory in
the size-controlled synthesis of RuNPs from Ru(COD)(COT) in various imidazolium ILs (1-
tel-00708600, version 1 - 15 Jun 2012
Scheme 3. The synthesis of RuNPs from the decomposition under H 2 of Ru(COD)(COT) in ILs
58
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Interconnexion of
Linear crystal growth non-polar domains
Polar medium
in non-polar domains
2.5 2.3 ±0.8
nm
2
1.9 ±0.6
nm
Mean size (nm)
1.5
Nonpolar domains
0.5
tel-00708600, version 1 - 15 Jun 2012
(lit.)
0
0 2 4 6 8 10
Alkyl chain length (n carbons) 3
Figure 2. Graph to show the variation in RuNP size with repect to imidazlolium side alkyl chain length. (circles)26
Also plotted, experimentally measured mean diameter of non-polar domains with respect to alkyl chain length.
(squares) 15
The RuNP/IL solutions obtained via this method remain stable indefinitely under inert
atmosphere and ambient conditions, with no coalescence or precipitation observed. Given the
apparent absence of ligands, the reason behind this stability is unknown.
When carried out at low temperatures (0 °C) in the absence of stirring in 1-butyl-3-
methylimidazolium bis(trifluorosulphonyl)imide, [C1C4Im][NTf2], this procedure consistently
yields the smallest small NPs with the narrowest size distribution (mean size: 1.1 ± 0.2 nm
calculated from TEM images realised in situ).25 These conditions have been replicated
throughout this work, to obtain fresh IL-stabilised colloidal solutions, on which NMR and mass
spectroscopy experiments have been performed, to determine the key points in their stabilisation.
2.2.1 In situ evidence of surface hydrides
The presence of surface hydrides has already been reported on RuNPs in organic
solvents.23, 27-29 Furthermore, in heterogeneous catalysis, it is well-known that under hydrogen
atmosphere, the surface of late transition metals is covered by hydrides.30-32 In ILs, is such
59
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
coverage still possible? Given the current and general stabilisation theory concerning NPs in ILs,
(strong interaction between the NPs and anion or cation, DLVO model),6, 33
the formation of
surface hydrides could be inhibited. Surface hydrides on NPs in ILs have to the best of our
knowledge never been reported or envisaged.
a) HD 200Hz
H2 1200Hz (3)
(2)
tel-00708600, version 1 - 15 Jun 2012
(1)
10 9 8 7 6 5 4 3 2 1 0 ppm
b)
C2 -D C4,5 -D
(3)
(2)
0.40
0.75
3.00
C4,5 -H methyl-H
C2 -H
methyl-H
C2 -H C4,5 -H
(1)
9 8 7 6 5 4 3 2 1 ppm
1
Figure 3a) H NMR of gas phase (1) pure H2, (2) gas phase after 1 hour dynamic vacuum and 1 day under
deuterium, (3) gas phase after 4 hours dynamic vacuum and 4 days stirring under deuterium; b) (1) 1H NMR
spectrum of IL after NP formation under H2 and evacuation of H2 and cyclooctane, (2) 1H and (3) 2H NMR spectra
of the reaction medium after addition of D2.
60
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
into an airtight Young NMR tube. The 1H NMR spectrum obtained exhibited a broad peak,
(w½ = 1200 Hz) corresponding to H2, superimposed with a finer peak (w½ = 200 Hz)
corresponding to HD, the intensity increasing with reaction time, Figure 3a. This is in agreement
with previous findings in the literatue for RuNPs in organic solvents,28 or supported on oxides,31,
32
and allows us to assign the HD desorption as resulting from deuterium activation on RuNPs
and recombination with bound H atoms, as observed on metal surfaces, Scheme 4.
tel-00708600, version 1 - 15 Jun 2012
The fact that there is ~ 60 % H/D exchange on positions 2, 4 and 5 of the imidazolium
ring, as shown by 1H and 2H NMR spectra of the IL, Figure 3b, could suggest that the HD
formed is due to this exchange and not from putative surface hydrides. However, when the ionic
liquid is treated under D2 in the absence of RuNPs, no H/D exchange is observed. The observed
H/D exchange must inevitably involve RuNP surface. To prove the presence of these surface
hydrides, further experiments have been performed characteristic of metal surface reactivity,
such as hydrogenation, as detailed in the following section.
2.2.1.2 Reaction with ethylene
The reaction of alkenes with PVP- (polyvinylpyrrolidone-) and ligand-stabilised RuNPs
has recently been used to quantify the surface hydrides.29 A similar approach was transposed to a
solution of RuNPs in IL1. In the following discussion, the RuNPs were generated in 2mL of IL1
from 86 μmol of Ru(COD)(COT). Given that the dispersion of NPs, (D, i.e the ratio surface
atoms Ns per total number of atoms Nt, D = Ns/Nt), is correlated to their size,30, 35, 36 we could
estimate that for RuNPs of 1.1 nm the dispersion is 83 %, corresponding for 86 μmol of
Ru(COD)(COT) to NS ≈ 71 μmol of Ru surface atoms. This IL1-stablised RuNP solution was
flushed overnight with argon to remove dissolved hydrogen. Then the solution was stirred under
an ethylene atmosphere (105.5 mbar) at 25°C. After 24 h of reaction the system reached
equilibrium, i.e. the pressure remained constant. In the gas phase, analysed by GC, 11 ± 1 μmol
of ethane and 0.1 μmol of iso-butane were found, corresponding to at least 22 ± 1 μmol of
61
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
surface hydride: (C2H4 + H2 C2H6). Note that butenes (trans- and 1-) were also observed
(0.2 μmol), however no hydrogen would be consumed in their formation. In summary, IL1-
stabilised RuNPs reacted with ethylene at room temperature to give ethane (hydrogenation) and
surface-bound alkyl species (homologation).37, 38
To quantify the total amount of surface-bound alkyl species, the gas phase was evacuated
and substituted by H2 (120 mbar) and the medium was heated at 100 °C for 12 h. In the gas
phase, methane (4.6 μmol), ethane (8.2 μmol), propane (2.7 μmol), n-butane (1.0 μmol) and n-
pentane (0.1 μmol) were detected, corresponding to at least a further 16 μmol of hydrides.
Knowing the total amount of surface Ru atoms available (65 μmol) and the number of surface
hydrides consumed in all the hydrogenation reactions (38 μmol) it is possible to calculate the
ratio of hydrides per surface Ru atom: c.a. 0.6 H /RuS.* The combination of these results provides
tel-00708600, version 1 - 15 Jun 2012
substantial evidence for the existence of surface hydrides on the IL1-stabilised RuNPs, although
the level of error in the quantitative results means that it is difficult to establish accurately the
quantity. The amount of H adsorbed on PVP- and ligand- stabilised RuNPs in organic solvent is
reported between 1.1-1.3 H/RuS,28and between 1 and 2 H/RuS for RuNPs supported on oxides.31,
32
Note that if the IL1-stabilised solutions of RuNP were treated under high vacuum (10-6 mbar)
overnight at room temperature, instead of being flushed with argon to remove the dissolved H2,
ethylene conversion is negligible. This suggests that the surface hydrides on RuNPs are labile
and almost completely desorbed under high vacuum.39-41
Indeed, hydrides bound to metal surfaces are known to be labile and the dissociation of
H2 adsorbed onto a metal surface is reversible: M + H2 ↔ MH2 ↔ 2 M–H. In the case of Ru,
about 2/3 of the total amount of surface hydrides are reversibly bound. These H species
correspond to more weakly adsorbed hydrogen on the surface, which is easily removed by
treating the sample under vacuum at 25 °C.31, 32.This explains why the hydrogenation of ethylene
(in the absence of hydrogen) does not occur over NPs previously treated under high dynamic
vacuum (10-5 mbar) for an extended period (12 h), as very few surface hydrides remain.39-41 It is
nevertheless very interesting that the presence of hydrides on ruthenium particles in ionic liquids
is confirmed, since this is in agreement with the presence of these nanoparticles in uncharged
lipophilic domains.
*
For calculations see experimental section 2.5.6
62
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
explained by the elimination of the hydrogen atoms reversibly bound to the metal surface, which
is made likely by heating under an inert atmosphere or treatment under high vacuum. In
conclusion, the presence of surface Ru–H stabilises the RuNPs. The results of these stability
experiments are also consistent with the lability of the hydrides which can be desorbed from the
nanoparticles in vacuo or at 100 °C.
20nm
20 nm
120 20 nm
100 RuNPs
H2
Ar
80 Vacuum
Frequency
20 nm
60
40
20
0
0 1 2 3 4 5 6
Size (nm)
Figure 4. TEM images captured in situ in ionic liquids and comparative size histograms of RuNPs/[C1C4Im][NTf2]
(1.1 ± 0.2 nm) heated to 100 °C for 24h under H2, under Ar and kept under dynamic vacuum (10-5 mbar) for 24 h at
25 °C.
63
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
In further tests, samples of these NP solutions were stirred under pure CO or pure O2
atmosphere (1 bar) during a period of 24 h. It was seen that under O2 the NPs underwent slight
coalescence, growing to a size of 1.5 ± 0.4 nm, and those resulting from stirring under CO
exhibited an interesting bimodal size distribution centred at 1.3 and 1.8 nm, Figure 5. It is worth
noting that the number of RuNPs smaller than 1 nm increases probably due to the formation of
ruthenium carbonyl clusters. Due to the low concentration of RuNPs in solution, (CO) for
coordinated CO was not observed by IR spectroscopy. In comparison with H2, O2 and CO appear
to have a destabilising effect, inducing slight agglomeration.
70
50
60
40 RuNP size CO
50
tel-00708600, version 1 - 15 Jun 2012
RuNP size O2
Frequncy
Frequency
30 40
30
20
20
10
10
0 0
0.5 1.0 1.5 2.0 2.5 3.0 0.5 1.0 1.5 2.0 2.5
Size/nm Size/nm
Figure 5. Size distribution histograms of RuNPs formed in IL1 after treatment under dry O 2 (left) and CO (right).
2.2.3 Evidence for close proximity between the RuNP surface and the alkyl chains
The fact that in [C1CnIm][NTf2], the non-polar domains control the local concentration of
Ru(COD)(COT) and consequently the size of RuNPs generated in situ, could allow us to
consider them as nanoreactors in which the phenomenon of crystal growth occurs.26 If the RuNPs
are indeed encapsulated in these non-polar pockets, it would be expected that the side alkyl
groups of the ILs be in close proximity or even in interaction with the RuNP surface.
To demonstrate this, the RuNPs were synthesised under deuterium instead of hydrogen
(following otherwise identical reaction conditions), since an H-D exchange on the side group of
the IL would prove the existence of the NPs in these domains. In the case of IL1, 2H NMR
spectra show us that, as expected, deuteration occurs at all positions on the imidazolium ring due
to the relatively high acidity of these protons. Furthermore, one small peak is apparent at
~1.5 ppm, which may correspond to deuteration at position C8, i.e. the penultimate carbon of the
butyl chain, however this is not irrefutable. The intensities of the peaks in the 1H NMR spectrum
64
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
were reduced by roughly 30 % at positions C2, C4 and C5, but no significant reduction in
intensity was noted for any of the alkyl peaks. ESI mass spectra indicate the formation of mono-,
di-, tri- and a trace amount of tetra-deuterated species. This is in agreement with the higher
reactivity of the hydrogen bound to the imidazolium ring than to the alkyl side chain. Note that,
5
in contrast to findings for IrNPs, where the deuteration was found mainly at position C2, for
RuNPs deuterium incorporation occurs equally at positions C2, C4 and C5 on the imidazolium
ring.
This experiment was then carried out in ILs with unsaturated functional side groups: 1-
(but-3-enyl)-3-methylimidazolium bis(trifluorosulphonyl)imide [C1C4//Im][NTf2], IL2 and 1-
benzyl-3-methylimidazolium bis(trifluorosulphonyl)imide [C1BzIm][NTf2], IL3. These
functionalised side groups are more susceptible to H-D exchange.
tel-00708600, version 1 - 15 Jun 2012
80
70
60
50
20 nm
40
30
20
10
0
0.0 0.5 1.0 1.5 2.0 2.5
size/ nm
Figure 6. Comparative size distribution curves for RuNPs synthesised in IL1 and IL2 and TEM image captured in
situ IL for RuNPs synthesised in IL2
For IL2, 2H NMR spectra show that deuteration takes place at positions C8 and C9, i.e. the
olefinic positions, as well as at position C2. Surprisingly, isotopic exchange is observed neither at
65
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
positions C4 and C5 of the imidazolium ring, nor at positions C6 and C10. Instead, and more
interestingly, a new set of peaks appears at high field, corresponding to a deuterated butyl group
due to the deuteration and reduction of the C8=C9 double bond, Figure 7. Further heating of this
solution under deuterium at 50 °C for 12 h led to a pronounced growth of all peaks already
deuterated, but no deuteration at other positions. The intensities of the peaks in the 1H NMR
spectrum of IL2 after reaction are reduced significantly at positions C2, C8 and C9 indicating that
24 % of the ionic liquid had undergone at least one H-D exchange and NMR and ESI mass
spectra indicate that around a further 15 % had been reduced to a deuterated butyl species
corresponding to a TON of ~16 with respect to Rus at 0 °C.† Also, besides the peaks due to IL1
(labelled * in Figure 8), new peaks appeared (labelled ° in Figure 8) in the Csp2 and Csp3 regions,
corresponding to IL2isom in which there had been an isomerisation of the butenyl chain,
tel-00708600, version 1 - 15 Jun 2012
Scheme 5.
Figure 7 a) 1H NMR spectrum of [C1C4//Im][NTf2], neat IL2; b) 2H NMR spectrum of IL2 after RuNP formation at
0°C under deuterium; c) 2H NMR spectrum of IL2 after RuNP formation at 50 C under deuterium; d) 1H NMR
spectrum of [C1C4Im][NTf2], neat.
†
ESI mass spectra and calculations of TON are given in the experimental section 2.5
66
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Figure 8. 1H NMR spectra of neat IL2; IL2 after RuNP formation at 0 °C under deuterium; neat IL1.
67
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Rus.‡ Further heating of this solution under deuterium at 50 °C for 12 h led to a pronounced
growth of all peaks mentioned in the 2H NMR spectrum.
tel-00708600, version 1 - 15 Jun 2012
Figure 9. 2H NMR spectra of [C1BzIm][NTf2], IL3, after RuNP formation at 0 °C under deuterium (lower) and after
a further 12 h of heating to 50 °C under deuterium (upper).
The hydrogenation of an aryl substituent, which does not occur with any molecular
ruthenium complexes derived from Ru(COD)(COT), must occur on the nanoparticle surface.43
This observed reduction therefore clearly demonstrates the proximity between the aryl groups of
the ILs and the RuNP surface.
During the formation of RuNPs under deuterium in both IL2 and IL3, the side groups of
the imidazolium rings are saturated with deuterium. In neither case is the C2 position the primary
site for H–D exchange, in contrast to literature results.13 Furthermore, in the case of IL2, no
exchange is observed at positions C4 and C5. Note that no exchange occurs at either position C6
or C10 in either IL. These results indicate that the side chains of the imidazolium are situated in
close proximity of the RuNP surface, where activation of deuterium and C-H occurs.
2.2.4 Role of the continuous 3-D network of ionic channels in the isolation of RuNPs
Addition of coordinating substrates, such as water, solvate the rigid ionic channels of the
IL through the formation of strong hydrogen bonds with the anions and also weaker hydrogen
bonds between the H bound to the C2 of the cation.20, 44, 45 Solvation of this ionic network alters
the structure of the ionic liquid causing aggregation of the non-polar domains.46-49 The following
question arises: what happens to the RuNPs if the structure of the IL is disrupted?
‡
ESI mass spectra and calculations of TON are given in the experimental section 2.5
68
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
40
30
Frequency
20
10
tel-00708600, version 1 - 15 Jun 2012
0
0.5 1.0 1.5 2.0 2.5 3.0
Size/nm
Figure 10. TEM image of RuNPs in IL1 after treatment with water and resultant size distribution histogram.
69
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
by their solvation in different domains of the ionic liquid. TEM images of the NPs obtained are
shown in Figure 11.
a) IL1/Ru(COD)(COT) + 3%D2O + H2 à d) IL1/3%D2O + Ru(COD)(COT) + H2 à
tel-00708600, version 1 - 15 Jun 2012
Figure 11. TEM images of solutions of RuNPs synthesised in ILs in the presence of water.
70
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
As can be seen in the TEM images in Figure 11 the presence of water has a significant
impact on the organisation of the resultant NPs. For the solutions prepared with water below the
limit of solubility, whether the water be added before or after the Ru(COD)(COT) the same
pattern is observed. For 3 % water (molar), RuNPs of around 2 nm group in circular aggregates
of around 20-30 nm in diameter (a and d in Figure 11), and for 30 % water RuNPs also of around
2 nm, group in larger circular aggregates of 30-50 nm (b and e in Figure 11). NPs formed from
the dissolution of Ru(COD)(COT) into water-saturated IL1 before decomposition exhibit the
same type of agglomeration of NP, here the diameter of the circular aggregates varying from
between 30 and 150 nm (f in Figure 11). This findings support the idea of “water-pools” in wet
ILs i.e. micellar-type aggregates of water, as through diffusion measurements water has found to
not be completely miscible on the microscale.47 In this case, RuNPs appear to decorate the water-
tel-00708600, version 1 - 15 Jun 2012
§
Given the relative volatility of water with respect to the IL, we assume the light regions correspond to water, which
under the electron beam and UHV conditions will evaporate.
71
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
route and/or the ionic liquid used. We have already demonstrated the growth, size control and
stabilisation mechanism of RuNPs from the decomposition under H2 of the zero-valent
organometallic ruthenium compound Ru(COD)(COT) in imidazolium ionic liquids.
This leads to the question of whether the same would be true of NPs generated in the
same way from an analogous organometallic compound of another metal. For this reason, in this
chapter we investigate the possibility of generalising these findings to the generation of nickel
nanoparticles, NiNPs, as the organometallic complex bis(1,5-cyclooctadiene)nickel, Ni(COD)2,
analogous to Ru(COD)(COT), is commercially available.
Furthermore, as zero-valent ogranometallic complexes are often difficult to come by,
particularly those of the early transition metals, it may be interesting to look at the reductive
decomposition of alkyl-metals. Therefore, the generation of tantalum nanoparticles, TaNPs from
tel-00708600, version 1 - 15 Jun 2012
72
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Nickel is one of the most abundant and therefore least costly of the catalytically active
late transition metals – 17 USD/kg (nickel) vs. 5340 USD/kg (ruthenium) – so presents a clear
economic advantage if its catalytic potential is exploited. Furthermore, bulk nickel being
ferromagnetic, magnetic nanomaterials often exhibit interesting optical and photoelectronic
properties,51 as well as being superparamegnetic. Their magnetic properties vary with size, shape
and composition and therefore magnetic measurements are often used as complementary
characterisation.
NiNPs have been prepared by the reduction of organometallic precursors such as nickel
tetracarbonyl (Ni(CO)4), bis-(cyclopentadienyl) nickel, (Ni(C5H5)2), bis-(acetylacetate) nickel
(Ni(acac)2),52 and bis-(1,5-cyclooctadiene)nickel (Ni(COD)2,53-57 in different stabilising media.
tel-00708600, version 1 - 15 Jun 2012
For the purpose of our work Ni(COD)2 is chosen due firstly to its zero-valence and also to the
fact that the by-products of decomposition are light hydrocarbons that are easily removed under
vacuum and will not react with the surface of the NiNPs, unlike any halide or carbonyl precursor.
Furthermore, this complex is analogous to Ru(COD)(COT), facilitating the comparison of
results.
The aim of this section is to (1) determine whether, as for RuNPs, the size control of
NiNPs generated in situ in ILs may be achieved through use of the intrinsic 3-D structure, and
(2) determine the stabilisation factors of the resultant NiNPs.
Figure 12. Typical synthesis of PVP-stabilised NiNPs dispersed in organic solvent (Chaudret’s method). TEM
image of PVP-stabilised NiNPs dispersed in MeOH and subsequent size distribution histogram.53
73
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Ni(COD)2, when decomposed under hydrogen in organic solvents such as THF in the
presence of hexadecylamine (HDA) as the stabiliser ligand or polyvinylpyrrolidone (PVP) as
polymer support,53, 54 produces NiNPs of a face centred cubic structure with a size distribution
centred around 4 – 5 nm, as found by Chaudret’s group. Cyclooctane is produced as a side
product. These NPs may be removed from the media by evaporation of the solvent and
homogenised by redispersion into other solvents, such as methanol or pentane, as seen in Figure
12. The NiNPs display a magnetisation value comparable to that of bulk Ni, although the
magnetic behaviour may be altered by adding more coordinating substrates to the medium, e.g.
CO.53, 54
Clearly research into NiNPs is well advanced, however the challenge remains to generate
sizes of less than 4 nm, which have not yet been successfully prepared from organometallic
tel-00708600, version 1 - 15 Jun 2012
precursors. To our knowledge the smallest recorded NiNPs are 2.5 nm, prepared by means of
laser electrodispersion techniques.58
H2 (4bar), 75 °C
Ni NiNPs
C1CnIm NTf2
Benzene
Figure 13 TEM micrographs and respective histograms showing the size-distribution of Ni nanoparticles prepared
and dispersed in [C1C4Im][NTf2] [a], [C1C8Im][NTf2] [b], [C1C10Im][NTf2] [c].56, 59
74
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Dupont et al. have already reported the preparation of NiNPs from Ni(COD)2 in
imidazolium based ILs in order to control the size of the NiNPs by varying the length of the
imidazolium alkyl substituent.56, 59
Here, the size of the NPs formed was found not to vary
greatly on changing the nature of the ionic liquid employed, i.e the alkyl side chain length, and in
contrast to previous findings from our group for RuNPs,26 the average size of NiNPs was found
to vary inversely with alkyl chain length, Figure 13. However, it is worth noting that a
substantial amount of benzene was used to aid the dissolution of Ni(COD)2 (6 mL of benzene for
2 mL of IL) and although effort was made to remove the volatiles before decomposition, we
cannot be sure of the complete removal of benzene and therefore the media in which the
decomposition actually occurred. Furthermore, the Ni-NiO core-shell particles produced bring
into question the purity of the solvents used. Therefore, the results are not comparable with the
tel-00708600, version 1 - 15 Jun 2012
work in our laboratory, where pure neat ILs are used as the solvent.
2.3.1 Size-controlled synthesis of NiNPs
The synthesis of nickel nanoparticles, NiNPs, from the decomposition of bis-(1,5-
cyclooctadiene) nickel (Ni(COD)2), under molecular hydrogen (4 bars) at 0 °C or 25 °C, was
attempted firstly in the following imidazolium based ionic liquids (ILs); [C1CnIm][NTf2], with
n = 2, 4, 6, 8, 10, varying the length of the alkyl chain in order to investigate the effect on the
size of the resulting NiNPs.
2.3.1.1 Decomposition in [C1C6Im][NTf2], [C1C8Im][NTf2] and [C1C10Im][NTf2]
Firstly, Ni(COD)2 (50 mg, 0.14 mmol) was stirred into the ionic liquid (10 mL) until a
light yellow solution resulted (14 mmol L-1 in each case). For all three ILs, the solutions were
exposed to H2 (4 bar) at 0 °C or 25 °C producing after 3 days black suspensions of NiNPs,
Scheme 6, except in the case of [C1C6Im][NTf2] at 0 °C, where the reaction proceeded very
slowly, and the starting product was still present with the NPs. The resulting volatiles were
removed in vacuo, trapped and characterised by 1H and 13
C NMR spectroscopy. The spectra
obtained showed the predominance of cyclooctane, with a small amount of remnant 1,5-COD
and even a trace amount of its isomer 1,3-COD. We may conclude that in this case the
decomposition occurs mainly via the hydrogenation of 1,5-COD, and the small amount
unchanged ligand may undergo isomerisation in the presence of Ni-H.
75
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
H2 (4bar), 25 °C
NiNPs +
Ni
C1C4<nIm NTf2
70
60
60
50 C8C1Im NTf2 0C
C8C1Im NTf2 RT 50
40
C10C1Im NTf2 0C
frequency
frequency
40
C10C1Im NTf2 RT
30
30
20
20
10 10
0 0
0 2 4 6 8 10 12 2 4 6 8 10 12
size / nm size/nm
Figure 14. Size distribution histrograms of NiNPs synthesized in [C1C8Im][NTf2] (left),and [C1C10Im][NTf2](right),
under 4 bars of H2 in various at 0 °C and at 25 °C
76
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
70
70
60
60
50
C6C1Im NTf2
50 C8C1Im NTf2 0C C8C1Im NTf2
C10C1Im NTf2 0C C10C1Im NTf2
frequency
40
frequency
40
30
30
20
20
10 10
0
tel-00708600, version 1 - 15 Jun 2012
0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
size/ nm size/ nm
Figure 15. Size distribution histograms of NiNPs size synthesised under 4 bars of H2, at 0° and 25°C without
stirring in different ILs.
At room temperature, the observed mean diameter increases with the side alkyl chain
length: in [C1C6Im][NTf2], (5 ± 2 nm), in [C1C8Im][NTf2] (6 ± 2 nm). and in [C1C10Im][NTf2],
(7 ± 2 nm). It was noted that a "mirror" appeared on the bottom of the glass auto-clave during the
reaction in [C1C10Im][NTf2] at 25°C corresponding to the deposition of bulk metallic nickel.
2.3.1.4 Comparison with RuNPs
The synthesis of NiNPs under H2 in these ILs, was seen to give similar trends to RuNPs,
i.e. mean size of resultant NPs increasing with both increasing temperature and increasing alkyl
chain length. In the case of RuNPs, when n = 10 we observe the formation of large “sponge-like
superstructures” with a size of 100–150 nm containing small aggregated particles. This
phenomenon could result from a diffusive process of RuNPs between the different non-polar
domains due to the chain length. For NiNPs, this phenomenon is not observed, although at 25 °C
in the same IL, the deposition of a metallic mirror as well as the formation of NiNPs was
recorded.
Unlike RuNPs generated in situ the sizes of NiNP obtained were large, and with a
relatively broad size distribution. Indeed, the NiNPs obtained are too large to be correlated to the
measured size of the non-polar domains. The fact that they cannot be contained within these
77
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
domains may explain their gradual agglomeration and precipitation, observed in each case
between 5 and 10 days after formation, and storage under inert atmosphere.
These differences between RuNPs and NiNPs could be related to a different mechanism
involved in generation. The complex Ru(COD)(COT) reacts readily with H2.63 Consequently, its
decomposition under H2 rapidly affords nuclides which subsequently coalescence to generate
small RuNPs of controlled size.
On the other hand, a higher energy barrier is involved in its decomposition of Ni(COD)2.
Consequently, the formation of nuclides is slow. Few nickel nuclides therefore coexist with a
great deal of Ni(COD)2. The nuclides induce the decomposition of Ni(COD)2 meaning that the
growth process in this case is much faster than the nucleation and therefore fewer larger NPs are
formed with less control, as illustrated in Scheme 7.
tel-00708600, version 1 - 15 Jun 2012
Similarly to RuNPs, when a solution of freshly prepared NiNPs was treated under
ethylene, in the absence of H2, ethane was detected in the atmosphere, indicating the presence of
hydrides on the nickel surface.30, 65 However, unlike RuNPs when heated and stirred under an
atmosphere of H2, NiNPs undergo rapid coalescence and precipitation. Recently, it has been
reported that the size of NiNPs formed by “dewetting” of Ni metal film drastically increases with
increasing H2 pressure. By analogy with Raney nickel, this result has been related by the authors
to the formation of a surface hydride leading to a modification of the surface free energy.66
2.3.2 Spontaneous decomposition on solvation
In the absence of hydrogen, in neat [C1C2Im][NTf2] and [C1C4Im][NTf2] at room
temperature, and in [C1C6Im][NTf2] at 40 °C, spontaneous decomposition of Ni(COD)2 occurred
on dissolution. TEM was performed on the resulting black solutions and revealed in each case a
mixture of NiNPs (< 10 nm) and sponge-like agglomerates of larger particles, Figure 16. It can
be seen from the TEM image that these sponge-like structures clearly consist of individual
78
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
particles. This is completely different from Dupont’s work,56 emphasising the important role that
a co-solvent such as benzene plays in the medium.
tel-00708600, version 1 - 15 Jun 2012
Note that 1H NMR spectroscopy and gas chromatography performed on the resulting
solution showed only a trace of cyclooctane COA, resulting from the hydrogenation 1,5-COD,
the by-product of decomposition reported in the literature.53 Instead, the presence of both 1,5-
cyclooctadiene and its isomer 1,3-cyclooctadiene (1,3-COD) was detected by both gas
chromatography and 1H and 13
C NMR spectroscopy. 1,3-COD must be a result of 1,5-COD
isomerisation, which could only take place at a metal centre in the presence of [Ni]-H bond,
Scheme 8.
79
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Ni H Ni H
Ni H
Ni Ni
Ni H
With RuNPs, the presence of surface hydrides has been proven by hydrogenation of
tel-00708600, version 1 - 15 Jun 2012
ethylene without the addition of H2 by a solution of RuNPs in IL, as shown in Scheme 9.67 We
repeated this experiment for the NiNP solutions formed by autodecomposition of Ni(COD)2 in
the media.
Scheme 9. Hydrogenation of ethylene over RuNPs without addition of H2, - evidence for surface hydrides.
After treatment under ethylene atmosphere (4 bars, 100 °C, 24 h) no ethane was detected
by GC as a result of ethylene hydrogenation, however significant amounts of butenes and
hexenes were detected, probably a result of oligomerisation of the ethylene.
The formation of an ethyl substituted IL [C1C2C4Im][NTf2] was also observed by NMR
spectroscopies; 1H shown in Figure 17, COSY, HETCOR and DOSY; and confirmed by
electrospray mass spectrometry, where cations at m/z = 139 and 167 were observed with similar
abundances, corresponding to [C1C4Im]+ and [C1C2C4Im]+, respectively. Note that a small
amount of the cation [C1C4C4Im]+ was also observed by mass spectrometry, m/z = 195, probably
a result of oligomerisation of ethylene to 1-butene before reaction with the imidazolium cycle.
80
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
C12
C11
C5' C4'
C10’-H C9-H
C6’-H C8-H
C11-H
C9’-H
C2-H C8’-H
C4-H C5-H C7-H
C12-H C7’-H
C4’-H C5’-H C6-H
C2 C6 C8
C10
N N C7 C9 C10-H
C2-H C9-H
C4-H C6-H C7-H C8-H
C5-H C5 C4
tel-00708600, version 1 - 15 Jun 2012
9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 ppm
Figure 17. 1H NMR spectrum of [C1C4Im][NTf2] containing NiNPs from auto-decomposition, before (lower) and
after (upper) treatment under ethylene.
The evidence gathered suggests that the observed decomposition could be due to the
cleavage of the very acidic C2-H bond and the consequent in situ generation of a nitrogen
heterocyclic carbene (NHC), as already reported in the case of Ir NP preparation in ILs3, 5, 13 and
with homogeneous complexes of Ni68, Pd,69, 70 and of Rh and Ir.71
Cavell et al. have proposed a scheme for the possible catalytic cycle for the
imidazolium/alkene coupling reaction where the organometallic starting material they studied
was Ni(OAc)2.72 The same phenomenon has also been observed in work by Lecocq et al. who
were investigating the oligomerisation behaviour of Ni in imidazolium ionic liquids.68
At this point it is impossible to determine whether the observed reactions (isomerisation
of 1,5-COD and formation of C1C2C4ImNTf2) occur on molecular or colloidal species. However,
these results do prove that the cleavage of the C2-H bond does occur during the spontaneous
decomposition on dissolution. The proposed mechanism of the reaction of imidazolium salts
with low valent M0 (M = Pd and Ni) hypothesises the formation of a molecular carbene-M-H
species,73, 74
as also proposed in the catalytic cycle for oligomerisation and formation of
trialkylimidazolium species in Scheme 10.
81
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Ni(COD)2
NiNP IL
COD Bu
N N
Ni(0)(COD) NTf2
R
Bu
N N
NTf2
R = Et, Bu
N N H
II II
Ni (COD) Ni (COD)
tel-00708600, version 1 - 15 Jun 2012
N N
NTf2 NTf2
Bu Bu
N
NiII(COD)
N
NTf2
Bu
Scheme 10. The proposed catalytic cycle for oligomerisation of ethylene and formation
of trialkylimidazolium species.
82
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
chains, involves attack on the acidic C2-H. This may be inhibited in the longer chain ionic liquids
due to the greater distance between the imidazolium ring and the non-polar domains, where
Ni(COD)2 will dissolve preferentially. If we are to achieve the generation of small size-
controlled NiNPs, we must find a way of inhibiting this auto-decomposition in imidazolium ILs
with short chains. Several strategies were attempted in order to circumvent the problem.
Firstly, [C1C1C4Im][NTf2] which does not contain the most acidic C2–H proton was used
as the solvent. Surprisingly, the Ni(COD)2 still decomposed on stirring, but this time afforded
well dispersed NiNPs (7 ± 2 nm), Figure 18. This can only be explained by attack on the two less
acidic protons C4-H,C5-H of the imidazolium ring and generation of transitory non-classical
NHC ligands.75
tel-00708600, version 1 - 15 Jun 2012
30
25
C4C1C1Im NTf2
20
7 ± 2 nm
frequency
15
10
0
2 4 6 8 10 12
size/ nm
Figure 18. TEM image of NiNPs formed by autodecomposition of Ni(COD) 2 in [C1C1C4Im][NTf2] and size
distribution histogram.
83
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
40
30 BF4 RT
Frequency
20
10
0
1 2 3 4 5 6 7
Size / nm
Figure 19. TEM image of NiNPs formed by decomposition of Ni(COD)2 under H2 (4 bars at 25 °C) in
[C1C4Im][BF4] and size distribution histogram.
tel-00708600, version 1 - 15 Jun 2012
2.3.4 Conclusion
The synthesis of NiNPs from Ni(COD)2 under H2 (4 bar) has been carried out in various
imidazolium based ionic liquids at 0 °C and 25 °C. Imidazolium cations substituted with longer
chains [C1C6Im][NTf2], [C1C8Im][NTf2] and [C1C10Im][NTf2], the IL media were found to offer
fairly good media for the preparation and stabilisation of nickel nanoparticles. NiNP size
increased with increasing chain-length and in this way we were able to control NiNP size.
Additionally, upon decreasing the reaction temperature to 0 °C, we were able to produce smaller
NiNPs than those produced in the same IL media at 25 °C. The presence of surface hydrides on
on these NiNPs as on RuNPs has been evidenced by hydrogenation of ethylene in the absence of
hydrogen. The resulting NiNPs in these ILs were found to be larger than RuNPs produced in the
same media. This can be largely attributed to differences in the nucleation and growth processes
for the two metals.
Unexpectedly, spontaneous decomposition of Ni(COD)2 occurred without the addition of
hydrogen for imidazolium ILs with short alkyl chains; [C1C2Im][NTf2]. [C1C4Im][NTf2] and
[C1C1C4Im][NTf2]. In [C1C2Im][NTf2] and [C1C4Im][NTf2], TEM micrographs showed NiNPs
of fairly large diameter were formed as well as sponge-like super agglomerates. In
[C1C1C4Im][NTf2] well dispersed NiNPs are formed. In these cases, an explanation concerning
the activation of the acidic protons on the imidazolium ring and the consequent nitrogen
heterocyclic carbene formation leading to rapid decomposition of the complex, has been
proposed, Scheme 11.
84
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Bu + Bu 1,3-COD
+ 2 (1,5-COD)
N N
N
A- N
Bu Me
C2
A- + Ni(1,5-COD)2 C2 [Ni] H N
H
Me
N
H 1,5-COD
Me
C2H4
?
Bu N + N N + N
Bu Me Bu Me
N C2 C2
+ H A- A-
C Et Bu
N 2 H
Me = Ni
Scheme 11. Reactions occuring during the autodecomposition of Ni(COD)2 in imidazolium ionic liquids.
tel-00708600, version 1 - 15 Jun 2012
85
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Note that of the early transition metals, only chromium, molybdenum and tungsten NPs
in ILs are reported in literature. They are obtained by by thermal or photolytic decomposition
under argon of mononuclear metal carbonyl precursors M(CO)6 (M = Cr, Mo, W).82
The number of publications describing the synthesis of Ta(0)NPs, is limited. Due to high
melting and boiling points, Ta(0)NPs are difficult to prepare by conventional physical methods.
For example, the preparation of Ta(0)NPs by a hydrogen rich plasma (water used as the
plasmatic medium as it dissociates under the effect of an electric arc) led to a mixture of
Ta(0)NPs and TaO NPs.83 Studies by XRD indicate that the TaO NPs present a hexagonal phase
with interplanar distances of 0.31 and 0.39 nm depending on the face under study (100 and 001
respectively). The Ta(0)NPs however, have a face centred cubic structure much like bulk Ta
metal. Their average size was estimated to be 10 nm and interplanar distances of the (110) face
were measured as being 0.24 nm.84
Chemical methods are rarely used. They involve the reduction of TaCl5 or K2TaF7 at high
temperature or in liquid ammonia in the presence of sodium.85, 86 Such techniques result in Ta
particles of sizes ranging between 1 – 100 μm. More recently, Ta(0)NPs supported on silica have
been described. The reaction of pentabenzyl tantalum, Ta(CH2C6H5)5, with the silica surface,
results in the formation of surface complex(es) that under hydrogen atmosphere at 573 K and
723 K, lead to the formation of Ta(0)NPs on SiO2 Scheme 12.87 Such Ta(0)NPs, of mean size
1 nm, contain Ta-Ta bond lengths of 2.93 Å (determined by EXAFS) and Ta-O-Si bond
lengths of 1.9 Å. The latter would suggest that the Ta(0)NPs are chemically bonded to the
86
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
surface. The question of the oxidation state of these surface Ta(0)NPs remains unanswered,
having not yet obtained XPS data.
H2
Ta(CH2Ph)5 + Si O2 ( SiO)nTa(CH3Ph)5-n [SiO2][Ta(0)]
nPhCH3 alkanes,
aromatics
Scheme 12. The synthesis of silica supported TaNPs87
Although ILs are known to be good media for the synthesis and stabilisation of NPs, few
papers regarding oxophilic metals are available.82 To our knowledge, no Ta(0)NPs have yet been
synthesised in ionic liquid media, although several examples of the electroreduction of tantalum
halides are reported leading to electroplated Ta on Pt electrodes, without reduction of the ionic
liquids.88-90 Here, a thin layer of halides is present at the surface, outlining the consequences and
tel-00708600, version 1 - 15 Jun 2012
87
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
However, the NPs were not monodisperse as larger species of up to ~ 18 nm were also detected
(c.a. 30 % of NPs larger than 10 nm).
35
30
25
Frequency
20
15
10
0
tel-00708600, version 1 - 15 Jun 2012
0 2 4 6 8 10 12 14 16 18
Size/ nm
88
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
50
45
40
with toluene 36%
35
with pentane 2%
30 no co-solvent
Frequency 25
20
15
10
0
0 5 10 15
Size/ nm
tel-00708600, version 1 - 15 Jun 2012
Figure 21. Superimposed size distribution histograms for TaNPs produced in [C1C4Im][NTf2] neat and in the
presence of co-solvents, pentane (2 % wt) and toluene (36 % wt).
Table 1. Resulting sizes of TaNPs, produced in different media under H 2 (4 bars) and 25 °C
Ionic Liquid Co-solvent Mean size NPs (nm)
C1C4Im NTf2 none 5 ± 3 (70%) (<10 (30%))
C1C4Im NTf2 toluene (36% wt) 5±3
C1C4Im NTf2 n-pentane (2% wt) 5±1
89
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
80
60
Frequency
C1C6Im NTf2
C1C4Im NTf2
40
20
0
0 2 4 6 8 10 12 14 16 18
Size/ nm
Figure 22. Comparative size distribution histograms of TaNPs produced in [C1C4Im][NTf2] and [C1C6Im][NTf2]
and TEM image of those formed in the latter.
90
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
atmosphere to remain zero-valent. The NP suspensions produced were not stable and decanted
over time. This presents the possibility of facile extraction of the NPs produced for use in other
media.
tel-00708600, version 1 - 15 Jun 2012
Figure 23. HREM images of crystalline TaNPs produced in [C1C4Im][NTf2]. Interplanar distances indicated by the
arrows are of 2.2 Å.
2.4.5 Conclusion
In conclusion, we have demonstrated that small monodisperse and zero-valent tantalum
nanoparticles, Ta(0)NPs, may be generated under mild conditions, by decomposition of the
organometallic complex tris(neopentyl)neopentylidene tantalum(V), under H2 in imidazolium
derived ionic liquids. The control of the size and the factors of stabilisation of these NPs in ionic
liquids are currently under investigation. This method has opened up the possibility of facile
access to other oxophilic metal nanoparticles, which we are currently studying, and could be of
great potential interest in the drive for smaller and more efficient electronic devices.
91
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
48 h at 65 °C. The hot solution was then transferred dropwise via a canula into toluene (200mL)
at 0 °C under vigorous mechanical stirring. The white precipitate formed was then filtered and
washed repeatedly with toluene (3 × 200 mL) and dried over night in vacuo giving a white
powder (95.6 g, 87 %). 1H-NMR (CD2Cl2): δ (ppm) : 11.05 (s, 1H, C2H) ; 7.33 (d, 1H, C4H) ;
7.28 (d, 1H, C5H) ; 4.31 (t, 2H, NCH2) ; 4.07 (s, 3H, NCH3) ; 1.90 (qt, 2H, CH2CH2CH2) ; 1.41
(st, 2H, CH2CH2CH3) ; 0.96 (t, 3H, CH2CH3) ; 13C{1H}-NMR (CD2Cl2) : δ (ppm) : 138.3 (C2H) ;
122.3 (C4H) ; 119.8 (C5H) ; 50.1 (NCH2) ; 36.8 (NCH3) ; 32.5 (CH2CH2CH2) ; 19.8
(CH2CH2CH3) ; 13.6 (CH2CH3)
1-butyl-3-methylimidazolium bis(trifluoromethylsulphonyl)imide [C1C4Im][NTf2] A
solution of lithium bis(trifluorosulphonylimide) [LiNTf2] (50 g, 0.17 mol) in water (50 mL) was
added to a solution of BMICl (30.4 g, 0.17 mol) in water (100 mL). The solution was stirred for
2 h at room temperature, then dichloromethane (50 mL) was added and the mixture transferred to
a separating funnel. The lower phase was collected and washed repeatedly with water
(8 × 100 mL) until no chloride traces remained in the washings (tested with silver nitrate
solution). The ionic liquid in dichloromethane was purified through a short alumina column and
the solvent removed in vacuo giving a colourless viscous liquid. 1H-NMR (CD2Cl2): δ (ppm) :
8.73 (s, 1H, C2H) ; 7.28 (d, 1H, C4H) ; 7.24 (d, 1H, C5H) ; 4.15 (t, 2H, NCH2) ; 3.90 (s, 3H,
NCH3) ; 1.83 (qt, 2H, CH2CH2CH2) ; 1.32 (st, 2H, CH2CH2CH3) ; 0.94 (t, 3H, CH2CH3) ;
13
C{1H}-NMR (CD2Cl2) : δ (ppm) : 134.3 (C2H) ; 124.3 (C4H) ; 122.4 (C5H) ; 118.2 (CF3) ; 50.3
(NCH2) ; 37.1 (NCH3) ; 33.2 (CH2CH2CH2) ; 19.8 (CH2CH2CH3) ; 13.5 (CH2CH3)
92
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
1-Methylimidazole (> 99 %) was purchased from Aldrich and distilled prior to use.
Chlorobutane, benzyl chloride and 4-chlorobut-1-ene (> 99 %, Aldrich) were used without
further purification. Bis(trifluoromethanesuflonyl)imide lithium salt (> 99 %, Solvionic) was
used without further purification.
Organometallic complexes
( 4-1,5-cyclooctadiene)(
)( 6-1,3,5-cyclooctatriene)ruthenium(0) Ru(COD)(COT) was
95
synthesised as reported. Degassed cycloocta-1,5-diene (60 mL, 0.49 mol) filtered over alumina
and zinc dust (10.0 g) were added to a solution of RuC13.3H2O (3.6 g, 13.7 mmol) in distilled
degassed methanol (50 mL). The mixture was heated under reflux with magnetic stirring for 6 h
and then filtered, washing the solid residue with dry degassed toluene. The resulting brown
solution was evaporated in vacuo until the residue was completely dry. Extraction with n-pentane
tel-00708600, version 1 - 15 Jun 2012
(2 mL) gave a red solution which was filtered through a column of alumina (20 cm), resulting in
a yellow solution. The volume of the solvent was reduced to ca. 5 mL in vacuo and the yellow
solution was cooled to - 78 °C to give, in 6 h, yellow crystals of Ru(COD)(COT) (1.7 g, 43 %),
m.p. 92 - 94 °C. 1H-NMR (C6D6) δ (ppm) : 4.78 (dd, 2 H), 5.22 (m, 2 H), 6.21 (m, 2 H), 7.08 (m,
4 H), 7.78 (m, 8 H), 8.36 (m, 2 H), 9.10 (m, 2 H).
s( 4-1,5-cyclooctadiene)nickel(0), Ni(COD)2, (Strem > 99 %) was used as recieved.
Bis(
Tris(neopentyl)(neopentylidene)tantalum(V), Ta Schrock, was prepared according to the
literature procedure. 93 A 1.56 M solution of NpMgCl in diethylether (100 mL) was added drop-
wise over 0.5 h to a stirred suspension of TaCl5 (6.0 g, 16 mmol) in 150 mL of diethylether at
room temperature. Any solid TaC15 rapidly dissolved as the reaction proceeded through stages
characterized by greenish-yellow, yellow, and finally, orange-brown colours. The final solution
contained some magnesium chloride, which was removed by filtration over celite. All solvent
was removed in vacuo and the residue sublimed in two stages at 90 °C and ~ 10-8 bar to give
1.7 g (22 %) of deep orange crystals on the probe. 1H-NMR (C6D6) δ (ppm): 8.09 (s, 1H) , 8.57
(s, 9H), 8.85 (s, 27H), 9.16 (s, 6H)
Nanoparticles
Ruthenium In a typical synthesis, Ru(COD)(COT) (66 mg, 0.21 mmol) was stirred
virorously into the dry degassed ionic liquid (5 mL). The resulting bright yellow suspension was
then transferred via canula to a Fischer-Porter bottle, in a thermostatic bath (0° C or 25° C),
which was pressurised under pure H2 or D2 (4 bars) over 3 days (0° C) or 18 h (25° C). The
93
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
volatiles (cyclooctane) were subsequently removed under reduced pressure and the resultant
black solution was stored under inert (Ar) atmosphere in the glovebox.
Nickel – decomposition under H2.The same procedure as above was applied to Ni(COD)2
(50 mg, 0.14 mmol) in ionic liquid [C1Cn>4Im][NTf2] (10 mL).
Nickel – decomposition on dissolution. Ni(COD)2 (50 mg, 0.14 mmol) was stirred into
ionic liquid [C1Cn≤4Im][NTf2] (10 mL) producing a pale yellow solution at 25 °C. After 1 h the
solution had turned green and after a further hour the solution a black solution resulted.
Tantalum. Ta Schrock (25 mg, 5.4×10-5 mol) was dissolved in the ionic liquid (10 mL) or
ionic liquid-organic solvent mixture (dry, degassed toluene or pentane). The resulting orange
solution was canulated into a Fischer-Porter bottle, which was pressurised under 4 bars of H2 at
25 °C during 4 days. A black solution resulted, which was stored in a glove box under argon.
tel-00708600, version 1 - 15 Jun 2012
[C1C4//Im][NTf2]:
94
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
137.1
100
90
80
70
60
50
40
30
138.1
tel-00708600, version 1 - 15 Jun 2012
20
141.1
140.1
10
139.1 142.1
143.1 144.2 145.2 146.2 147.2
0
138 139 140 141 142 143 144 145 146 147
m/z
[C1BzIm][NTf2]:
95
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
90
80
70
60
50
40
174.1
30
20
10 175.1
m/z
175.2
90
80 186.2
184.2
70
60
183.2
50 187.2
181.2
182.2
40
188.2
30
180.2
176.1
20 189.2
179.2
190.2
10 178.0 191.2 193.2 193.8 194.4
177.1
0
176 178 180 182 184 186 188 190 192 194
m/z
96
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
orientation.
2.5.5 GC analyses
The products were quantitatively analysed by gas chromatography on a P6890
chromatograph equipped with a flame ionisation detector (FID) and an Al2O3/KCl column (L )
50 m, φint ) 0.32 mm, film thickness ) 5 μm). The injector and detector temperature was 230° C,
and the injection volume was 1 μL. The temperature was fixed at 190° C.
2.5.6 Hydrogenation of Ethylene by RuNPs in the IL
A 2 mL sample of the RuNP solution in the IL was introduced under argon into a Schlenk
of known volume. The sample was treated under flow of argon for 18 h. The argon atmosphere
was replaced with a known pressure of ethylene using a vacuum line and the system was stirred.
A decrease of the internal pressure was observed, and the composition of the gas phase was
monitored by gas chromatography. After 12 h, the atmosphere was replaced by a H2 atmosphere
and the system was heated and stirred for 12 h. The composition of the gas phase was again
monitored by gas chromatography.
Ethylene Hydrogenation Calculations
pV = nRT
Vreactor = 125.8 cm3 = 1.258 × 10-4 m3
pr = 129 mbar = 1.2 × 104 Pa
R = 8.314 m3 Pa K-1 mol-1
T = 298 K
97
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
n = (1.29 ×104 ×1.258 ×10-4) / (8.314 × 298) m3 Pa m-3 Pa-1 K mol K-1 =
6.54 × 10-5 mol ethylene introduced
GC calibration results: 2.3×10-12 × A = no. moles of C
300µL injections = 3×10-7 m3
Table 2. GC results following treatment under ethylene
Species Ret time/ mins Peak Area No. Moles of C No. of Moles (in No. of Moles (in
300 µL) reactor)
C2H4 + H2 à C2H6:
1 mole ethane produced = 1 mole H2 consumed, i.e. 2 moles of surface hydrides
2C2H4 + H2 à C4H10:
1 mole isobutane produced = 1 mole H2 consumed, i.e. 2 moles of surface hydrides
2C2H4 à C2H8
1 mole butene produced = No H2 consumed
Table 3. GC results following treatment under H2 liberating surface alkyls
Species Ret time/ mins Peak Area No. Moles of C No. of Moles (in No. of Moles (in
300 µL) reactor)
½ C2H4 + ½ H2 à -CH3
1 mole methyl produced = ½ mole H2 consumed, i.e. 1 mole of surface hydrides
C2H4 + ½ H2 à -C2H5
1 mole ethyl produced = ½ mole H2 consumed, i.e. 1 mole of surface hydrides
1½ C2H4 + ½ H2 à -C3H7
98
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
6.54 ×10-5 mol (introduced) - 6.02×10-5 mol (detected by GC) = 5.2 µmol ~ 8 % error
2.5.7 H-D exchange
A 2mL sample of the RuNP solution in the IL was introduced under argon into a Schlenk,
which was then treated under dynamic vacuum (10-5 mbar) during a period of 1h or 4h. This was
tel-00708600, version 1 - 15 Jun 2012
then charged with deuterium (1 bar) and stirred at 25°C during a period of 1 or 4 days. The gas
phase was subsequently expanded into an airtight Young tube for NMR analysis.
2.5.8 TON calculations
Butenyl IL à butyl IL
Ru(COT)(COD) – 2.2 × 10-4 moles
RuNPs, 1.1 ± 0.2 nm, dispersion ~75%
Rus – 1.7 × 10-4 moles
Butenyl IL 7.5g/ 417g mol-1 = 1.8 × 10-2 moles
15 % converted à 2.7 × 10-3 moles
TON à 2.7 × 10-3 moles/ 1.7 × 10-4 moles = 16
Benzyl IL à methylenecyclohexyl IL
Ru(COT)(COD) – 2.2 × 10-4 moles
RuNPs, 3.2 ± 0.7 nm, dispersion ~35%
Rus – 7.7 × 10-5 moles
Benzyl IL 7.5g/ 453g mol-1 = 1.7 × 10-2 moles
30% converted à 4.9 × 10-3 moles
TON à 4.9 × 10-3 moles/7.7 × 10-5 moles = 64
99
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
2.6 References
100
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
101
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
38. M. Leconte, A. Theolier, D. Rojas and J. M. Basset, J. Am. Chem. Soc., 1984, 106, 1141-
1142.
39. I. M. Ciobica, A. W. Kleyn and R. A. Van Santen, J. Phys. Chem. B 2003, 107, 164-172.
40. T. H. Rod, A. Logadottir and J. K. Norskov, J. Chem. Phys., 2000, 112, 5343–5347.
41. J. Wang, C. Y. Fan, Q. Sun, K. Reuter, K. Jacobi, M. Scheffler and G. Ertl, Angew.
Chem. Int. Ed., 2003, 42, 2151-2154.
42. M. P. Stracke, M. V. Migliorini, E. Lissner, H. S. Schrekker, D. Back, E. S. Lang, J.
Dupont and R. S. Goncalves, New J. Chem., 2009, 33, 82-87.
43. A. Roucoux, A. Nowicki and K. Philippot, in Nanoparticles, ed. D. Astruc, Wiley-VCH,
Weinheim, 2007, p. 351.
44. L. Cammarata, S. G. Kazarian, P. A. Salter and T. Welton, Phys. Chem. Chem. Phys.,
tel-00708600, version 1 - 15 Jun 2012
2001, 3, 5192-5200.
45. Y.-Y. Jiang, Z. Zhou, Z. Jiao, L. Li, Y.-T. Wu and Z.-B. Zhang, J. Phys. Chem. B, 2007,
111, 5058-5061.
46. A. Mele, C. D. Tran and S. H. De Paoli Lacerda, Angew. Chem. Int. Ed., 2003, 42, 4364-
4366.
47. A.-L. Rollet, P. Porion, M. Vaultier, I. Billard, M. Deschamps, C. Bessada and L.
Jouvensal, J. Phys. Chem. B, , 2007, 111 11888 -11891.
48. B. L. Bhargava and M. L. Klein, J. Phys. Chem. B, 2009, 113, 9499-9505.
49. B. L. Bhargava and M. L. Klein, Soft Matter, 2009, 5, 3475-3480.
50. S. Pickering, J. Chem. Soc., 1907, 91, 307.
51. L. Lu, M. L. Sui and K. Lu, Science, 2000, 287, 1436.
52. H. Wang, X. Jiao and D. Chen, J. Phys. Chem. C, 2008, 112, 18793-18797.
53. N. Cordente, C. Amiens, B. Chaudret, M. Respaud, F. Senocq and M.-J. Casanove, J.
Appl. Phys. , 2003, 94, 6358-6365.
54. N. Cordente, M. Respaud, F. Senocq, M.-J. Casanove, C. Amiens and B. Chaudret, Nano
Lett., 2001, 1, 565-568.
55. T. Ould Ely, C. Amiens and B. Chaudret, Chem. Mater., 1999, 11, 526-529.
56. P. Migowski, G. Machado, S. R. Texeira, M. C. M. Alves, J. Morais, A. Traverse and J.
Dupont, Phys. Chem. Chem. Phys., 2007, 9, 4814-4821.
102
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
Svedlindh, M. Haase and H. Weller, J. Am. Chem. Soc., 2005, 125, 9092-9101.
65. E. A. Crespo, M. Ruda and S. Ramos de Debiaggi, Int. J. Hydrogen Energy, 2008, 33,
3561-3565.
66. A. Geissler, M. He, J.-M. Benoit and P. Petit, J. Phys. Chem. C, 2010, 114, 89-92.
67. P. S. Campbell, C. C. Santini, D. Bouchu, B. Fenet, K. Philippot, B. Chaudret, A. A. H.
Padua and Y. Chauvin, Phys. Chem. Chem. Phys., 2010, 12, 4217-4223.
68. V. Lecocq and H. Olivier-Bourbigou, Oil & Gas Sci. Tech, 2007, 62, 761-773.
69. L. Magna, Y. Chauvin, G. P. Niccolai and J.-M. Basset, Organometallics, 2003, 22,
4418-4425.
70. C. M. Crudden and D. P. Allen, Coord. Chem. Rev., 2004, 248, 2247-2273.
71. U. Hintermair, T. Gutel, A. M. Z. Slawin, D. J. Cole-Hamilton, C. C. Santini and Y.
Chauvin, J. Organometal. Chem., 2008, 693, 2407-2414.
72. D. S. McGuinness, W. Mueller, P. Wasserscheid, K. J. Cavell, B. W. Skelton, A. H.
White and U. Englert, Organometallics, 2002, 21, 175-181.
73. N. D. Clement, K. J. Cavell, C. Jones and C. J. Elsevier, Angew. Chem. Int. Ed., 2004, 43,
1277-1279.
74. D. C. Graham, K. J. Cavell and B. F. Yates, Dalton Trans., 2007, 41, 4650-4658.
75. S. Gruendemann, A. Kovacevic, M. Albrecht, J. W. Faller Robert and H. Crabtree, Chem.
Commun., 2001, 2274-2275.
76. P. Ireland, Thin Solid Films 1997, 304, 1-12.
103
Chapter 2 Generation and Stabilisation of Transition Metal NPs in Imidazolium ILs
77. S. Murarka and S. Hymes, Crit. Rev. Solid State Mater. Sci. 1995, 20, 87-124.
78. J. Dufourcq, S. Bodnar, G. Gay, D. Lafond, P. Mur, G. Molas, J. Nieto, L. Vandroux, L.
Jodin, F. Gustavo and T. Baron, App. Phys. Lett., 2008, 92.
79. C. C. Santini, J.-M. Basset, T. Gutel, P. Campbell, S. Deleonibus and P.-H. Haumesser,
CEA-LETI /Univ Lyon 1 -CNRS, Fr0901463, 2009.
80. C. C. Santini, J.-M. Basset, T. Gutel, P. Campbell, S. Deleonibus and P.-H. Haumesser,
CEA-LETI /Univ Lyon 1 -CNRS, Fr0901464, 2009.
81. G. Schmid, Nanoparticles: From Theory to Application, Wiley-VCH, Weinheim, 2004.
82. E. Redel, R. Thomann and C. Janiak, Chem. Commun., 2008, 1789-1791.
83. Y. Wang, Z. Cui and Z. Zhang, Mater. Lett., 2004, 58, 3017-3020.
84. A. F. Wells, Structural Inorganic Chemistry, Clarendon Press, Oxford, 1986.
tel-00708600, version 1 - 15 Jun 2012
104
III
tel-00708600, version 1 - 15 Jun 2012
Chapter 3
105
tel-00708600, version 1 - 15 Jun 2012
106
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Chapter 3
Imidazolium Ionic Liquids as Media for Homogeneous
Catalysis
3.2 The influence of the ionic liquid medium on the catalytic hydrogenation of 1,3-cyclohexadiene by
an Osborn-type catalyst............................................................................................................................. 116
tel-00708600, version 1 - 15 Jun 2012
3.2.1 Solvation of 1,3-cyclohexadiene and molecular structure of the mixtures .............................. 117
107
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
108
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
3.1 Introduction
Ionic liquids are up-and-coming solvents for catalysis, with the number of publications in
this field growing at an exponential rate.1-10,11,12 Indeed, the “green” nature of ILs is often cited,
and recyclability is a recurring theme to justify such great interest. This recyclability can be two-
fold, as ILs may act to immobilise certain ionic,13, 14or colloidal catalysts,15 preventing catalyst
leaching; and due to their negligible vapour pressure, they themselves may be recycled more
easily than traditional solvents, the products being recovered by simple distillation or liquid-
liquid extraction.16, 17 For the latter, the extraction solvent is often present during the reaction,
thus many papers describe “biphasic catalysis”.18-22
ILs of course are unlike traditional solvents. As reaction media, they have specific
tel-00708600, version 1 - 15 Jun 2012
properties that may have consequences on the catalytic process. In some catalytic reactions, the
difference between the use of ILs and traditional solvents can be related to the chemical role of
ILs, which may interfere with the reaction. For instance, imidazolium moieties of the IL may be
deprotonated in situ and coordinate to metal centres such as Pd, Ir and Rh as imidazolylidene N-
heterocyclic carbene ligands, as shown in Scheme 1.23, 24 Due to the acidic C2-H, this reaction
can occur under mild conditions and in the absence of a base.25 This has a significant effect on
catalytic reactions. In the case of palladium-catalysed telomerisation of butadiene with methanol,
catalyst deactivation occurs due to the formation of highly stable palladium imidazolylidene
complexes.26 During biphasic Ni-catalysed ethylene oligomerisation in ILs, the formation of
NHCs in situ has also been evidenced by the formation of an imidazolium cation substituted with
an ethyl group on C2, Scheme 2.27
109
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Scheme 2. Ethyl substituted ionic liquid resulting from the in situ formation of NHC
However, it has also been reported by Hintermair et al. that the nature of such ligands can
have positive effect in on both the thermal stability and selectivity of the resulting catalyst
compared to phosphine ligands.28 Thus the potential for chemical interference may be considered
either an advantage or disadvantage, depending on the desired outcome of the catalytic process.
tel-00708600, version 1 - 15 Jun 2012
It is well worth mentioning that certain ILs themselves exhibit catalytic activity. For
instance Lewis acidic ILs, particularly dialkylimidazolium chloroaluminates, which are amongst
the first described ionic liquids, can be used successfully as both solvent and catalyst
simultaneously. The acidity of such ionic liquids is easily tuned by simply altering the
stoichiometry of the dialkylimidazolium chloride versus the aluminium trichloride; an excess of
the latter leads to the presence of Al2Cl7-. Due both to this and to the fact that aromatics are
highly soluble in IL media, Friedel-Crafts alkylations and acylations occur readily in these dual-
purpose solvents, Scheme 3.6, 29
110
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
with varying chain lengths, have found structural heterogeneities that varied in size with the
alkyl chain length, Figure 1.
tel-00708600, version 1 - 15 Jun 2012
Figure 1. X-ray diffraction patterns from the series of supercooled liquid RTILs: [C 1CnIm]Cl, n ) 3, 4, 6, 8, 10, at
25 °C. In the inset the spatial correlation L ) 2π/QMAX, QMAX being the interference peak position, is plotted (circles)
as a function of n, the alkyl chain length,for 4 ≤ n. The corresponding data are also reported for the cases n = 6 and 8
for [C1CnIm][BF4] (squares). 31
Indeed ionic liquids, especially those based on imidazolium, are widely claimed to
exhibit a non-homogenous structure, with close anion-cation association existing throughout the
solution, as demonstrated through molecular simulation,35 and coarse grain modelling.36 The
order and structure therefore depends greatly on both the coordination strength of the anion and
on the alkyl chains which group together in lipophilic domains. Short alkyl chains (n < 4) are too
short to associate with one another, whereas intermediate length chains (4 ≤ n < 10) will form
“pockets”, as seen in Figure 2,35, 37 explaining the heterogeneities observed by X-ray diffraction.
Longer chains lead to mesophases and ionic liquid crystals as demonstrated through SAXS
experiments.38 This therefore leads to the question: how do these self-organised media interact
with different substrates and what is the upshot in catalysis?
111
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
tel-00708600, version 1 - 15 Jun 2012
Figure 2. Molecular simulations of [C1CnIm][PF6] ionic liquid structures. In a) [C1C2Im][PF6], colouring is used
to distinguish anions from cations, anions being yellow. In b)-f) colouring distinguishes charged species from non-
polar species, red being charged and green being non-polar: b) [C1C2Im][PF6], c) [C1C4Im][PF6], d) [C1C6Im][PF6],
e) [C1C8Im][PF6 and f) [C1C10Im][PF6]
Ionic liquids have been shown to exhibit specific interactions with themselves, so specific
interactions between the ionic liquid and the substrate may also exist and play an important role.
Indeed, molecular simulation and determination of radial distribution functions predict that non-
polar solutes dissolve in non-polar domains, polar solvents dissolve in polar regions, and
amphiphilic solvents, e.g. methanol and acetonitrile display more complicated behaviour.39
Polarity is not the only consideration however, as X-ray structural determination of benzene-
112
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
[C1C2Im][NTf2] mixtures with various compositions have shown evidence for strong π-cation
interactions, as shown in Figure 3.40 Similarly, through combined ROESY NMR and molecular
simulation studies, such interactions have been shown to exist in toluene-IL mixtures for the
trialkyl IL [C1C1C4Im][NTf2]. By contrast, for the dialkyl IL [C1C4Im][NTf2, π-cation
interactions with toluene are inhibited by the strong H-bond network between anion and cation,
induced by the relatively acidic C2-H. In this case, toluene penetrates the non-polar alkyl regions
of the IL.41 Another NMR study has provided supporting evidence that the nature of the IL, and
thus the interactions within, is crucial in determining the placement of thiiophene in thiophene-IL
mixtures. Figure 4 shows the different hypothesised arrangements for thiophene in the ILs
[C1C1Im][MeOPHO2] and [C1C4Im][SCN] based on this NMR evidence.42. Furthermore, the
arrangement of CS2 dissolved in the IL [C1C5Im][NTf2] has been studied as a function of
tel-00708600, version 1 - 15 Jun 2012
Figure 3. Crystal structure of benzene-[C1C2Im][NTf2] inclusion crystal for a 1:1 molar ratio
113
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
tel-00708600, version 1 - 15 Jun 2012
Figure 4. Proposed structure of thiophene in the ionic liquids [C1C1Im][MeOPHO2] (upper) and [C1C4Im][SCN]
(lower), 42
Figure 5. CS2 molecules are isolated from each other and mainly localised in the nonpolar domains of the IL.43
Finally, issues such as density and viscosity in ionic liquids, which are more significant
than in most traditional solvents, will play an important role in catalysis through dictating the
114
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
diffusion of molecules in the medium.44, 45, 46 The viscosity varies greatly with the nature of the
IL and, although related to the strength of the cation-anion association, is often difficult to
predict.47-48
In this chapter, we study how in ILs, the activity and selectivity of a model catalytic
reaction (hydrogenation of 1,3-cyclohexadiene, CYD, Scheme 4) are dependent on the solvation
of substrate and of the physical-chemical properties of the resultant catalytic medium.
H2
Cat / IL
Scheme 4. Catalytic hydrogenation of CYD
tel-00708600, version 1 - 15 Jun 2012
115
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
stacking existing for the latter due to disruption of the anion-cation H-bonding caused by the
methyl group. For a non-aromatic π-system, such as 1,3-cyclohexadiene, solvation is probably a
combination of more subtle interactions with the ionic liquid.
The ILs [C1C4Im][NTf2] and [C1C1C4Im][NTf2] are chosen in order to avoid impurities
such as chloride and water, since they are hydrophobic, liquid at room temperature, and their
purification is well controlled.26 Firstly, the molecular interactions between the ILs and CYD
must be assessed through:
NMR characterisation
study of the microscopic structure of the IL-substrate mixtures
measurement of the thermodynamic properties of mixing (such as the solubility
and the heat of mixing)
determination of the viscosity of the reaction media
measurement of diffusivity of CYD in ILs
Following this, the difference in catalytic performance in the two different media must be
quantified, and explained using the findings from the previously conducted analyses. (Anion
exchange between the IL and the catalyst can be circumvented by replacing Osborn’s complex
with [Rh(COD)(PPh3)2]NTf2, COD = 1,5-cyclooctadiene).
116
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Figure 6. Liquid-liquid equilibrium diagrams for the mixture of toluene and [C 1C4Im][NTf2] (o) or
[C1C1C4Im][NTf2 1C4Im][NTf2] ( ) or [C1C1C4Im][NTf2] ( ) xIL = IL molar
fraction: [IL]/([IL]+{CYD]).
117
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
equilibrium lines. In comparison with aromatic molecules in these ILs, CYD is much less
soluble. For instance, the solubility of toluene at 303.15 K expressed in mole fraction is 0.73 in
[C1C4Im][NTf2] and 0.72 in [C1C1C4Im][NTf2],41 the solubility of benzene in [C1C4Im][NTf2]
40
being even higher (mole fraction of 0.78 at 303.15 K). The small difference in solubility of
toluene with respect to benzene can be attributed to the presence of an additional methyl group,
whereas the much lower solubility of CYD is explained by less favourable interactions between
the solute and the ILs. Disruption of the aromatic system changes the interactions between solute
and ionic liquid and thus decreases the solubility.
3.2.1.2 Enthalpy of mixing
The energy involved in the interactions of different solutes dissolved in ILs can be
assessed by measuring the excess molar enthalpy. This was carried out for both systems; CYD-
tel-00708600, version 1 - 15 Jun 2012
[C1C4Im][NTf2] and CYD-[C1C1C4Im][NTf2]. The excess molar enthalpy of mixing ΔHmixE was
determined by isothermal titration calorimetry, from the heat effect involved in injections of
small quantities of CYD into the ionic liquid, QCYD. The partial molar excess enthalpy of solute,
HCYDE, was calculated according to equation (1).
E
E H mix QCYD
HCYD (1)
nCYD nCYD
nIL, p,T
where nCYD and nIL denote the quantity of CYD and IL, respectively, ΔHmixE is the excess molar
enthalpy of the entire system (enthalpy of mixing) and ΔnCYD is the quantity of solute per
injection. ΔnCYD was calculated from the injected volumes and the density of CYD was obtained
from the literature. In these calculations, heat due to evaporation of the solute from IL solution is
assumed to be negligible. Hence, no correction for the vapour pressure of the solute was made.
Figure 7 represents partial molar excess enthalpies of CYD in both ILs at 303.15 K as a
function of composition. A larger dispersion of the values (up to 3%), especially at higher
concentration, was observed. At times, before each injection of CYD into the IL, an exothermic
effect was detected, mainly originating from insufficient mixing of both components at the
51
beginning of the injection period. As already observed in the case of aromatic hydrocarbons,
the rate of dissolution of non-polar, low density and low viscosity CYD in the IL is slow and
sometimes leads to formation of a solute-rich layer on the surface of the solution. This is often
118
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
accompanied by partial evaporation or even polymerisation of CYD before the mixing process
is complete, hence disturbing the measurements. A large number of injections were made to
ensure reliability.
tel-00708600, version 1 - 15 Jun 2012
Figure 7. Partial molar excess enthalpies vs. mole fraction of CYD in the binary mixture with [C 1C4Im][NTf2]( ) or
[C1C1C4Im][NTf2]( ) at 303.15 K. The lines represent functions: HCYDE=1.6285+0.2481xCYD ( ) and
HCYDE=0.8574-0.1518xCYD ( ).xCYD = CYD molar fraction: [CYD]/([IL]+{CYD])
The formation of a polymer was observed and confirmed by NMR analysis.52 This side-reaction is attributed to the
fact that CYD has been distilled to eliminate the stabiliser BHT before the physical chemical measurements.
119
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
120
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
tel-00708600, version 1 - 15 Jun 2012
Figure 8. ROESY NMR spectrum for the mixture of 1,3-cyclohexadiene and [C1C4Im][NTf2] at R = 0.5.
Figure 9. ROESY NMR spectrum for the mixture of 1,3-cyclohexadiene and [C1C1C4Im][NTf2] at R = 0.5.
121
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Molecular mechanics calculations have been performed, using the SYBYL software with
the TRIPOS force field,54on isolated pairs of CYD and imidazolium cations. Intermolecular
distances were fixed using results from NMR ROESY experiments and the energy of both
systems, CYD-[C1C4Im] and CYD-[C1C1C4Im], were minimised. The average distances between
molecules of CYD and the imidazolium cations were determined from their geometry and
orientation at the potential energy minimum. This model, pictured in Figure 10, suggests that
CYD is probably located nearer to the butyl chain of [C1C4Im][NTf2], the π-bonds being oriented
towards the imidazolium cation, a configuration similar to that previously observed for toluene.41
tel-00708600, version 1 - 15 Jun 2012
Figure 10. Representation of molecular positions in saturated solutions of CYD-[C1C4Im][NTf2] (upper) and CYD-
[C1C1C4Im][NTf2] (lower) from ROESY NMR extrapolation.
122
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
molecule or ion. The molecular simulation details are given in the experimental section 3.5.10. In
Figure 11 are plotted the site-site radial distribution functions — the probability of finding pairs
of atoms at a given distance, compared to the average — between the hydrogens on Cβ of CYD,
Hβ, and selected atoms of the [C1C4Im] and [C1C1C4Im] cations at R = 0.5 (atoms are labelled as
indicated in Scheme 5). In both cases, there is a higher probability (a stronger association) of
finding all hydrogen atoms of CYD near the side chain than in the vicinity of the aromatic
nucleus region of imidazolium cation. This means that CYD is preferentially solvated in the non-
polar domain of the ILs, as already observed for saturated hydrocarbons and for the methyl group
of toluene. This orientation effect relative to the cation is less distinct for CYD than for toluene,
41
as expected given the weaker -cation interactions.
tel-00708600, version 1 - 15 Jun 2012
Figure 11. Radial distribution functions between hydrogen on Cβ of CYD, Hβ, and selected sites in the [C 1C4Im]
cation (top) and [C1C1C4Im] cation (bottom), for R = 0.5. Atoms are labeled as indicated in Scheme 5.
Figure 12 represents the comparison of the radial distribution functions of the mixtures
CYD-[C1C4Im][NTf2] and CYD-[C1C1C4Im][NTf2], again at R = 0.5, between hydrogen on Cβ
123
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
and Cα of CYD and C2 of the imidazolium rings. It can be seen that both these hydrogen atoms
are found with a higher probability closer to C2 in [C1C1C4Im][NTf2] than in [C1C4Im][NTf2].
The strong hydrogen bond between C2-H and the anion in [C1C4Im][NTf2] prevents such an
interaction with CYD. This observation is consistent with the distances between the IL cations
and CYD obtained from ROESY experiments and molecular mechanics calculations on isolated
pairs, as seen in Figure 10.
tel-00708600, version 1 - 15 Jun 2012
Figure 12. Comparison of site-site radial distribution functions between chosen atoms in CYD and the cations, in
CYD-[C1C4Im][NTf2] (top) and CYD-[C1C1C4Im][NTf2] (bottom) for R = 0.5. Atoms are labeled as indicated in
Scheme 5.
124
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
CYD molecule. Oxygen atoms of the NTf2- anion (plotted in red) are found in the plane of CYD
interacting with the hydrogen atoms of the double bonds. Similar results have been previously
obtained for toluene in the same ILs. 41 This shows that although in CYD the π-system is smaller
than that of a fully aromatic system, it still determines the structure of the solvation shell in ILs
with cations positioned above and below the plane of CYD and anion in the plane.
tel-00708600, version 1 - 15 Jun 2012
Figure 13. Spatial distribution functions around the C2 carbon of CYD in CYD-[C1C4Im][NTf2] (left) and CYD-
[C1C1C4Im][NTf2] (right) at R = 0.5. In blue is plotted the iso-surface corresponding to a local density of twice the
average density of C2 carbon of the imidazolium cations. In grey is plotted the iso-surface corresponding to a local
density of twice the average density of terminal methyl carbons from the butyl side chain, C9. In red is plotted the
iso-surface corresponding to a local density of twice the average density of oxygen atoms from the NTf 2- anion.
Oxygen atoms are located in the CYD plane interacting with hydrogen atoms of double bonds, Hα and Hβ.
Figure 14 shows spatial distribution functions around the [C1C4Im] and the [C1C1C4Im]
cations. With both cations, CYD (the white regions) is preferentially located above and below
the plane of the imidazolium ring at distances greater than those of the closest cation-anion pairs.
125
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
In red is plotted the iso-surface corresponding to a local density of 4 times the average density of
the oxygen atoms from [NTf2] anions. As can be seen, for [C1C4Im][NTf2] the probability of
finding the oxygen of [NTf2] near the C2-H of the imidazolium cation is higher than the carbon
of CYD, indicating a strong H-bond. The situation is completely different for [C1C1C4Im][NTf2],
where the probability of finding the carbon of CYD near the C2 of the cation is higher than that
of the oxygen atoms of [NTf2]. As already found in case of solvation of toluene,41interactions of
the [C1C1C4Im] cation with the anion are mainly through H4, H5 and also through the nitrogen
atoms, N1 and N3, whereas in [C1C4Im] cation-anion interactions are mainly through C2-H bonds.
tel-00708600, version 1 - 15 Jun 2012
Figure 14. Spatial distribution functions around the C2 carbon of the cations in CYD-[C1C4Im][NTf2] (left) and
CYD-[C1C1C4Im][NTf2] (right) at R = 0.5. Above and below are different views of the same iso-surfaces. In white is
plotted the iso-surface corresponding to a local density of twice the average density of Cα of CYD. In red is plotted
the iso-surface corresponding to a local density of 4 times the average density of oxygen atoms from the NTf 2–
anion.
126
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Table 1. The 1H NMR chemical shifts of neat [C1C4Im][NTf2], and those of the IL, in the presence of CYD.
R = [CYD]/[IL]. Protons numbered as indicated in Scheme 5.
Proton. R0=0 R1=0.11 Δ(R1- R0) R2=0.21 Δ(R2- R0) R3=0.34 Δ(R3- R0) R4=0.54 Δ(R4- R0) R5=0.81 Δ(R5- R0)
2 8.60 8.62 0.02 8.64 0.04 8.67 0.07 8.71 0.11 8.76 0.16
4 7.45 7.52 0.07 7.54 0.09 7.56 0.11 7.59 0.14 7.63 0.18
5 7.38 7.44 0.06 7.46 0.08 7.49 0.11 7.52 0.14 7.55 0.17
6 4.14 4.22 0.08 4.24 0.10 4.27 0.13 4.30 0.16 4.34 0.20
1 3.91 3.95 0.04 3.97 0.06 3.99 0.08 4.02 0.11 4.07 0.16
7 1.86 1.91 0.05 1.93 0.07 1.95 0.09 1.98 0.12 2.03 0.17
tel-00708600, version 1 - 15 Jun 2012
8 1.36 1.39 0.03 1.42 0.06 1.44 0.08 1.48 0.12 1.53 0.17
9 0.95 0.95 0.00 0.98 0.03 1.01 0.06 1.05 0.10 1.10 0.15
mean Δ 0.04 0.07 0.09 0.13 0.17
Table 2. The 1H NMR chemical shifts of neat [C1C1C4Im][NTf2], and those of the IL, in the presence of CYD.
R = [CYD]/[IL]. Protons numbered as indicated in Scheme 5.
Proton. R0=0 R1=0.08 Δ(R1- R0) R2=0.18 Δ(R2- R0) R3=0.30 Δ(R3- R0) R4=0.40 Δ(R4- R0) R5=0.69 Δ(R5- R0)
4 7.35 7.36 0.01 7.39 0.04 7.41 0.06 7.42 0.07 7.46 0.11
5 7.28 7.29 0.01 7.32 0.04 7.34 0.06 7.36 0.08 7.40 0.12
6 4.08 4.09 0.01 4.12 0.04 4.14 0.06 4.16 0.08 4.20 0.12
1 3.76 3.78 0.02 3.81 0.05 3.83 0.07 3.85 0.09 3.89 0.13
2 2.57 2.59 0.02 2.62 0.05 2.64 0.07 2.66 0.09 2.71 0.14
7 1.77 1.79 0.02 1.82 0.05 1.85 0.08 1.87 0.10 1.91 0.14
8 1.36 1.37 0.01 1.41 0.05 1.44 0.08 1.46 0.10 1.51 0.15
9 0.91 0.93 0.02 0.97 0.06 1.00 0.09 1.02 0.11 1.08 0.17
mean Δ 0.02 0.05 0.07 0.09 0.14
Unlike in the case of toluene-IL mixtures,41 it is found that all peaks are generally shifted
downfield with increasing molar ratio of CYD, R = [CYD]/[IL], demonstrating the deshielding
effect CYD has on all hydrogen nuclei. The upfield shift of IL proton NMR shifts in toluene-IL
mixtures was attributed to π-cation interactions, which according to our molecular models do not
exist in the case of CYD.
Excluding the first point of each series, Δδ versus R follows a straight line, Figure 15.
The slopes (0.17 ppm for [C1C4Im][NTf2]; 0.18 ppm for [C1C1C4Im][NTf2]) are almost identical
for both ILs. The similar linear variation of the deshielding effect for both ILs shows that no
127
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
0.2
0.18
0.16
0.14
Mean Δδ / ppm
0.12
0.1
0.08 [C1C4Im][NTf2]
0.06 [C1C1C4Im][NTf2]
0.04
0.02
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
R
Figure 15. Mean deviation of proton chemical shifts in IL-CYD mixtures with respect to R
The variation of the CYD chemical shifts when changing R were also studied and are
tabulated in Table 3. As for the shifts of the IL moieties, CYD shifts are shifted downfield
demonstrating once again the deshielding effect of CYD, even on itself. At low concentration the
CYD molecules interact mostly with the alkyl chains of the ILs, but as the concentration
increases, CYD molecules will also interact with one another.
128
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Table 3. 1H NMR chemical shifts of CYD in the ILs. R = [CYD]/[IL]. Protons labelled as indicated in Scheme 5.
[C1C4Im][NTf2]/CYD [C1C1C4Im][NTf2]/CYD
R R
α β γ α β γ
0.11 5.84 5.79 2.08 0.08 5.81 5.75 2.06
0.21 5.88 5.81 2.11 0.18 5.85 5.78 2.10
0.34 5.92 5.85 2.15 0.30 5.87 5.82 2.13
0.54 5.94 5.88 2.19 0.40 5.91 5.83 2.16
0.81 6.00 5.93 2.41 0.69 5.97 5.90 2.22
Table 4. Density and viscosity, at 298.15 K and atmospheric pressure, of mixtures of CYD-[C1C4Im][NTf2] and
CYD-[C1C1C4Im][NTf2] at R = CYD/IL.
R xCYD ρ (g/mL) η (mPas) R xCYD ρ (g/mL) η (mPas)
[C1C4Im][NTf2] [C1C1C4Im][NTf2]
129
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Diffusivity
Another possible way to obtain information on how the media influences the mobility of
molecules is to determine their diffusion coefficients, D, by NMR spectroscopy. Diffusion Order
SpectroscopY, DOSY*, is an NMR diffusion experiment, which provides a way to separate the
different compounds in a mixture, based on the differing translational diffusion coefficients (and
therefore differences in the size and shape of the molecule, as well as physical properties of the
surrounding environment such as viscosity, temperature, etc.) of each chemical species in
solution.
Generally, according to the Stokes-Einstein equation (2), the diffusion coefficient is
inversely proportional to the viscosity of the medium.
k BT
D (2)
tel-00708600, version 1 - 15 Jun 2012
6 rH
kB = Boltzman constant, rH = hydrodynamic radius
For instance, when R is around 0.5, the diffusion coefficients found for CYD are
187.5 µm2/s in [C1C4Im][NTf2] against 97 µm2/s in [C1C1C4Im][NTf2], i.e. diffusion of CYD in
[C1C4Im][NTf2] is 1.9 times faster than in [C1C1C4Im][NTf2] in agreement with their viscosity
ratio.
The diffusion coefficients (D) of CYD in both ILs while increasing R were thus
determined from extrapolation of DOSY NMR data, and are plotted in Figure 16. The diffusion
coefficients were found to increase with R in both ILs.
*
see experimental section 3.5.9 for further details
130
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
3.00E-10
[C1C4Im][NTf2]
2.50E-10
[C1C1C4Im][NTf2]
2.00E-10
D / µm2 /s
1.50E-10
1.00E-10
5.00E-11
0.00E+00
tel-00708600, version 1 - 15 Jun 2012
3 1.8
[C1C4Im][NTf2] 1.6
2.5 [C1C1C4Im][NTf2]
grad 2 = 3.8 10 -9 1.4
grad 2 = 3.4 10 -9
D / 10 -10 µm2 s-1
D / 10 -10 µm2 s-1
2 1.2
1
1.5
0.8
grad 1 = 5.6 10 -9 grad 1 = 5.6 10 -9
1 0.6
0.4 grad 1: grad 2 = 1.6
0.5 grad 1: grad 2 = 1.5
0.2
0 0
0 0.02 0.04 0.06 0.08 0 0.01 0.02 0.03 0.04
η-1 / (mPa.s)-1 η-1 / (mPa.s)-1
Figure 17. Calculated inverse viscosity η-1 plotted as a function of D for CYD in CYD-IL mixtures.
This may indicate different mobility mechanism for low and high concentration mixtures
i.e. for different viscosity. Indeed, in certain cases the Stokes-Einstein model may not be
131
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
k BT 1 2 / rH
D 6 (3)
rH 1 3 / rH
k BT
D (4)
4 rH
Indeed, a recent study of the diffusion of ferrocene in imidazolium ionic liquids found
tel-00708600, version 1 - 15 Jun 2012
that due to the lack of specific interactions between ferrocene and the surrounding medium, the
Sutherland coefficient of 4 was more applicable.46 We have already demonstrated in section
3.2.1 that CYD does not display strong specific interactions with the IL. This brings into
question the suitability of the Stokes-Einstein equation in the case under study.
When we calculate the ratio of the gradients of the curves of η-1 plotted as a function of
D, Figure 17, at low vs. high concentration of CYD, we find a ratio of 1.5 for [C1C4Im][NTf2]
and 1.6 for [C1C1C4Im][NTf2]. This could be explained by Stokes-Einstein behaviour at high
concentration (i.e. low viscosity) θ = 6, and “sliding surface” behaviour at low concentration,
θ = 4. This corroborates the lack of intermolecular interactions found between the CYD and the
ILs. More measurements are underway to confirm this.
Nevertheless, it is clear that the viscosity is a complex issue that must be taken into
account when ILs are used as media for catalysis.
3.2.3 Hydrogenation of 1,3-cyclohexadiene
In organic solvents, during the hydrogenation of 1,3-cyclohexadiene, the catalyst first
undergoes a ligand exchange yielding [Rh(CYD)(PPh3)2]+. This occurs more rapidly than the rate
at which it then reacts with H2.56, 57
The two reaction steps, depicted in Scheme 6, were
investigated in both ILs.
3.2.3.1 Monitoring by UV-vis spectroscopy in both ILs
Firstly, [Rh(CYD)(PPh3)2]NTf2 was synthesised from [Rh(COD)(PPh3)2]NTf2 and fully
characterised. As [Rh(COD)(PPh3)2]NTf2 is yellow ( max = 450 nm) and [Rh(CYD)(PPh3)2]NTf2
132
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
is red ( max = 500 nm), the ligand exchange COD-CYD could be monitored by UV-Vis
spectroscopy ( = 500 nm) in both ILs, as shown in Figure 18. This reaction is much slower in
[C1C1C4Im][NTf2] than in [C1C4Im][NTf2].
tel-00708600, version 1 - 15 Jun 2012
Scheme 6. Ligand exchange and catalytic cycle for the hydrogenation of CYD using an Osborn-type catalyst in ILs
133
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
tel-00708600, version 1 - 15 Jun 2012
Figure 18. Evolution as a function of time of the ligand exchange reaction COD-CYD for
[Rh(COD)(PPh3)2]NTf2 in [C1C4Im][NTf2] (full line) and [C1C1C4Im][NTf2] (dashed line) monitored at = 500 nm
by UV-Vis spectroscopy. UV-visible spectra of [Rh(COD)(PPh3)2]NTf2 and [Rh(CYD)(PPh3)2]NTf2 are provided in
section 3.5.3
134
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
tel-00708600, version 1 - 15 Jun 2012
Figure 19. Evolution of the absorbance at = 500 nm as a function of time during the hydrogenation reaction of
CYD in the presence of [Rh(CYD)(PPh3)2]NTf2 in [C1C4Im][NTf2]( ) and [C1C1C4Im][NTf2]( ).
Scheme 7. The two steps involved in the catalytic hydrogenation of CYD, the attack by H 2 being rate-
determining.
135
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
entire reaction system after dissolution in a mixture of acetonitrile and toluene 99:1. Each point,
repeated at least twice, presented in Figure 20, corresponds to a different experiment
tel-00708600, version 1 - 15 Jun 2012
Figure 20. Hydrogenation of CYD at 303 K under 1.2 bar H2 (R = 0.5 and r = n(CYD)/n(Rh) = 500) in
[C1C4Im][NTf2]( ) and [C1C1C4Im][NTf2]( ).
136
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
In these conditions, at 303 K, and under 1.2 bar H2, conversion curves for different values
of R are plotted in Figure 21. It can clearly be seen that the rate of this reaction increases with
increasing molar ratio R. When R = 0.1, the conversion rate is found to be very low, values
obtained falling within the experimental error, thus this result is not depicted in Figure 21.
45
40
R = 0.5
35
R = 0.4
30
R = 0.3
Conversion %
25 R = 0.2
20
15
10
0
0 10 20 30 40 50 60 70
Time / mins
Figure 21. Conversion curves for the hydrogenation of CYD with varying molar ratio R = CYD/IL, in
[C1C4Im][NTf2]
137
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
From literature data and previous results obtained by UV, we hypothesised that the rate
determining step in this reaction is the attack by H2 on the catalyst and not the coordination of
the substrate. In such situations the rate law should be expressed as (5). 58
As the coordination of the substrate to the catalyst is rapid, we can consider the
concentration of substrate-coordinated catalyst, [MS] in (5), equal to the concentration of
catalyst.
tel-00708600, version 1 - 15 Jun 2012
The concentration and diffusivity of H2 in these media (IL:CYD mixture) has been
neither measured nor calculated. Even in neat ILs, no experimental data were found on the
diffusivity of hydrogen. Nevertheless, the diffusivity is expected to be inversely proportional to
the viscosity of the liquid medium since H2 interacts weakly with ILs59-62
In an attempt to establish a rough approximation to the rate law of this reaction, a graph is
plotted of the initial reaction rate versus a combination of these known and controlled variables,
potentially affecting the rate of reaction, i.e. concentration of CYD and catalyst, also including a
-1
term for the viscosity, . As seen in Figure 22, this gives us a near linear relationship, passing
close to origin, indicating that the rate varies inversely with viscosity as expected from previous
results. The rate also varies linearly with the concentration of both catalyst and substrate. (Note
that plots accounting for only one or two of these variables do not give straight lines.) Our curve
-1
therefore seems to indicate that the rate is proportional to [Rh][CYD] .
138
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
160
140
initial rate /(10-6 mol L-1 s-1 )
120
100
80
60
40
20
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
tel-00708600, version 1 - 15 Jun 2012
Figure 22. linear relationship of initial rate against the inverse viscosity measured experimentally
It is clear that the system under investigation is complex. To establish the correct rate law
for this reaction, the solubility of H2 in ILs, CYD and IL-CYD binary mixtures must be
accurately measured. We must also determine the effect of changing pressure on both the
solubility of H2 in the media and the viscosity. Furthermore, conductivity must be taken into
account as this may affect the mobility of the ionic catalyst.63
139
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
3.3 Conclusion
140
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
reaction the viscosity and thus mass transport is indeed the deciding factor when performed in
ILs.
The exact kinetic rate law cannot be determined without precise knowledge of H2
solubility and diffusivity in the media and also studies of conductivity and its potential effect on
the mobility of the ionic catalyst. This work is currently underway.
We have successfully demonstrated in this chapter that to fully understand the kinetics of
catalytic processes carried out in IL media, a comprehensive study of all physical-chemical
variables is necessary. Consequently, when comparing literature data on catalytic activity in ILs
the differences in physical-chemical properties must be taken into account, and may shed light
on differences noted. In the case under study it has been found that mass transport factors play a
decisive role.
3.4 Outlooks
For the hydrogenation of CYD in these ILs, we have demonstrated that the rate
determining step involves the insertion of H2. Future work is planned to determine the rate law of
this reaction through a full kinetic study varying conditions such as H2 pressure. For this, due to
the complicated viscosity-dependent behaviour and the interesting solvation properties of ILs, a
study of the viscosity and H2 solubility dependence on H2 pressure must first be undertaken.
Also, a study of conductivity in CYD-IL mixtures, with varying R may give useful information
on the mobility of the ionic catalyst in the different media.
141
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Finally, as aromatic substrates such as toluene have already been shown to display more
complicated behaviour when dissolved in ILs, i.e. π-cation interactions, it may be interesting to
observe the effect of such interactions in catalysis.
142
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Yield: 100 mg (99%). 1H NMR (300 MHz, CD2Cl2): = 2.1 (s, 2H), 5.3 (s, 2H), 7.5 (m, 30H).
31
P NMR (121 MHz, CD2Cl2): = 25.2ppm with JRh-P = 17 Hz. Mass spectrometry (ESI):
positive mode: m/z=707 Rh(cyclohexadiène)(PPh3)2+; m/z=627 Rh(PPh3)2+. UV-Vis
spectroscopy (6.6 mg of [Rh(CYD)(PPh3)2]NTf2 dissolved in 2 mL of one of ionic liquids,
[C1C4Im][NTf2] or [C1C1C4Im][NTf2]); spectrum recorded at max = 500 nm.
3.5.3 Ligand exchange
Ligand exchange in [Rh(CYD)(PPh3)2]NTf2 was followed by UV-Vis spectroscopy.
CYD (0.15 mL, 1.6 mmol) was added to a system of [Rh(COD)(PPh3)2]NTf2 (1.6 mg, 1.6 µmol)
in one of the ionic liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2] (1 mL). The UV-visible spectra
of 1 mL of the solution were recorded on the Perkin-Elmer LAMBDA 950 Spectrophotometer in
stirred and closed UV cells at max = 500 nm every second. The spectra for both complexes are
tel-00708600, version 1 - 15 Jun 2012
Absorbance
2.0 2.0
1.5 1.5
1.0 1.0
0.5 0.5
0.0 0.0
400 450 500 550 600 650 700 750 800 400 450 500 550 600 650 700 750 800
λ (nm) λ (nm)
Figure 23. UV-vis absorption spectra for [Rh(COD)PPh3] (left), and [Rh(CYD)PPh3] (right)
3.5.4 Reaction with H2
[Rh(CYD)(PPh3)2]NTf2 (8 mg, 8.1×10-6 mmol) was dissolved in one of the ionic liquids,
[C1C4Im][NTf2] or [C1C1C4Im][NTf2] (5 mL) and pressurised under 1.2 bar of hydrogen. The
UV-visible spectra of 1 mL of the solution were recorded at a given time at max = 500 nm.
3.5.5 Hydrogenation of 1,3-cyclohexadiene
The hydrogenation of CYD was carried out at 1.2 atm of H2 and 30 °C. CYD (0.15 mL, 1.6
mmol) (0.06 mL, 0.63 mmol) was dissolved in a system of [Rh(COD)(PPh3)2]NTf2 (3.2 mg, 3.2
µmol) in one of ionic liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2] (1mL), under argon
resulting in red homogeneous solutions. The reaction mixture was kept under hydrogen
143
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
atmosphere (1.2 atm, constant pressure) until 4 mL of acetonitrile were added to the catalytic
solution. The product distribution in the reaction mixture and the conversion were determined by
GC analyses in presence of toluene as an internal standard. A HP6890 chromatograph equipped
with FID detector and an Al2O3/KCl column (L = 50 m, int = 0.32 mm, film thickness = 5 µm)
was used. Injector and detector temperature were set to 230 °C. Samples were injected in volume
of 1 µL. The temperature of the column was fixed at 190 °C. From those relative response
factors, the mass of each product can be determined by thegeneral formula:
A(x)
A
M(x) = K(x) M(s)
A(s)
A
M(x); mass of product x
K(x); relative response factor of product x
A(x); peak area of product x
tel-00708600, version 1 - 15 Jun 2012
144
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
temperature was monitored using a platinum resistance thermometer with a precision of 0.1 K.
The temperature of the bath was first increased slowly until one phase was observed. The clear
homogenous system was then cooled down very slowly (5 K/h) under continuous stirring. The
temperature at which the first sign of turbidity (first cloudiness) appeared was considered as the
temperature of the liquid-liquid phase transition. The overall accuracy in the measurement of
cloud-point temperatures is estimated to be ± 2 K.
3.5.7 Isothermal titration calorimetry
The heat effects resulting from mixing of aliquots of CYD with the ionic liquid were
measured at 303.15 K using an isothermal titration nanocalorimeter equipped with 4 mL glass
cells in a Thermal Activity Monitor TAM III from TA Instruments. An electrical calibration was
done before each experiment and the instrument was chemically calibrated 5 times by titration of
tel-00708600, version 1 - 15 Jun 2012
0.01 M aqueous solution of 18-crown-6 ethers with 0.2 M aqueous solution of BaCl2.
Approximately 2.75 mL of degassed ionic liquid was introduced into 4 mL glass
measuring and reference cells. The liquid in the measuring cell was stirred by a turbine stirrer at
160 rpm and volumes of 4 μL of CYD were injected during 180 seconds using a motor driven
pump (Thermometric 3810 Syringe Pump) equipped with a 100 μL gas-tight Hamilton syringe.
In all experiments, the intervals between consecutive injections were of 35 to 40 minutes, which
provided a good thermal stabilisation of the ionic liquid solution and the return to a stable
baseline. To minimise the undesirable effects of diffusion of the ionic liquid into the canula
linking the CYD syringe to the cell, the canula was immersed in the sample 10 minutes prior the
first injection.
A peak with an area proportional to the resulting heat effect Qi translates to the thermal
effect due to each injection of CYD. The integration of peaks from the recorded calorimetric
plots was performed using the TAM III Assistant software. Each experiment was repeated four
times to obtain reproducible values of Qi at different concentrations within the error bar of ± 2 %.
3.5.8 Density and viscosity
The mixtures of ionic liquid and CYD at different compositions were prepared
gravimetrically following the procedure already described, including the precautions to minimize
vapour head-space.41The viscosity of the mixture was measured at 298.15 K (controlled to within
± 0.005 K and measured with the accuracy better than ± 0.05 K) using a rolling ball viscometer
from Anton Paar, model AMVn, equipped with capillary tubes of 3.0 and 1.8 mm in diameter.
145
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Before starting the measurements, the 3.0 mm diameter capillary tube was calibrated as a
function of temperature and angle of measurement with standard viscosity oil from Cannon
(N35). The 1.8 mm diameter tube was calibrated with water by the manufacturer. The overall
uncertainty of the viscosity is estimated as ± 2.0 %.
The densities of the mixtures, necessary to calculate the viscosities were measured in an
Anton Paar vibrating tube densimeter model 512 P, at 298.15 K (measured by a calibrated PRT
with an accuracy of ± 0.02 K). The densimeter was calibrated using n-heptane, bromobenzene
and 2,4-dichlorotoluene. The overall uncertainty of the density is estimated as ± 0.01 %.
3.5.9 NMR spectroscopy
1
H, 13C and 31P solution NMR data were collected at room temperature on a Bruker AC
300 MHz spectrometer with the resonance frequency at 300.130 MHz for 1H nucleus. 103
Rh
tel-00708600, version 1 - 15 Jun 2012
solution NMR was carried out on a Bruker DRX 500 instrument at 298 K (nominal) with a
resonance frequency at 500.130 MHz. The solvent used (CD2Cl2) was distilled and kept in a
rotaflo with molecular sieves. Chemical shifts are reported in ppm (singlet = s, doublet = d,
doublet of doublet = dd, and multiplet = m) and were measured relative to the residual proton of
the solvent to CHDCl2 for 1H, to CD2Cl2 for 13C and to H3PO4 for 31P spectra.
For 1H 1D NMR spectroscopy, the samples with molar ratio R = 0.5 (R is the molar ratio
between the amount of substance of the hydrocarbon and the amount of substance of IL) were
prepared. The mixtures of ionic liquids and hydrocarbon were prepared in closed vials in a glove
box by adding the appropriate amount of CYD to each ionic liquid, [C 1C4Im][NTf2] or
[C1C1C4Im][NTf2]. The resulting systems were stirred for 24 h at 303 K resulting in
homogeneous monophasic solutions. Approximately 0.3 mL of the sample was then introduced
into a 5 mm NMR tube. A stem coaxial capillary tube loaded with CD2Cl2 was inserted into the 5
mm NMR tube to avoid any contact between the deuterated solvent and the analysed mixture.
The deuterium in CD2Cl2 was used for the external lock of the NMR magnetic field and the
residual CHDCl2 in CD2Cl2 was used as the 1H NMR external reference at 5.32 ppm. When 1H
NMR data are obtained in this way, the reference signal of CHDCl2 remains constant and is not
affected by changes in sample concentration.
For ROESY (Nuclear Overhauser Effect SpectroscopY) experiments in the rotating
frame, the 2D sequence was built with the scheme proposed by Bodenhausen (the pulse sequence
shown in Figure 24).64 The mixing time (200 ms) is split into two parts separated by a p-pulse.
146
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
At each side of the spin lock the B1 field is ramped linearly (4.5 ms) to ensure adiabatic
conditions for spinlock. During the first spinlock pulse, the frequency is shifted to O1 + Df
whereas the second pulse frequency is set to O1 – Df; O1 is the offset frequency and Df set to
give a B1 field at the magic angle. By this way, the ROESY response is roughly constant across
the spectra of interest. 1D sequence, PFGSE (Pulse Field Gradient Spin Echo for Selective
Excitation) have been used and spinlock followed the same scheme as previously.
Molecules in liquid or solution state move. This translational motion is, in contrast to
rotational motion, known as Brownian molecular motion and is often simply called diffusion or
self-diffusion. It depends on a lot of physical parameters like size and shape of the molecule,
temperature, and viscosity. Assuming a spherical size of the molecule the diffusion coefficient D
is described by the Stokes-Einstein equation (2):
Pulsed field gradient NMR spectroscopy can be used to measure translational diffusion of
molecules and is sometimes referred to as q-space imaging. By use of a gradient, molecules can
be spatially labelled, i.e. marked depending on their position in the sample tube. If they move
after this encoding during the following diffusion time D, their new position can be decoded by a
second gradient. The measured signal is the integral over the whole sample volume and the NMR
signal intensity is attenuated depending on the diffusion time D and the gradient parameters (g,
δ). This intensity change is described by equation (4):
2 2 2
D g
I I 0e (4)
I; the observed intensitu
I0; reference intensity (unattenuated signal intensity)
D; diffusion coefficient
g; gradiant strength
γ; gyromagnetic ratio
δ; length of the gradient
Δ; diffusion time
147
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
To simplify this equation some parameters are often combined to emphasize the
exponential decay behaviour.
Dq2 ( )
3 DQ
I I 0e or I I 0e (8)
If bipolar gradients are used for dephasing and rephasing a correction for the time t
between those bipolar gradients has to be applied (6).
D 2g 2 2
( / 3 r / 2)
I I 0e (9)
2D DOSY experiments used a slightly modified Bruker experiment ledbpgp2s to improve
lineshape and trapezoidal gradients where implemented for shorter pulses gradients (the pulse
sequences shown in Figure 25). The diffusion evolution time was set to 100 ms, the constant
amplitude part of the gradient was set to 3 ms, and the cosine raising and falling part of gradient
tel-00708600, version 1 - 15 Jun 2012
where set to 150 µs. The diffusion space where sampled by 32 linearly spaced gradients. Values
obtained for D using this technique are tabulated in Tables 4 and 5, as well as viscosity values
and the hydrodynamic radii calculated from these experimental data according to the Stokes-
Einstein relationship (2).
148
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
Although this may be important to correctly reproduce dynamic properties of ILs, structural
features and thermodynamic quantities have been described to equivalent levels of accuracy
using fixed-charge models.69
Molecular dynamics simulations of condensed-phase CYD-[C1C4Im][NTf2] and CYD-
[C1C1C4Im][NTf2] mixtures were performed using the DL_POLY program.70 System sizes were
chosen so as to contain about 10000 atoms, and so the numbers of cations, anions and CYD
molecules varied according to composition (e.g. 128 ion pairs and 64 molecules of CYD for
R = 0.5). Initial low-density configurations, with ions and molecules placed at random in
periodic cubic boxes, were equilibrated to attain liquid-like densities and structures at 400 K and
1 bar. Temperature and pressure were maintained using Nosé-Hoover thermostat and barostat,
respectively. Production runs then took 500 ps with an explicit cut-off distance of 16 Å for non-
bonded interactions, and long-range corrections applied for repulsive-dispersive interactions.
Electrostatic energies were calculated using the Ewald summation method with a relative
accuracy of 10-4. Structural quantities such as radial and spatial distribution functions were
calculated from configurations generated during the production runs
149
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
3.6 References
150
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
26. L. Magna, Y. Chauvin, G. P. Niccolai and J.-M. Basset, Organometallics, 2003, 22,
4418-4425.
27. V. Lecocq and H. Olivier-Bourbigou, Oil & Gas Sci. Tech, 2007, 62, 761-773.
28. U. Hintermair, T. Gutel, A. M. Z. Slawin, D. J. Cole-Hamilton, C. C. Santini and Y.
Chauvin, J. Organometal. Chem., 2008, 693, 2407-2414.
29. J. A. Boon, J. A. Levisky, J. L. Pflug and J. S. Wilkes, J. Org. Chem., 1986, 51, 480-483.
30. A. Mele, G. Romano, M. Giannone, E. Ragg, G. Fronza, G. Raos and V. Marcon, Angew.
Chem. Int. Ed., 2006, 45, 1123-1126.
31. A. Triolo, O. Russina, H.-J. Bleif and E. Di Cola, J. Phys. Chem. B, 2007, 111, 4641-
4644.
32. A. Triolo, O. Russina, B. Fazio, G. B. Appetecchi, M. Carewska and S. Passerini, J.
Chem. Phys., 2009, 130, 164521/164521-164521/164526.
33. J. Fuller, R. T. Carlin, H. C. De Long and D. Haworth, J. Chem. Soc., Chem. Commun.,
1994, 299-300.
34. C. Hardacre, J. D. Holbrey, S. E. J. McMath, D. T. Bowron and A. K. Soper, J. Chem.
Phys., 2003, 118, 273-278.
35. J. N. A. Canongia Lopes and A. A. H. Padua, J. Phys. Chem. B, 2006, 110, 3330-3335.
36. Y. Wang, S. Izvekov, T. Yan and G. A. Voth, J. Phys. Chem. B, 2006, 110, 3564-3575.
37. A. A. H. Padua, M. F. Costa Gomes and J. N. A. Canongia Lopes, Acc. Chem. Res., 2007,
40, 1087-1096.
151
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
44. T. Umecky, Y. Saito and H. Matsumoto, J. Phys. Chem. B, 2009, 113, 8466-8468.
45. T. Umecky, Y. Saito and H. Matsumoto, ECS Transactions, 2010, 25, 23-29.
46. M. A. Vorotyntsev, V. A. Zinovyeva and M. Picquet, Electrochim. Acta 2010, 55, 5063-
5070.
47. V. Kempter and B. Kirchner, J. Mol. Struct., 2010, 972, 22-34.
48. P. A. Hunt, J. Phys. Chem. B, 2007, 111, 4844-4853.
49. M. Blesic, J. N. Canongia Lopes, A. A. H. Padua, K. Shimizu, M. F. Costa Gomes and L.
P. N. Rebelo, J. Phys. Chem. B, 2009 113, 7631.
50. U. Domańska, Z. Żołek-Tryznowska and M. Królikowski, J. Chem. Eng. Data, 2007, 52,
1872.
51. W. Marczak, S. P. Verevkin and A. Heintz, J. Solution Chem., 2003, 32, 519-526.
52. I. Natori, K. Imaizumi, H. Yamagishi and M. Kazunori, J. Polym. Sci. B 1998, 36, 1657.
53. D. Frezzato, F. Rastrelli and A. Bagno, J. Phys. Chem. B, 2006, 110, 5676-5689.
54. M. Clark, R. D. Cramer, III and N. Van Opdenbosch, J. Comp. Chem., 1989, 10, 982-
1012.
55. W. Sutherland, Philos. Mag., 1905, 781.
56. R. R. Schrock and J. A. Osborn, J. Am. Chem. Soc., 1976, 98, 4450-4455.
57. R. R. Schrock and J. A. Osborn, J. Am. Chem. Soc., 1976, 98, 2134-2143.
58. P. W. N. M. Van Leeuwen, Homogeneous Catalysis Understanding the art, Kluwer,
Dordrecht, 2004.
152
Chapter 3 Imidazolium ILs as Media for Homogeneous Catalysis
59. J. Jacquemin, M. F. Costa Gomes, P. Husson and V. Majer, J. Chem. Thermodyn., 2006,
38, 490-502.
60. J. Jacquemin, P. Husson, V. Majer and M. F. Costa Gomes, Fluid Phase Equilib., 2006,
240, 87.
61. M. F. Costa Gomes, J. Chem. Eng. Data, 2007, 52, 472-475.
62. W. Shi, D. C. Sorescu, D. R. Luebke, M. J. Keller and S. Wickramanayake, J. Phys.
Chem. B, 2010, 114, 6531-6541.
63. K. Ueno, H. Tokuda and M. Watanabe, Phys. Chem. Chem. Phys. , 2010, 12, 1649-1658.
64. B. Cutting, R. Ghose and G. Bodenhausen, J. Magn. Reson., 1999, 138, 326-329.
65. J. N. Canongia Lopes, J. Deschamps and A. A. H. Pádua, J. Phys. Chem. B, 2004, 108,
2038-2047.
tel-00708600, version 1 - 15 Jun 2012
66. J. N. Canongia Lopes and A. A. H. Pádua, J. Phys. Chem. B, 2004, 108, 16893-16898.
67. J. N. Canongia Lopes, A. A. H. Padua and K. Shimizu, J. Phys. Chem. B 2008, 112,
5039-5046.
68. W. L. Jorgensen, D. S. Maxwell and J. Tirado-Rives, J. Am. Chem. Soc., 1996, 118,
11225-11236.
69. E. J. Maginn, Acc. Chem. Res., 2007, 40, 1200-1207.
70. W. Smith, T. R. Forester and I. T. Todorov, ed. T. D. P. m. s. package, STFC Daresbury
Laboratories, , Warrington , UK, 2007, p. The DL_POLY molecular simulation package.
153
tel-00708600, version 1 - 15 Jun 2012
154
IV
tel-00708600, version 1 - 15 Jun 2012
Chapter 4
Nanoparticle Catalysis in
Imidazolium Ionic Liquids
155
tel-00708600, version 1 - 15 Jun 2012
156
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
Chapter 4
157
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
158
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
4.1 Introduction
159
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
increasing size, which was attributed to the greater availability of the (111) face, the most active
site for this reaction.10 The activity in CO hydrogenation of supported RhNPs with hcp structure
has been found to increase drastically with NP size, due to the larger number of steps on the
11
surface. In contrast, in ethane hydrogenolysis to methane over Pt nanoparticles, the turnover
rate has been shown to decrease with particle size while activation energy increases, indicating
that the more unsaturated vertex and/or edge positions are responsible for activity.12
4.1.2 The influence of size on selectivity
When several different products are possible, selectivity becomes an important issue.
This has been shown to vary with particle size. For instance, the hydrogenation/dehydrogenation
of cyclohexene over PtNPs has been shown to exhibit size-dependent selectivity, smaller NPs
favouring benzene production, Figure 2.13
tel-00708600, version 1 - 15 Jun 2012
Furthermore, it has also been reported in a study of the hydrogenation of benzene over Pt,
that on the (111) face cyclohexane and cyclohexene are produced, whereas on the (100) face
only cyclohexene is produced. This result implies that cubic NPs, exhibiting uniquely (100)
faces, would produce selectively cyclohexene in the hydrogenation of benzene.6, 13
160
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
From the results mentioned above and many more examples in the literature,6 it is clear
that the size of NPs can be a deciding factor in catalytic processes. Nonetheless, the synthesis of
nanoparticles (NPs) with a controlled size in the range of 1–10 nm in order to corroborate this
theory is a challenging issue.14 The aim of this chapter is to use the ionic liquid route to
synthesise catalytically active RuNPs of distinct and controlled sizes, and thus investigate
thoroughly the consequences of changing NP size on the catalytic process, in particular
hydrogenation.
161
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
2 nm
5 nm
tel-00708600, version 1 - 15 Jun 2012
5 nm
Figure 3. Transition electron micrograph of RuNPs and high resolution electron micrograph examples showing
crystallinity for Ru0 (top left), Ru25 (top right), Ru50 (bottom left) and Ru75 (bottom right).
162
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
160
140 T = 0 °C
T = 25 °C
120
T = 50 °C
100 T = 75 °C
Frequency 80
60
40
20
0
1 2 3 4 5
Size/ nm
Figure 4. Size distribution histograms for RuNPs prepared in [C1C4Im][NTf2] at different temperatures.
tel-00708600, version 1 - 15 Jun 2012
163
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
Figure 5. XPS spectra of the Ru 3p region after NP filtration onto SiO2, experimental data and fitted peaks.
Left Ru0, middle Ru25 and right Ru50.
variables must remain unchanged. Therefore maintaining constant the initial ratio of substrate to
catalyst is imperative. In NP catalysis, as in heterogeneous catalysis, only the atoms at the
surface (Rus) take part in reaction. The dispersion (D) presents the ratio between surface atoms,
Rus and the total number of atoms, RuT (D = Rus/RuT) and varies with the size of the NPs,
smaller particles of course having a larger percentage of surface atoms. The different dispersion
values must therefore be taken into account for each size of nanoparticle formed.
Ruthenium is known to exhibit a hexagonal close-packed crystal structure, with the
following lattice parameters: a: 270.59 pm b: 270.59 pm c: 428.15 pm α: 90° β: 90° γ: 120°.19
Using these parameters, SYBYL software can be applied to extrapolate the lattice until the
measured diameters, in order to model the structure of the different size NPs, assuming
crystallinity. It is seen that crystalline hexagonal close-packed RuNPs would adopt a truncated
hexagonal bipyramid form, with two symmetric hexagonal faces (0001) and twelve irregular and
uneven trapezoid faces (10 –11).20 From these findings a curve of D with respect to diameter can
be plotted and then used to estimate D for each size of nanoparticle, Figure 6.
164
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
100%
90%
80%
70%
60%
Dispersion
50%
40%
30%
20%
Dispersion = 0.8622 Diameter-0.63
tel-00708600, version 1 - 15 Jun 2012
10%
0%
0 0.5 1 1.5 2 2.5 3 3.5
Diameter / nm
Figure 6. Curve of dispersion against mean diameter of crystalline hcp RuNPs.
4.3 Substrate
165
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
of the partially hydrogenated product.24 The same group has also described the possibility of
extracting cyclohexene during benzene hydrogenation using this solubility difference. 25 For this
reason, solubilities of CYD, CYE and CYA were measured. It was found that the solubility of
the hydrogenated products (6 ± 1 % wt - CYE, 4 ± 1 % wt- CYA) is much lower than that of
CYD (12 ± 2 % wt), therefore the medium may tend to a biphasic system during the course of
the reaction. As a result, the collection of aliquots from a single batch would render inaccurate
results. Consequently, each measure of conversion recorded in this work, hence each point
plotted, corresponds to a separate experiment, quenched after time t by opening the reaction
vessel to air, thus releasing the hydrogen, and dissolving the catalytic system entirely in a 1
molar solution of toluene in acetonitrile for gas phase chromatography.
4.3.2 Viscosity
tel-00708600, version 1 - 15 Jun 2012
Thermophysical properties of the reaction medium such as density and viscosity may also
influence the catalytic performance. As demonstrated in Chapter 3, and by others, reaction
kinetics in IL media are highly dependent on the mobility of molecules.21,26-28 Furthermore,
knowledge of the viscosity is very important from engineering point of view as it plays a major
role in stirring, mixing and pumping processes. Consequently, identical concentrations of
substrate must be used in each case in order to maintain constant viscosity and eliminate effects
due to mass transport. The densities and viscosities of the pure [C1C4Im][NTf2] and that of the
mixtures with CYD were measured at different molar ratio CYD/IL (R) at 25 °C and atmospheric
pressure. The results are presented in Table 1.
Table 1. Density, ρ, and viscosity, η, of CYD-IL mixtures of different compositions at 298 K and atmospheric
pressure. xIL= molar fraction of IL, R = molar ratio CYD/IL
R xIL ρ/g cm-3 η/mPa s
166
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
As can be seen, the viscosity of the mixtures of CYD in IL varies greatly with the
concentration of CYD. From the Stokes-Einstein equation, the diffusion coefficient, D, which
reflects the mobility of molecules, varies inversely with η, equation (1), see Chapter 3.
kT
D (1)
rH
For the purpose of our investigations the concentration of CYD is fixed at 10 % wt.
(corresponding to a molar ratio R = 0.59), which is lower than the limit of solubility, limiting
phenomena due to formation of a biphasic system. According to results reported in Chapter 3, a
molar ratio of R = 0.59 gives a perfectly “sticking” medium for which the Stokes-Einstein
equation (1) is followed with the Sutherland coefficient, = 6.
tel-00708600, version 1 - 15 Jun 2012
Table 2. Calculations for the composition of the catalytic systems. Column 1- Name of catalyst, Column 2- Average
RuNP diameter measured by TEM, Column 3- Calculated dispersion, Column 4- Initial Ru(COD)(COT)
concentration, Column 5- Consequent Rus concentration, Column 6- Volume of IL/RuNP solution ,Column 7-
Volume of pure IL added, Column 8- Consequent number of moles of Rus in the 5 mL mixture, Column 9- mass of
IL, Column 10- mass of substrate, Column 11- Substrate-Catalyst ratio, Column 12- Substrate-IL ratio.
1 2 3 4 5 6 7 8 9 10 11 12
Ru0 1.1 82% 43.0 35.2 2.42 2.58 9.52 7.0 0.78 105 0.59
Ru25 2.3 53% 43.0 22.9 3.74 1.26 9.52 7.0 0.78 105 0.59
Ru50 2.9 43% 43.0 18.5 5.00 - 9.52 7.0 0.78 105 0.59
In Table 2, are collected all data concerning the hydrogenation of CYD. In columns 1-3,
size and dispersion values of (Ru0), (Ru25), (Ru50) are reported. Given the concentration of NP
precursor (column 4) and the size of NPs generated (column 2), it is possible to calculate the
concentration of Rus for Ru0, Ru25 and Ru50 (column 5). The more concentrated solutions of Rus
(Ruo and Ru25) can be easily diluted to match the least concentrated (Ru50) by simple addition of
the appropriate amount of pure IL (columns 6 and 7), resulting in identical quantities of Ru s
167
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
(column 8). Identical concentrations of substrate may then be used, below the limit of solubility
to avoid demixing (10 % wt) (columns 9 and 10), permitting as a result both a constant substrate-
catalyst ratio (column 11) and a constant substrate-IL ratio (column 12) hence constant viscosity.
Using solutions prepared as described, the reaction was carried out in parallel in several
0.5 mL batches under 1.2 bars of pure molecular hydrogen, which were stirred and heated with
the aid of a thermostatic carousel, ensuring identical reaction conditions.
4.5.1 1,3-cyclohexadiene
It can be seen in Table 3 that the largest NPs (2.9 nm) are the most active in the
tel-00708600, version 1 - 15 Jun 2012
In a similar fashion, the three sizes of RuNP were tested in the catalytic hydrogenation of
other substrates to determine whether this size-related activity could be generalised.
4.5.2 Cyclohexene
To establish whether or not a difference in activity is also apparent in the case of a
monoene, the hydrogenation of cyclohexene (CYE) to cyclohexane (CYA) was studied. The
results tabulated in Table 4 show that the larger NPs are indeed more active, however the
difference in activity is less substantial than in the case of the conjugated diene CYD, and falls
within experimental error.
168
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
4.5.3 Limonene
The naturally occurring terpene, (R)-(+)-limonene, or 1-methyl-4-isopropenyl-1-
cyclohexene, Lim, was also chosen as a substrate. This molecule is also a non-conjugated diene
as can be seen in Scheme 2.
tel-00708600, version 1 - 15 Jun 2012
Scheme 2. (R)-(+)-limonene.
Kinetic tests for the hydrogenation of Lim were carried out in the same way as those for
CYD, using this time only 5 % wt. of substrate with respect to the IL, due to its lower solubility
in the medium. Mixture compositions are given in Table 5.
Table 5. Compositions of the catalytic systems for the hydrogenation of limonene and conversion after 120 minutes
Conversion
Vol.
[Ru]/ [Rus]/ Vol. IL- Ru s /
IL mass
RuNP mass at 120
d/nm D mmol mmol solution/ pure/ 10-5 Lim/ Lim/Rus Lim/IL
IL/ g mins
g
L-1 L-1 mL moles
mL %
Ru0 1.1 82% 43.0 35.2 2.42 2.58 9.52 7.0 0.37 28 0.16 90%
Ru25 2.3 53% 43.0 22.9 3.74 1.26 9.52 7.0 0.37 28 0.16 88%
Ru50 2.9 43% 43.0 18.5 5.00 - 9.52 7.0 0.37 28 0.16 92%
At 30 °C, after 2 hours of reaction ~ 90 % of Lim was converted in each case, Table 5.
Here, no remarkable difference in activity is apparent.
169
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
4.5.4 Styrene
The hydrogenation of styrene, Sty, affords either a partially (ethylbenzene, EBn) or fully
hydrogenated product (ethylcyclohexane, ECYA) (Scheme 3). Generally, the hydrogenation of
the olefinic position is faster than the aromatic ring.
tel-00708600, version 1 - 15 Jun 2012
As for CYD, CYE and Lim, kinetic tests for the hydrogenation of Sty were carried out
using the three differently sized RuNPs produced in the ILs. Mixture compositions and
conversions are given in Table 6.
Table 6. Compositions of the catalytic systems for the hydrogenation of styrene and conversion after 120 minutes.
Vol. IL- Vol.
[Ru]/ [Rus]/ Ru s / mas Conversion
d/ RuNP IL mass Sty Sty/
D mmol mmol 10-5 s IL/ at 120
nm solution/ pure/ Sty/ g /Rus IL
L-1 L-1 moles g mins%
mL mL
Ru0 1.1 82% 43.0 35.2 2.42 2.58 9.52 7.0 0.78 81 0.45 60 %
Ru25 2.3 53% 43.0 22.9 3.74 1.26 9.52 7.0 0.78 81 0.45 60 %
Ru50 2.9 43% 43.0 18.5 5.00 - 9.52 7.0 0.78 81 0.45 63 %
At 30 °C, the hydrogenation of the external double bond in each case was rapid, leading
to 70 % conversion after 2 h in all cases. Selectivity for EBn remained at 100% for all sizes of
NP. As was the case for limonene and CYE, at 30 °C, no significant size effect was observed.
4.5.5 Comparison of the substrates
According to literature results, the catalytic activity of NPs depends on their size, and
generally reaches a maximum for those of around 3 nm.29,30 Here, only a size effect on activity
170
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
was observed for the case of CYD. The fact that the hydrogenation of CYD is faster with larger
NPs can be related to two factors:
1) Larger NPs present the appropriate number of neighbouring surface sites to facilitate
the -bond activation of the conjugated system.31
2) Through this -bonding activation, similarly to benzene, 1,3-cyclohexadiene would
lose part of its resonance energy and react more readily.20
The coordination of monoenes such as CYE or the olefinic bond of styrene does not
necessitate large surfaces, explaining the less pronounced size effect in these cases.
In the case of limonene, unlike in CYD, the double bonds are not conjugated and so a
planar π-coordination to disrupt conjugation is not energetically advantageous, hence the result is
similar to that of a monoene.
tel-00708600, version 1 - 15 Jun 2012
4.6.1 1,3-cyclohexadiene
100%
1.1 nm
95%
2.3 nm
Selectivity for CYE
90% 2.9 nm
85%
80%
75%
70%
0% 20% 40% 60% 80% 100%
Conversion
Figure 7. Selectivity for cyclohexene as a function of conversion for the three different sizes of RuNPs
171
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
In the hydrogenation of CYD, CYE is obtained as the major product. Interestingly, the
selectivity for CYE diminishes with increasing NP size. Indeed, in the case of Ru0, selectivity for
CYE is 100% at low conversion and only slightly diminishes at high conversion (97%). In
contrast, for Ru50 the hydrogenation is unselective even at low conversion (Figure 7).
Assuming highly crystalline particles with hcp structure, a particle of diameter 1.1 nm
would have the vast majority of catalytic surface atoms occupying vertex or edge positions. Such
vertex ruthenium atoms Ruv, which under H2 atmosphere are ligated by hydrides,32 may
coordinate one C=C double bond of CYD. The product of the subsequent hydrogenation is CYE,
which must undergo a second coordination to give the fully hydrogenated CYA. Similarly, for a
larger particle of average diameter 2.9 nm assuming high crystallinity and a hcp structure, it is
evident that most of the catalytically active surface ruthenium atoms are found in facial
tel-00708600, version 1 - 15 Jun 2012
positions, Ruf. Indeed, here such crystallinity has already been observed by HREM, Figure 1. Ruf
may hydrogenate the olefin using the mechanism previously discussed, but due to the planar
arrangement of Ruf, another mechanism may be envisaged involving the double coordination of
the diene, as generally found during the hydrogenation of 1,3-cyclohexadiene on metallic
surfaces,33,34 and thus rapid consecutive or simultaneous hydrogenation of both double bonds
leading to the fully hydrogenated CYA may be possible. In Figure 8 are represented simplified
SYBYL models of CYD molecules coordinating to the surface of perfectly crystalline RuNPs of
calculated average diameter 1.3 nm and 2.8 nm. This illustrates nicely the greater facility of
planar coordination to faces of the larger NPs. Through this -bonding activation similarly to
benzene, 1,3-cyclohexadiene would lose part of its resonance energy and react more readily. This
could explain the lower selectivity of the larger RuNPs despite identical reaction conditions.
Likewise, in the hydrogenation of 1,3-butadiene or 1-hexyne the selectivity of small NPs towards
1-butene or 1-hexene versus butane or hexane is still higher than that of larger NPs.31, 33-35 In this
work, in the hydrogenation CYD the selectivity in CYE versus CYA drops from 97% to 80%
when the RuNP size increases from 1.1 to 2.9 nm.
172
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
tel-00708600, version 1 - 15 Jun 2012
Figure 8. SYBYL representations of CYD coordinated to the face of highly crystalline RuNPs of mean diameter
1.3 nm (left) and 2.8 nm (right)
4.6.2 Limonene
Limonene, containing two double-bonds, both inside and outside of the cyclohexene ring,
may also yield different products, as depicted in Scheme 4. Complete hydrogenation would yield
mixture of cis- and trans-p-menthane, c and d, whereas partial hydrogenation may result in one
of the intermediates p-1-menthene or 1-methyl-4-isopropenylcyclohexane, a and b. Several
groups have reported however, that hydrogenation of the external double bond is preferential, so
only p-1-menthene is detected as a partially hydrogenated product of this reaction.36
173
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
When the catalytic hydrogenation was performed at 30 °C, high selectivity for the
tel-00708600, version 1 - 15 Jun 2012
partially hydrated product menthene was observed, although both menthanes are formed slowly
over time. As for CYD hydrogenation the selectivity was the poorest in the case of Ru50. Unlike
for CYD, 100% selectivity is only observed at very low conversion (≤ 20 %) with Ru0 (Figure
9).
Figure 9. Selectivity for p-1-menthene at 30°C as a function of conversion for the three different sizes of RuNPs.
In an attempt to increase the selectivity, the same reaction was therefore performed at low
temperature (-2 °C). For all sizes of nanoparticle, the selectivity for menthene is increased
174
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
compared to when carried out at ambient temperature. Interestingly the trend of selectivity
although poor at the start of the reaction improved between 10 – 20 % conversion, Figure 10.
This phenomenon must be due to the start of a biphasic system, thus leading to extraction of the
product, inducing a better selectivity as already reported in the hydrogenation of butadiene to
butenes,24 and benzene to cyclohexene.25
For the hydrogenation of limonene which contains two non-conjugated double-bonds, the
selectivity is NP size-dependent at 30 °C, 100 % selectivity for p-1-menthene being observed at
low conversion for the smallest NPs. At -2 °C, decreasing the solubility of both substrate and
products affords a higher selectivity in all cases, mainly due to the formation of biphasic media.
tel-00708600, version 1 - 15 Jun 2012
Figure 10. Selectivity for p-1-menthene at -2 °C as a function of conversion for the three different sizes of RuNPs.
4.6.3 Styrene
It is often reported that the activity and the selectivity in aromatic hydrogenation are size
and face dependant.4,6,13
A common probe for chemo-selectivity in the hydrogenation of aromatics is styrene, Sty.
Generally, the hydrogenation of the olefinic double bond is faster than the aromatic ring and
occurred at 30 °C with 100 % selectivity for EBn, vide supra. This is because a greater energy
barrier must be overcome when hydrogenating an aromatic compared to an olefin. Therefore to
175
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
force the hydrogenation of the aromatic ring, the same reaction must be carried out at higher
temperature and for a longer duration, here 75 °C during a period of 6 h.
Once again, the hydrogenation of the external double bond was rapid and occurred
quantitatively for each size of NP, Figure 11. Only at 100 % conversion was hydrogenation of
the aromatic cycle observed, indicating a successive hydrogenation of the different functions.
tel-00708600, version 1 - 15 Jun 2012
Figure 11. Hydrogenation of Sty at 75 °C under 1.2 bar H2 in [C1C4Im][NTf2], for three different sized RuNPs.
In Figure 12 is depicted the yield in ECHA with respect to time. Here it can be clearly
seen that the full hydrogenation of the aromatic occurs more readily on the large NPs, in
agreement with results for the hydrogenation of CYD. Here, a planar coordination is even more
crucial, as the inert aromatic ring must lose part of its resonance energy to be activated.20
Using styrene as a probe has emphasised the importance of planar π-coordination in
aromatic hydrogenation as the larger NPs fully reduced styrene to ethylcyclohexane more readily
than the smaller NPs
176
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
tel-00708600, version 1 - 15 Jun 2012
Figure 12. Hydrogenation of Sty at 75°C under 1.2 bar H2 in [C1C4Im][NTf2], for three different sized RuNPs.
It is also interesting to note that during the course of the reaction small amounts of 1-
ethylcyclohexene were produced, and identified by GC-MS: as much as 0.5 % for Ru0, by the
end of the reaction. Indeed the partial hydrogenation of aromatic cycles is an interesting and
ongoing challenge and has already been reported in IL media for benzene.25 It may be interesting
in the future to optimise conditions to favour the production and extraction of this molecule from
the medium.
177
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
adsorption of H2, in agreement with observations of labile surface hydrides on the NP surface,32
but the mobility and absorption of the substrate.21
Table 7. Results for the catalytic hydrogenation of CYD using Ru 0 and Ru0 under 1.2 bars and 4 bars H2
Pressure H2 / Conversion at 90 Selectivity
TON TOF / h-1
bar min CYE
4.8 Recycling
tel-00708600, version 1 - 15 Jun 2012
TEM images of the reaction media after hydrogenation of CYD show a reorganisation of
the NPs, probably due to the presence of organic substrates, poorly soluble in the IL. For Ru0 and
Ru25, circular patterns are present, possibly caused by the formation of aggregates, where the
NPs would decorate the organic-IL interface Figure 13. In the case of Ru50 phase separation is
apparent, which may be indicative of a larger amount of CYE and CYA present as a result of the
higher activity and lower selectivity. In all cases, the mean RuNP size measured Table 8 does not
differ greatly from the original size, in accordance with the stability of RuNPs in ILs under
molecular hydrogen,32 however the size distribution is larger, probably as an effect of stirring.15
Figure 13. Transition electron micrographs of RuNPs after CYD hydrogenation for Ru0 (left), Ru25 (middle) and
Ru50 (right)
178
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
The apparent resistance to coalescence of the NPs means that they may be tested for their
recyclability. Subsequently, using the most selective catalyst, Ru0, recycling experiments were
performed, by extracting in vacuo and quantifying the volatiles after each 90 minute run. More
CYD was then added for hydrogenation. From the results (Table 8 and Figure 14) it can be seen
that the activity and selectivity both remain high after 5 recycles, diminishing only slightly with
each run. This small decrease of course is attributable to the gradual coalescence of the NPs,
leading to a diminution in the number of active surface sites and larger, less selective NPs.
Indeed, TEM images obtained of the NPs after all recycling experiments showed that the NPs
had undergone coalescence to attain an average diameter of 1.8 ± 0.5 nm, approaching the size of
Ru25, and of course exhibiting a similar selectivity.
tel-00708600, version 1 - 15 Jun 2012
Table 8. Results from recycling experiments in the hydrogenation of CYD using Ru0 and comparison with results
obtained for Ru25 and Ru50
Pressure H2 / Conversion at 90 Selectivity Size after
TON TOF / h-1
bar min CYE catalysis / nm
179
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
55 100
TOF
Selectivity for cyclohexene
50
45
80
-1
TOF / h
40
70
35
60
30
tel-00708600, version 1 - 15 Jun 2012
25 50
1 2 3 4 5 6
Cycle
4.9 Conclusion
180
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
For the hydrogenation of Lim, as for Sty, the effect of NP size on activity is not
significant. However, the selectivity is NP size-dependent at 30 °C, 100 % selectivity for p-1-
menthene being observed at low conversion ( 20 %) for the smallest NPs – Ru0.
Conversely, it was found that for catalytic hydrogenation of CYD, the activity of catalyst
increases with the NP size in agreement with the literature results on heterogeneous catalysts.
Contrarily to activity, in the hydrogenation of CYD, the selectivity for CYE versus CYA drops
from 97 % to 80 % when the RuNP size increases from 1.1 to 2.9 nm.
Both results, activity and selectivity, are in agreement with a mechanism involving a π-
bond activation and a double coordination of diene substrates, necessitating several neighbouring
surface atoms only found in facial positions on the larger NPs. Furthermore, these RuNPs show a
high level of recyclability with neither loss of activity nor significant agglomeration.
tel-00708600, version 1 - 15 Jun 2012
4.10 Outlooks
A commonly reported advantage to generating NPs in ILs is that the resulting surfaces
are ligand free, and therefore the potential in catalysis is maximised. However, certain ligands
may have an important electronic effect on the catalyst. In our laboratory, we have recently
reported a synthesis of RuNPs involving amine ligands in ionic liquids, giving very small NPs
(1.2 nm) with a narrow size distribution, active in aromatic hydrogenation at 75 °C without
coalescence.38 It is planned therefore to compare the catalytic performance, i.e. activity,
selectivity and recyclability of ligated and ligand-free NPs.
The partial hydrogenation of aromatics is a challenging issue. The complicated Asahi
process, involving a tetraphasic mixture incorporating colloidal ruthenium as the catalyst, is the
only industrial process, and can convert benzene to CYE with a yield of 60%.39 In this work and
in similar systems,25 the partial hydrogenation of aromatics has been reported, using the
solvation properties of ILs to extract partially hydrogenated species. Future work is planned to
firstly determine whether these species are a result of partial hydrogenation or dehydrogenation
of fully hydrogenated species, and then attempt to optimise conditions, for instance using wet
ILs, to selectively extract the desired product in higher yields.
181
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
182
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
the “TEMSCAN” centre of the Université Paul Sabatier Toulouse 3, Toulouse, France, using a
JEOL JEM 200CX electron microscope with acceleration voltage of 200 kV.
4.11.4 X-ray Photoelectron Spectroscopy (XPS)
X-ray photoelectron spectroscopy was performed in a Kratos Axis Ultra DLD
spectrometer, using a monochromatic AlKα X-ray with a pass energy of 20 eV and a coaxial
charge neutraliser. The base pressure in the analysis chamber was better than 5×10−8 Pa. XPS
spectra of Ru3p, C1s, Si2p and O1s levels were measured at a normal angle with respect to the
plane of the surface. High resolution spectra were corrected for charging effects by assigning a
value of 284.6 eV to the C1s peak (adventitious carbon). Binding energies were determined with
an accuracy of ± 0.2 eV. The data were analysed using Casa-XPS (v 2.3.13) employing a Shirley
background subtraction prior to fitting and a peak shape with a combination of Gaussian and
tel-00708600, version 1 - 15 Jun 2012
Lorentzian (30% Lorentzian). High resolution spectra were acquired in the region of Ru 3p as the
Ru 3d region overlaps with the C 2p region of the residual ionic liquid.
4.11.5 Catalytic tests
Catalytic solutions were made in a glove box and left stirring for 12 h in a closed system
to ensure homogenity. 0.5 mL aliquots were transferred to identical Schlenk tubes containing
cross-shaped magnetic stirrer bars. The argon atmosphere was removed and the solution
degassed in vacuo whilst cooling in liquid nitrogen (−196 °C). For reactions at 30 °C, 6 of these
Schlenk tubes were placed in a thermostatic carousel to ensure identical temperature and stirring
conditions. After 30 minutes, when the temperature had stabilised, the Schlenk tubes were
opened to 1.2 bars of H2. After t minutes, a Schlenk tube was isolated and opened to air,
releasing the H2 atmosphere thus stopping the reaction. The solution was entirely dissolved in
10 mL of acetonitrile containing a 1 molar concentration of toluene. The composition of the
mixture was determined by gas phase chromatography using toluene as the internal standard. For
reactions at 75 °C the same procedure was used as above. For reactions at -2 °C, the Schlenk
tubes were cooled independently in a cryostatic bath and stirred. After 30 minutes, when the
temperature had stabilised, the Schlenk tubes were opened to 1.2 bars of H2. After t minutes, a
Schlenk tube was isolated and opened to air, releasing the H2 atmosphere thus stopping the
reaction. The solution was entirely dissolved in 10 mL of acetonitrile containing a 1 molar
concentration of toluene. The composition of the mixture was determined by gas phase
chromatography using toluene as the internal standard.
183
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
0.25 μm film thickness. The carrier flow (He) was maintained at 1.2 mL min-1. The column
temperature programme was as follows: initial temperature 70°C, maintained for 2 min, then
ramped at 15 °C min-1 to 310 °C, which was maintained for 10 min.
4.11.8 Density and viscosity
The mixtures of and CYD at different compositions were prepared gravimetrically
following the procedure already described.[27] The viscosity of the mixture was measured at
298.15 K (controlled to within ± 0.005 K and measured with the accuracy better than ± 0.05 K)
using a rolling ball viscometer from Anton Paar, model AMVn.[27] The overall uncertainty of
the viscosity is estimated as ± 2.0 %. The densities of the mixtures, necessary to calculate the
viscosities were measured in an Anton Paar vibrating tube densimeter model 512 P, at 298.15 K
(measured by a calibrated PRT with an accuracy of ± 0.02 K). The overall uncertainty of the
density is estimated as ± 0.01 %.
4.11.9 Solubility
To measure the solubility, 1 mL of the substrate was stirred with the ionic liquid in a
closed system at 298.15 K for 12 h then left to settle for a further 2 h. A 0.1 mL sample of the
ionic liquid phase was weighed and its composition determined by GC using the procedure
described in section 4.11.6. Tests were repeated 4 times for each substrate, to guarantee
reproducibilty.
184
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
4.12 References
185
Chapter 4 NP catalysis in Imidazolium ILs – The influence of size on catalytic performance
27. T. Umecky, Y. Saito and H. Matsumoto, Electrochem. Soc. Trans., 2010, 25, 23-29.
28. M. A. Vorotyntsev, V. A. Zinovyeva and M. Picquet, Electrochim. Acta 2010, 55, 5063-
5070.
29. D. Y. Murzin, Chem. Eng. Sci., 2009, 64, 1046-1052.
30. A. Binder, M. Seipenbusch, M. Muhler and G. Kasper, J. Catal., 2009, 268, 150-155.
31. N. Semagina, A. Renken and L. Kiwi-Minsker, J. Phys. Chem. C, 2007, 111, 13933-
13937.
32. P. S. Campbell, C. C. Santini, D. Bouchu, B. Fenet, K. Philippot, B. Chaudret, A. A. H.
Padua and Y. Chauvin, Phys. Chem. Chem. Phys., 2010, 12, 4217-4223.
33. M. Saeys, M.-F. Reyniers, M. Neurock and G. B. Marin, Surf. Sci., 2006, 600, 3121-
3134.
34. W. L. Manner, G. S. Girolami and R. G. Nuzzo, J. Phys. Chem. B, 1998, 102, 10295-
10306.
35. J. Silvestre-Albero, G. Rupprechter and H.-J. Freund, J. Catal., 2006, 240, 58-65.
36. E. Bogel-Lukasik, I. Fonseca, R. Bogel-Lukasik, Y. A. Tarasenko, M. Nunes da Ponte,
A. Paiva and G. Brunner, Green Chem., 2007, 9, 427-430.
37. A. Ahosseini and A. M. Scurto, Int. J. Thermophys. , 2008, 29, 1222-1243.
38. G. Salas, Santini C. C., K. Philippot, V. Collière, B. Chaudret and B. Fenet, Dalton
Trans., submitted 2010.
39. H. Nagara, Appl. Surf. Sci., 1997, 121/122, 448-451.
186
tel-00708600, version 1 - 15 Jun 2012
General Conclusion
187
tel-00708600, version 1 - 15 Jun 2012
188
General Conclusion
General Conclusion
This thesis focuses on the peculiar physical-chemical properties of imidazolium based ionic
liquids – more precisely, how we may exploit these properties in catalysis and how they may
influence the outcome of the catalytic process. In this respect, the following two important
factors were predominantly investigated:
1. The specific solvation properties of ionic liquids (ILs) – How may these be utilised to
generate and stabilise in situ catalytically active metal nanoparticles (NPs) of
controlled size?
2. The distinctive thermophysical properties of ILs (viscosity and diffusivity) – What is
the influence on catalytic activity in both homogeneous and NP catalysis performed
tel-00708600, version 1 - 15 Jun 2012
in situ?
First of all, a review of the most significant literature results in the synthesis and catalytic
applications of ruthenium nanoparticles (RuNPs) highlighted the dependence of the catalytic
properties on their size and surface properties. The ability to predict and control the size of the
NPs produced, is therefore of the utmost importance. Furthermore, producing NPs free of ligands
and other surface contaminants can only be advantageous in catalysis. Decomposition of
organometallics under H2 presents the advantage of producing only inert alkane by-products,
easily removed under vacuum, while the use of ILs produces stable NPs of controlled size
without the need for stabilising additives. Marrying these two techniques leads to a route to
stable NPs of controlled size, with maximum surface availability. Consequently, throughout this
work, NPs are produced by the decomposition of organometallic complexes under H2 in ILs.
It has previously been explained that the phenomenon of crystal growth of RuNPs,
generated in situ in imidazolium-based ILs from Ru(COD)(COT), is controlled by the size of
non-polar domains created by the grouping of lipophilic alkyl chains and segregated by a rigid 3-
D network of ionic channels. However, no information was available to explain the remarkable
stability of the resulting systems. Is the RuNP surface really contaminant-free?
In this work, we have brought to light firstly the close proximity of the RuNPs to the non-
polar R group of the cation, through in situ labelling experiments (deuteration and reduction),
thus corroborating the hypothesis that the RuNPs are located within these non-polar pockets.
Addition of water, destroying the ionic network, induces agglomeration of the RuNPs,
189
General Conclusion
supporting the fact that the RuNPs are originally isolated within non-polar pockets by the 3-D
network of ionic channels. Secondly, the presence of surface hydrides and their role in the
stabilisation of the nanoparticles is demonstrated. The stabilisation mechanism is thus analogous
to that described in colloidal organic solutions.
A
HO N
OH N
Me Me methanol A
N
R N
H
A
R
H N
R
H
N
H
Me Me OH THF
N N
H R A
N
R H A
HO A
N
N
A
H by analogy N
N
RH R
H
N
H N
N
R H H
H H N
A A
R
A A
N N
R N
N
N RH R A
N
N N N
H H
R
H A
R H
cyclooctane H A
H H N R N
N
A
N
Literary results give conflicting views on the mechanism for the growth and stabilisation
tel-00708600, version 1 - 15 Jun 2012
of transition metal nanoparticles in ionic liquids, leading to the conclusion that this depends on
the nature of the metal, its precursor complex, the synthetic route and/or the ionic liquid used.
This led to the question of whether the same would be true of NPs generated in the same way
from an analogous organometallic compound of another metal. For this reason, we have also
investigated the possibility of generalising these findings to the generation of nickel
nanoparticles, NiNPs, as the organometallic complex bis(1,5-cyclooctadiene)nickel, Ni(COD)2,
analogous to Ru(COD)(COT), is commercially available.
The synthesis of NiNPs from Ni(COD)2 under H2 (4 bar) was carried out in various
imidazolium based ionic liquids at 0 °C and 25 °C. Imidazolium cations substituted with longer
chains [C1C6Im][NTf2], [C1C8Im][NTf2] and [C1C10Im][NTf2], were found to offer fairly good
media for the preparation and stabilisation of nickel nanoparticles. NiNP size increased with
increasing chain-length and in this way we were able to control NiNP size. Additionally, upon
decreasing the reaction temperature to 0 °C, we were able to produce smaller NiNPs than those
produced in the same IL media at 25 °C. The presence of surface hydrides on these NiNPs, as on
RuNPs, has been evidenced by hydrogenation of ethylene in the absence of hydrogen. The
resulting NiNPs in these ILs were found to be larger than RuNPs produced in the same media.
This can be largely attributed to differences in the nucleation and growth processes for the two
metals.
Unexpectedly, spontaneous decomposition of Ni(COD)2 occurred without the addition of
hydrogen upon dissolution into imidazolium ILs with short alkyl chains; [C1C2Im][NTf2],
190
General Conclusion
Bu + Bu 1,3-COD
+ 2 (1,5-COD)
N N
N
A- N
Bu Me
C2
A- + Ni(1,5-COD)2 C2 [Ni] H N
H
Me
N
H 1,5-COD
Me
tel-00708600, version 1 - 15 Jun 2012
C2H4
?
Bu N + N N + N
Bu Me Bu Me
N C2 C2
+ H A -
A-
C Et Bu
N 2 H
Me = Ni
191
General Conclusion
(1) (2)
(1) (2)
IL1
IL1
IL2
IL2
rate α η-1
192
General Conclusion
In summary, ILs have been shown to be excellent media for the controlled growth and
stabilisation of metal NPs, and the importance of the size control in catalysis has been proven.
193
General Conclusion
Outlooks
During this work, the synthesis and stabilisation of size-controlled RuNPs in imidazolium
ionic liquids has been well established and understood. Their use in catalysis has been touched
upon using some model reactions, during which, the partial hydrogenation of styrene giving
ethylcyclohexene was observed. Although the yield was poor, this is nonetheless an interesting
result, as the partial hydrogenation of aromatic cycles to cyclic olefins is an important and
challenging industrial process. By investigating further this reaction, in order to understand the
mechanism, it may be possible to optimise conditions in order to increase the yield of this
interesting product. Is this product a result of partial hydrogenation, or full hydrogenation
followed by dehydrogenation? Can we extract this product by using additive containing ILs (e.g.
tel-00708600, version 1 - 15 Jun 2012
H2O to alter the thermophysical properties and the solubility or organic moieties)?
It may also be interesting to use these well defined catalytic systems to attempt more
interesting catalytic reactions, such as alkane oxidations or nitrogen activation.
Finally, some important and interesting preliminary results have been obtained regarding
the possiblility of producing zero-valent NPs of oxophilic metals, such as tantalum. This opens
up new innovative pathways to produce nano-objects with reduced process cost and could be of
great potential interest in the drive for smaller and more efficient electronic devices. This is
currently under development in collaboration with the “Commissariat à l’Énergie Atomique
(CEA)”, and is the subject of two Ph.D. theses.
194
i
tel-00708600, version 1 - 15 Jun 2012
Appendix 1
Characterisation of Nickel
Nanoparticles by XPS
195
tel-00708600, version 1 - 15 Jun 2012
196
Appendix 1 Characterisation of Nickel Nanoparticles by XPS
atmosphere and eliminating as much IL as possible. The XPS spectrum in the Ni 2p region for
NPs prepared by decomposition under H2 is shown in Figure 1. It is clear that two species exist,
Ni0 and NiII.
The data gathered during XPS analyses are tabulated in Table X. It can be seen that in
both cases a large peak is present corresponding to NiII. It should be noted that when transferring
the sample into the sample treatment chamber, it was subjected to air and ambient conditions for
a short period of time (< 5 minutes). The high level of oxidation may be a result of air exposure.
For this reason we rerun the analyses of the samples after a longer exposure period in air
197
Appendix 1 Characterisation of Nickel Nanoparticles by XPS
(~ 5 hours). In the case of NiNPs produced through the H2 decomposition of Ni(COD)2, the
intensity of the Ni 2p bands corresponding to NiII was greatly increased after prolonged air
exposure, where as for the NiNPs produced from spontaneous decomposition, the amount of Ni II
remained constant. This could be explained by two factors; 1) the NiNPs produced by
spontaneous decomposition exhibit less Ni surface susceptible to oxidation; 2) these NiNPs were
already oxidised at the surface by oxidative addition of the imidazolium ring producing carbene
ligands.
Table 1. XPS data gathered for NiNPs produced by autodecomposition and H 2 decomposition of Ni(COD)2 in the
IL [C1C6Im][NTf2]
Sample Ni 2p peak 2p 3/2 2p 1/2 Doublet Area Ratio
separation
Ni(0):NiII
tel-00708600, version 1 - 15 Jun 2012
The results obtained are inconclusive, as we cannot be sure whether the oxidised nickel
was due to air exposure or if it already existed in the NiNP/IL mixture. Future tests are planned
to run XPS using a system for transferring the sample into the chamber without exposure to air.
Magnetic measurements using a sensitive SQUID (Superconducting Quantum
Interference Device) may also be useful in characterising the nickel nanoparticles. FC-ZFC (field
cooled – zero field cooled) measurements of the magnetisation can be used to find the blocking
temperature (Tb) of the superparamagnetic NPs, which may be related to their size. The
percentage of non-oxidised Ni0 may also be estimated from the magnetic hysteresis curve at
obtained at low temperature (generally 2 K). This work is currently in progress.
198
tel-00708600, version 1 - 15 Jun 2012
Publications
ii
Appendix 2
199
tel-00708600, version 1 - 15 Jun 2012
200
Appendix 2 Publications
Publications
The work reported in this thesis has contributed to the following papers and patents. Copies of
the papers can be found in this appendix.
Papers
1) “A novel stabilisation model for ruthenium nanoparticles in imidazolium ionic liquids: in situ
spectroscopic and labelling evidence.”
3) “Olefin hydrogenation by ruthenium nanoparticles in ionic liquid media: Does size matter?”
4) “Imidazolium ionic liquids as promoters and stabilising agents for the preparation of metal(0)
nanoparticles by reduction and decomposition of organometallic complexes.”
Patents
1) “Mémoires et Interconnections à base de Nanotubes de Carbone.” Catherine Scampucci (ep.
Santini), Jean-Marie Basset, Thibaut Gutel, Paul Campbell, Simon Deleonibus, Paul Haumesser,
Brevet CEA-LETI /Univ Lyon 1 CNRS/LCOMS : Fr 0901464 27/03/09
2) “Procédé de réalisation d’un dispositif mémoire à grille flottante.” Catherine Scampucci (ep.
Santini) , Jean-Marie Basset, Thibaut Gutel, Paul Campbell, Simon Deleonibus, Paul Haumesser,
Brevet CEA-LETI /Univ Lyon 1 CNRS/LCOMS : Fr 0901463 27/03/09
201
tel-00708600, version 1 - 15 Jun 2012
202
Appendix 2
Publications
PAPER www.rsc.org/pccp | Physical Chemistry Chemical Physics
In situ labelling and spectroscopic experiments are used to explain the key points in the stabilisation
of ruthenium nanoparticles (RuNPs) generated in imidazolium-based ionic liquids (ILs) by
decomposition of (Z4-1,5-cyclooctadiene)(Z6-1,3,5-cyclooctatriene)ruthenium(0), Ru(COD)(COT),
under dihydrogen. These are found to be: (1) the presence of hydrides at the RuNP surface and, (2)
the confinement of RuNPs in the non-polar domains of the structured IL, induced by the rigid 3-D
organisation. These results lead to a novel stabilisation model for NPs in ionic liquids.
Metal nanoparticles (MNPs) have lately become a new and include ligand-type stabilisation by the IL moieties. Indeed,
exciting area of research interest. Defined as particles, whose this has been illustrated for IrNPs in an imidazolium based IL,
dimensions are on the nanoscale, they coincide with a transition where H/D labelling experiments provided evidence for
between bulk and molecular states, and exhibit unique charac- the formation of N-heterocyclic carbenes on the surface of
teristics related to the discontinuities and quantum effects of the NPs.8,9
such a transition, e.g. enhanced magnetic and catalytic proper- Very recently, a relationship between the size of IL
ties, etc. The controlled synthesis of nanoparticles in the range non-polar domains calculated by molecular dynamics simu-
of 1–10 nm is still an ongoing challenge, as is the under- lation and the mean diameter of ruthenium nanoparticles
standing of their stabilisation and agglomeration.1–4 Indeed, (RuNP), (generated in situ from the organometallic complex
transition-metal nanoparticles remain only kinetically stable, (Z4-1,5-cyclooctadiene)(Z6-1,3,5-cyclooctatriene) ruthenium(0),
the thermodynamic minimum being bulk metal. Consequently, Ru(COD)(COT), has been found.10 This suggests that the
substantial effort has been centred on stabilising transition- phenomenon of crystal growth occurs in these non-polar
metal nanoparticles. Furthermore, their formation and stabilisation domains of the ILs and is controlled by the local concentration
(inhibition of coalescence) are closely related issues.3 of the precursor complex. This can be supported by the fact
In ionic liquids (ILs), particularly those based on the that imidazolium-based ILs exhibit an extended hydrogen-
imidazolium cation, NPs can be generated by several physical bond network in the liquid state and a continuous three
and chemical routes.5 Unlike traditional solvents, IL media dimensional network of ionic channels, coexisting with non-
are able to stabilise nanoparticles in the absence of further polar domains created by the grouping of lipophilic alkyl
additives such as ligands, inhibiting metal agglomeration to chains.11–13 We can thus hypothesise that the RuNPs are
the bulk.6 However, the question of precisely how ILs stabilise surrounded by these non-polar pockets.
transition metal NPs remains under debate. Likewise, some of us have reported ‘‘solvent-only’’ stabi-
Many papers have claimed possible electrostatic stabilisation lised nanoclusters prepared from the anion-free precursor
by interactions between nanoparticles and weakly coordinating Ru(COD)(COT) in THF–MeOH.14 In this work, ‘‘nano-
reactors’’ or ‘‘pockets’’ are hypothesised to be formed with
a
Universite´ de Lyon, Institut de Chimie de Lyon, cyclooctane (the by-product of the decomposition) in which
UMR 5265 CNRS-Universite´ de Lyon-ESCPE Lyon,
LC2P2, Equipe Chimie Organométallique de Surface, ruthenium is confined and the particles stabilised, Scheme 1a.15
ESCPE 43 Boulevard du 11 Novembre 1918, F-69616 Villeurbanne, Moreover, the same authors have proved the presence and the
France. E-mail: santini@cpe.fr; Fax: 33(0)472431795; stability of hydrogen atoms on the surface of RuNPs.16,17
Tel: 33(0)472431794 By analogy, novel stabilising factors for RuNPs in ILs are
b
Centre Commun de Spectrométrie de Masse, UCB Lyon 1-ESCPE
Lyon, 43 Boulevard du 11 Novembre 1918, F-69616 Villeurbanne, reported here, such as the presence of surface hydrides, and
France their encapsulation and segregation in the non-polar domains
c
Centre Commun de RMN, UCB Lyon 1-ESCPE Lyon, of the structured IL induced by the rigid 3-D organisation,
43 Boulevard du 11 Novembre 1918, F-69616 Villeurbanne, France
d
Laboratoire de Chimie de Coordination, UMR CNRS, 205, Scheme 1b. It is worth noting that ‘‘stabilising factors’’ are
route de Narbonne, F-31077-Toulouse cedex 04, France often claimed based on ex situ methods such as TEM images
e
Laboratoire de Thermodynamique des Solutions et des Polymères, and XPS experiments performed on isolated samples. In
Universite´ Blaise Pascal, Clermont-Ferrand, 24 av. des Landais, contrast, in this work, only in situ labelling methods and
63177 Aubie`re, France
w Electronic supplementary information (ESI) available: Mass spectra spectroscopic measurements made directly in the reaction
and calculations. See DOI: 10.1039/b925329g/ mixture are used to identify these stabilising factors.
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 | 4217
Scheme 3 IL1: C4C1Im NTf2, carbon numbering of the cation.
of stirring in 1-butyl-3-methylimidazolium bis(trifluorosulfonyl)- H/D exchange must involve RuNPs’ surface. To prove, there-
imide, [C1C4Im][NTf2], Scheme 3, this method yields very small fore, the presence of these surface hydrides, further experi-
NPs with a narrow size distribution (mean size: 1.1 0.2 nm ments are performed characteristic of metal surface reactivity,
calculated from TEM images captured in the ILs).21 These such as hydrogenation.
conditions are replicated throughout this work, obtaining
fresh IL-stabilised colloidal solutions, on which NMR and
mass spectroscopy experiments are performed, to determine
the key points in their stabilisation. Note that no deposition of
the RuNP suspension is observed after several months under
argon atmosphere.
4218 | Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 This journal is
c the Owner Societies 2010
21 Hydrogenation of ethylene It is, nevertheless, very interesting that the presence of
hydrides on ruthenium particles in ionic liquids is confirmed,
The reaction of alkenes with PVP (polyvinyl pyrrolidone)- and
since this is in agreement with the presence of these nano-
ligand-stabilised RuNPs has recently been used to quantify the
particles in uncharged lipophilic domains.
surface hydrides.17 A similar approach is transposed to a
solution of RuNPs in IL1. In the following discussion, the
RuNPs are generated in 2 mL of IL1 from 86 mmol of Stabilising effect of surface hydrides
Ru(COD)(COT). Given that the dispersion of NPs, (D, i.e.
In order to ascertain the stabilising effect of hydrogen, two
the ratio surface atoms Ns per total number of atoms Nt,
aliquots of the RuNP/IL1 solution are taken and stirred,
D = Ns/Nt), is correlated to their size,24,25 we could estimate
at 100 1C, for a period of 24 h, either under H2 or under Ar
that for RuNPs of 1.1 nm the dispersion is 75%, corresponding,
(1 bar). The resulting solutions are analysed by TEM. For the
for 86 mmol of Ru(COD)(COT), to Ns E 65 mmol of Ru
aliquot heated under H2, the mean size of the RuNPs increases
surface atoms. This IL1-stabilised RuNP solution is flushed
to 1.4 0.7 nm, and for the aliquot heated under argon, the
overnight with argon to remove dissolved hydrogen. Then, the
mean size increases to 2.2 0.8 nm. This difference in size
solution is stirred under an ethylene atmosphere (105.5 mbar)
increase with the gas atmosphere proves the inferior stability
at 25 1C. After 24 h of reaction, the system reaches equili-
of the particles when heated under argon compared to under
brium, i.e. the pressure remains constant. In the gas phase,
H2. In addition, when the RuNP/IL1 solution is treated under
analysed by GC, 11 1 mmol of ethane and 0.1 mmol of high vacuum (10ÿ6 mbar) overnight at room temperature, the
n-butane are found, corresponding to at least 22 1 mmol of TEM images show an increase in size from 1.1 0.2 nm to
surface hydride : (C2H4 + H2 - C2H6). Note that butenes 1.7 0.4 nm, Fig. 2. From these results, it appears that there is
(trans- and 1-) are also observed (0.2 mmol); however, no less coalescence of the RuNPs when heated under H2 at 100 1C
tel-00708600, version 1 - 15 Jun 2012
hydrogen would be consumed in their formation. In summary, than under argon at the same temperature. This may be
IL1-stabilised RuNPs react with ethylene at room tempera- explained by the elimination of the hydrogen atoms reversibly
ture to give ethane (hydrogenation) and surface-bound alkyl bound to the metal surface, which is favoured by heating
species (homologation).26,27 under an inert atmosphere or treatment under high vacuum. In
To quantify the total amount of surface-bound alkyl conclusion, the presence of surface Ru–H stabilises the
species, the gas phase is evacuated and substituted with H2 RuNPs. These stability experiments are also consistent with
(120 mbar), and the medium is heated at 100 1C for 12 h. In the the lability of the hydrides which can be desorbed from the
gas phase, methane (4.6 mmol), ethane (8.2 mmol), propane nanoparticles in vacuo or at 100 1C.
(2.7 mmol), n-butane (1.0 mmol) and n-pentane (0.1 mmol) are
detected, corresponding to at least a further 16 mmol of
hydrides. Knowing the total amount of surface Ru atoms Evidence for close proximity between the RuNP surface and the
available (65 mmol) and the number of surface hydrides alkyl chains
consumed in all of the hydrogenation reactions (38 mmol), it The fact that in C1CnImNTf2, the non-polar domains control
is possible to calculate the ratio of hydrides per surface Ru the local concentration of Ru(COD)(COT) and consequently
atom: ca. 0.6 H per RuS, (see supplementary information for the size of RuNPs generated in situ could allow us to consider
calculationw). The combination of these results provides them as nanoreactors in which the phenomenon of crystal
substantial evidence for the existence of surface hydrides on growth occurs.10 If the RuNPs are contained within these non
the IL1-stabilised RuNPs, although the level of error in the polar pockets, it would imply that the side alkyl groups of the
quantitative results means that it is difficult to establish ILs and the RuNP surface are in close proximity or even in
accurately the quantity. The amount of H adsorbed on RuNPs interaction with one another.
is reported to be between 1.1–1.3 H/RuS for PVP- and ligand- To demonstrate this, RuNPs are synthesised under deuterium
stabilised in organic solvent,16 and between 1 and 2 H/RuS instead of hydrogen (following otherwise identical reaction
when supported on oxides.23 Note that if the IL1-stabilised conditions), since an H–D exchange on the side group of the
solutions of RuNP are treated under high vacuum (10ÿ6 mbar) IL would prove the existence of the NPs in these domains. In
overnight at room temperature instead of being flushed with the case of IL1, 2H NMR spectra show us that, as expected,
argon to remove the dissolved H2, ethylene conversion is deuteration occurs at all positions on the imidazolium ring due
negligible. to the relatively high acidity of these protons. One small peak
Indeed, hydrides bound to metal surfaces are labile and the is apparent at B1.5 ppm, which may correspond to deuteration
dissociation of H2 adsorbed on a metal surface is reversible: M + at position C8, i.e. the penultimate carbon of the butyl chain;
H2 2 MH2 2 2M–H. In the case of Ru, about 2/3 of the however, this is not irrefutable. The intensities of the peaks in
total amount of surface hydride is reversibly bound. These H the 1H NMR spectrum are reduced by roughly 30% at
species correspond to more weakly adsorbed hydrogen on the positions C2, C4 and C5, but no significant reduction in
surface, which are easily removed by evacuating the sample intensity is noted for any of the alkyl peaks. ESI mass spectra
under vacuum at 25 1C.23 Consequently, this explains why the indicate the formation of mono-, di-, tri- and a trace amount
hydrogenation (in the absence of hydrogen) of ethylene does of tetra-deuterated species. This is in agreement with the
not occur with NPs previously treated under high dynamic higher reactivity of the hydrogen bound to the imidazolium
vacuum (10ÿ5 mbar) for an extended period (12 h), because ring than to the alkyl side chain. Note that, in contrast to
very few surface hydrides remain.23,28–30 findings for IrNPs,9 where deuteration was mainly at position
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 | 4219
tel-00708600, version 1 - 15 Jun 2012
Fig. 2 TEM images captured in situ in ionic liquids and comparative size histograms of RuNPs/C1C4Im NTf2 (1.1 0.2 nm) heated to 100 1C for
24 h under H2, under Ar and kept under dynamic vacuum (10ÿ5 mbar) for 24 h at 25 1C.
4220 | Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 This journal is
c the Owner Societies 2010
Fig. 6 2H NMR spectra of BzC1Im NTf2, IL3, after RuNP formation
at 0 1C under deuterium (lower) and after a further 12 h of heating to
50 1C under deuterium (upper).
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 | 4221
purification. Bis(trifluoromethanesulfonyl)imide lithium salt
(Solvionic) was used without further purification.
Elemental analyses were performed at the CNRS Central
Analysis Department of Solaize.
Solution NMR spectra were recorded on Bruker Avance
300 spectrometer for 1H and 13C DRX 500 for 2H. Gas phase
1
H NMR spectra were obtained on a Bruker DRX 500
instrument at 298 K (nominal) with a resonance frequency
at 500.130 MHz.
Mass spectra were acquired on a ThermoFinnigan LCQ
Advantage ion trap instrument, detecting positive (+) an
negative (ÿ) ions in the ESI mode. Samples (1–10 mg mLÿ1
in acetonitrile) were infused directly into the source (5 mL minÿ1)
using a syringe pump. The following source parameters were
applied: spray voltage 3.0–3.5 kV, nitrogen sheath gas flow
5–20 arbitrary units. The heated capillary was held at 200 1C.
MS spectra were obtained by applying a relative collision
energy of 25–40% of the instrumental maximum. (Spectra in
supplementary information.w)
Transmission electron microscopy (TEM) experiments were
performed directly in the IL media.21 A thin film of RuNP
tel-00708600, version 1 - 15 Jun 2012
Conclusion GC analyses
The phenomenon of crystal growth of RuNPs generated in situ The products were quantitatively analysed by gas chromato-
from Ru(COD)(COT), in imidazolium-based ILs is found to graphy on a P6890 chromatograph equipped with a flame
be controlled by the size of non-polar domains created by the ionisation detector (FID) and an Al2O3/KCl column (L: 50 m,
grouping of lipophilic alkyl chains. Close proximity of the jint: 0.32 mm, film thickness: 5 mm). The injector and detector
RuNPs to the non polar R group of the cation is evidenced by temperature was 230 1C, and the injection volume was 1 mL.
in situ labelling experiments (deuteration and reduction), thus The temperature was fixed at 190 1C.
corroborating the hypothesis that RuNPs are surrounded by
these non-polar pockets. Addition of water, destroying the Hydrogenation of ethylene by RuNPs in the IL
ionic network, induces agglomeration of the RuNPs, supporting
A 2 mL sample of RuNP solution in IL was introduced under
the fact that the RuNPs are originally isolated within non-
argon into a Schlenk tube of known volume. The sample was
polar pockets by the 3-D network of ionic channels. Further-
treated under flow of argon for 18 h. The argon atmosphere
more, the presence of surface hydrides and their role in the
was replaced with a known pressure of ethylene using a
stabilisation of the nanoparticles is demonstrated. Finally, the
vacuum line and the system was stirred. A decrease of the
RuNPs are shown to be hydrophobic.
internal pressure was observed, and the composition of the gas
phase was monitored by gas chromatography. After 12 h, the
Experimental section atmosphere was replaced by a H2 atmosphere and the system
was heated and stirred for 12 h. The composition of the gas
All operations were performed in the strict absence of oxygen phase was again monitored by gas chromatography.
and water under a purified argon atmosphere using glovebox
(Jacomex or MBraun) or vacuum-line techniques. Imidazolium
H–D exchange
ionic liquids C4C1Im NTf2, IL1 C4//C1Im NTf2, IL2 BzC1Im
NTf2, IL3,39 Ru(COD)(COT),40 and RuNPs were synthesised A 2 mL sample of RuNP solution in IL was introduced under
as already reported.21,41 The halide content is under 200 ppm argon into a Schlenk tube, which was then treated under
(E.A), and water under 5 ppm, (limit of Karl Fischer titration). dynamic vacuum (10ÿ5 mbar) during a period of 1 or 4 h.
1-Methylimidazole (>99%) was purchased from Aldrich This was then charged with deuterium (1 bar) and stirred at
and distilled prior to use. Chlorobutane, benzyl chloride and 25 1C for a period of 1 or 4 d. The gas phase was subsequently
4-chlorobut-1-ene (>99%, Aldrich) were used without further expanded into an airtight Young tube for NMR analysis.
4222 | Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 This journal is
c the Owner Societies 2010
Acknowledgements 20 J. Dupont, G. S. Fonseca, A. P. Umpierre, P. F. P. Fichtner and
S. R. Teixeira, J. Am. Chem. Soc., 2002, 124, 4228–4229.
We would like to thank ANR CALIST and Ministe`re de 21 T. Gutel, J. Garcia-Anton, K. Pelzer, K. Philippot, C. C. Santini,
l’Enseignement Supe´rieur (P.S.C.) for financial support. Anne Y. Chauvin, B. Chaudret and J.-M. Basset, J. Mater. Chem., 2007,
17, 3290–3292.
Baudouin for 2H NMR experiments. 22 J. L. Anthony, J. L. Anderson, E. J. Maginn and J. F. Brennecke,
J. Phys. Chem. B, 2005, 109, 6366–6374.
23 R. Berthoud, P. Delichere, D. Gajan, W. Lukens, K. Pelzer,
Notes and references J.-M. Basset, J.-P. Candy and C. Coperet, J. Catal., 2008, 260,
1 (a) H. Bönnemann, K. S. Nagabhushana and R. M. Richards, 387–391; J. P. Candy, A. Elmansour, O. A. Ferretti, G. Mabillon,
in Nanoparticles and Catalysis, ed. D. Astruc, Wiley-VCH, J. P. Bournonville, J. M. Basset and G. Martino, J. Catal., 1988,
Weinheim, 2008; (b) M.-C. Daniel and D. Astruc, Chem. Rev., 112, 201–209.
2004, 104, 293–346. 24 R. Van Hardeveld and F. Hartog, Surf. Sci., 1969, 15, 189–230.
2 G. A. Somorjai and J. Y. Park, Top. Catal., 2008, 49, 126–135. 25 J.-M. Bassett, A. Baudouin, F. Bayard, J.-P. Candy, C. Copéret,
3 L. S. Ott and R. G. Finke, Coord. Chem. Rev., 2007, 251, A. de Mallmann, G. Godard, E. Kuntz, F. Lefebvre, C. Lucas,
1075–1100. S. Norsic, K. Pelzer, A. Quadrelli, C. Santini, D. Soulivong,
4 G. Schmid, Nanoparticles: From Theory to Application, Wiley- F. Stoffelbach, M. Taoufik, C. Thieuleux, J. Thivolle-Cazat and
VCH, Weinheim, 2004. L. Veyre, in Modern Surface Organometallic Chemistry,
5 (a) J. Dupont and D. d. O. Silva, in Nanoparticles and Catalysis, ed. ed. J. M. Basset, A. Psaro Rinaldo, A. Roberto Dominique and
D. Astruc, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, A. Ugo Renato, Wiley VCH, Weinheim, 2009, pp. 23–68.
2008, pp. 195–218; (b) A. Taubert and Z. Li, Dalton Trans., 2007, 26 M. Leconte, A. Theolier and J. M. Basset, J. Mol. Catal., 1985, 28,
723–727; (c) R. E. Morris, Angew. Chem., Int. Ed., 2008, 47, 217–231.
442–444; (d) J. Kraemer, E. Redel, R. Thomann and C. Janiak, 27 M. Leconte, A. Theolier, D. Rojas and J. M. Basset, J. Am. Chem.
Organometallics, 2008, 27, 1976–1978. Soc., 1984, 106, 1141–1142.
6 A. Corma, I. Dominguez, T. Rodenas and M. J. Sabater, J. Catal., 28 I. M. Ciobica, A. W. Kleyn and R. A. Van Santen, J. Phys. Chem.
2008, 259, 26–35. B, 2003, 107, 164–172.
tel-00708600, version 1 - 15 Jun 2012
7 A. P. Umpierre, G. Machado, G. H. Fecher, J. Morais and 29 T. H. Rod, A. Logadottir and J. K. Norskov, J. Chem. Phys., 2000,
J. Dupont, Adv. Synth. Catal., 2005, 347, 1404–1412. 112, 5343–5347.
8 L. S. Ott and R. G. Finke, Inorg. Chem., 2006, 45, 8382–8393. 30 J. Wang, C. Y. Fan, Q. Sun, K. Reuter, K. Jacobi, M. Scheffler and
9 L. S. Ott, M. L. Cline, M. Deetlefs, K. R. Seddon and R. G. Finke, G. Ertl, Angew. Chem., Int. Ed., 2003, 42, 2151–2154.
J. Am. Chem. Soc., 2005, 127, 5758–5759. 31 M. P. Stracke, M. V. Migliorini, E. Lissner, H. S. Schrekker,
10 T. Gutel, C. C. Santini, K. Philippot, K. Pelzer, Agilio Padua, D. Back, E. S. Lang, J. Dupont and R. S. Gonçalves, New J.
B. Chaudret, Y. Chauvin and J.-M. Basset, J. Mater. Chem., 2009, Chem., 2009, 33, 82–87.
19, 3624–3631. 32 A. Roucoux, A. Nowiki and K. Philippot, in Nanoparticles,
11 A. A. H. Padua, M. F. Costa Gomes and J. N. A. Canongia Lopes, ed. D. Astruc, Wiley-VCH, Weinheim, 2007.
Acc. Chem. Res., 2007, 40, 1087–1096. 33 L. Cammarata, S. G. Kazarian, P. A. Salter and T. Welton, Phys.
12 M. G. Del Popolo, C. L. Mullan, J. D. Holbrey, C. Hardacre and Chem. Chem. Phys., 2001, 3, 5192–5200.
P. Ballone, J. Am. Chem. Soc., 2008, 130, 7032–7041. 34 J. N. Canongia Lopes, M. F. Costa Gomes and A. A. H. Padua,
13 T. Gutel, C. C. Santini, A. A. H. Padua, B. Fenet, Y. Chauvin, J. Phys. Chem. B, 2006, 110, 16816–16818.
J. N. Canongia Lopes, F. Bayard, M. F. Costa Gomes and 35 Y.-Y. Jiang, Z. Zhou, Z. Jiao, L. Li, Y.-T. Wu and Z.-B. Zhang,
A. S. Pensado, J. Phys. Chem. B, 2009, 113, 170–177. J. Phys. Chem. B, 2007, 111, 5058–5061.
14 K. Pelzer, O. Vidoni, K. Philippot, B. Chaudret and V. Collière, 36 A. Mele, C. D. Tran and S. H. De Paoli Lacerda, Angew. Chem.,
Adv. Funct. Mater., 2003, 13, 118–126. Int. Ed., 2003, 42, 4364–4366.
15 K. Philippot and B. Chaudret, C. R. Chimie, 2003, 6, 37 A.-L. Rollet, P. Porion, M. Vaultier, I. Billard, M. Deschamps,
1019–1034. C. Bessada and L. Jouvensal, J. Phys. Chem. B, 2007, 111,
16 T. Pery, K. Pelzer, G. Buntkowsky, K. Philippot, H.-H. Limbach 11888–11891.
and B. Chaudret, ChemPhysChem, 2005, 6, 605–607. 38 B. L. Bhargava and M. L. Klein, J. Phys. Chem. B, 2009, 113,
17 J. Garcia-Anton, M. R. Axet, S. Jansat, K. Philippot, B. Chaudret, 9499–9505.
T. Pery, G. Buntkowsky and H.-H. Limbach, Angew. Chem., Int. 39 L. Magna, Y. Chauvin, G. P. Niccolai and J.-M. Basset, Organo-
Ed., 2008, 47, 2074–2078. metallics, 2003, 22, 4418–4425.
18 C. Pan, K. Pelzer, K. Philippot, B. Chaudret, F. Dassenoy, 40 P. Pertici, G. Vitulli, M. Paci and L. Porri, J. Chem. Soc., Dalton
P. Lecante and M.-J. Casanove, J. Am. Chem. Soc., 2001, 123, Trans., 1980, 1961–1964.
7584–7593. 41 T. Gutel, C. C. Santini, K. Philippot, A. A. H. Padua, K. Pelzer,
19 K. Pelzer, K. Philippot and B. Chaudret, Z. Phys. Chem., 2003, B. Chaudret, Y. Chauvin and J.-M. Basset, J. Mater. Chem., 2009,
217, 1539–1547. 19, 3624–3631.
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 4217–4223 | 4223
8156 J. Phys. Chem. B 2010, 114, 8156–8165
difference in reactivity, molecular interactions and the microscopic structure of ionic liquid +1,3-cyclohexadiene
mixtures were studied by NMR and titration calorimetry experiments, and by molecular simulation in the
liquid phase. Diffusivity and viscosity measurements allowed the characterization of mass transport in the
reaction media. We could conclude that the diffusivity of 1,3-cyclohexadiene is 1.9 times higher in
[C1C4Im][NTf2] than in [C1C1C4Im][NTf2] and that this difference could explain the lower reactivity observed
in [C1C1C4Im][NTf2].
1. Introduction solutes are solvated in the ionic liquid:14 nonpolar species will
be solvated within the hydrophobic domains whereas polar
The application of ionic liquids (ILs) in catalyzed reactions
substrates tend to interact with the polar network. The strength
is of increasing importance,1,2 in particular, they provide
of cation-anion association15 can also influence the possibility
excellent media for conducting catalytic hydrogenations.1-3
of establishing ion-solute specific interactions (e.g., π-cation
Several research groups report differences in the rates and
interactions), as has been shown in the case of toluene.9 For a
selectivities in ILs when compared to corresponding reactions
nonaromatic π-system, such as 1,3-cyclohexadiene, solvation
in molecular solvents.2
is probably a combination of more subtle interactions with the
As reaction media, ILs have specific properties that may have
ionic liquid. This information can be assessed through the study
consequences on the catalytic process. In some catalytic
of the thermodynamic properties of solution (solubility, enthalpy
reactions, the difference between the use of ILs and traditional
of solution) and from the characterization of the mass transfer
solvents has been related to the chemical role of ILs, which can
through viscosity and diffusivity data.
serve as new ligands for the catalytic metal center, as catalyst
The aim of this work is to study the influence of the nature
activators, as cocatalysts, or even as catalysts themselves.1,4,5 In
of the ionic liquid on the catalytic hydrogenation of 1,3-
other cases, differences in reactivity have a physical-chemical
cyclohexadiene (CYD) with [Rh(COD)(PPh3)2]NTf2 (COD )
origin,1,6 resulting from peculiar solvation phenomena including
1,5-cyclooctadiene). First, we have quantified the difference in
specific interactions between the IL and the substrate (H-bonds,
reactivity when the reaction is performed in two ionic liquids:
cation-π),7-9 mass transfer factors (viscosity, diffusivity),1 and
1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide,
effects attributed to the highly structured nature of ILs.10-12
[C1C4Im][NTf2], and 1-butyl-2,3-dimethylimidazolium bis(tri-
ILs have a high degree of self-organization due to the fluoromethylsulfonyl)imide, [C1C1C4Im][NTf2]. Following this,
coexistence of charged moieties and hydrophobic alkyl chains, we have attempted to rationalize the differences encountered
to the importance of both Coulombic and van der Waals through NMR characterization of the molecular interactions
interactions, and frequently to the presence of hydrogen bonds between the ionic liquids and the substrate, the study of the
between the cation and the anion. In butylimidazolium ionic microscopic structure of the IL-substrate mixtures, the measure-
liquids, the alkyl side chain is sufficiently long to allow the ment of the thermodynamic properties of mixing (such as the
formation of nonpolar domains that coexist with an ionic solubility and the heat of mixing), the determination of the
network.13 This heterogeneous structure influences the way viscosity of the reaction media, and the measurement of
diffusivity of CYD in the ionic liquids.
* To whom correspondence should be addressed. E-mail: C.C.S., santini@
cpe.fr; M.F.C.G., margarida.costa-gomes@univ-bpclermont.fr.
†
Université de Lyon. 2. Experimental Section
‡
CNRS.
§
Clermont Université and Université Blaise Pascal. Materials. 1-Methylimidazole (>99%, Aldrich) and 1,2-
|
Centre Commun de RMN. dimethylimidazole (>98%, Aldrich) were distilled prior to use.
10.1021/jp102941n 2010 American Chemical Society
Published on Web 05/26/2010
Catalytic Hydrogenation of 1,3-Cyclohexadiene J. Phys. Chem. B, Vol. 114, No. 24, 2010 8157
Anhydrous 1,3-cyclohexadiene (99.8%, Aldrich) and 1,3- in presence of toluene as an internal standard. A HP6890
cyclohexadiene stabilized (96%, Acros Organics, stabilized with chromatograph equipped with FID detector and an Al2O3/KCl
50 ppm of 2,6-di-tert-butyl-4-methylphenol, BHT) distilled over column (L ) 50 m, φint ) 0.32 mm, film thickness ) 5 µm)
NaK alloy and stored on zeolites were used for the reaction was used. The injector and detector temperatures were set to
and the physical chemical measurements, respectively. Bis(tri- 230 °C. Samples were injected in a volume of 1 µL. The
fluoromethanesulfonyl)imide lithium salt (>99%, Solvionic) and temperature of the column was fixed at 190 °C (see Supporting
[Rh(COD)Cl]2 (>99%, Strem) were used without further puri- Information for calculations).
fication. All other reagents and solvents were commercially Solubility and Phase Diagrams. Liquid-liquid phase equi-
available and were used as received. Ionic liquids, synthesized libria of the mixture of [C1C4ImNTf2] or [C1C1C4Im][NTf2] and
as previously reported,16 were dried overnight under high CYD at atmospheric pressure were determined using a dynamic
vacuum and stored in a glovebox (Jacomex) to guarantee method with visual detection of solution turbidity, as already
rigorously anhydrous products. described.18
Synthesis of the Catalysts. [Rh(COD)(PPh3)2]NTf2. The The mixtures of ionic liquid and CYD at different composi-
procedure followed was adapted from the literature.17 A mixture tions were prepared gravimetrically in a glass vial equipped with
of [Rh(COD)Cl]2 (100 mg, 0.2 mmol) dissolved in 2 mL of a stirring bar. First, the ionic liquid was introduced into a glass
dichloromethane and of LiNTf2 (107 mg, 0.37 mmol) dissolved vial, then the appropriate amount of CYD was added and the
in 2 mL of water was stirred vigorously while triphenylphos- vial was sealed. To minimize the volume of the vapor phase in
phine (405 mg, 1.5 mmol) was added. After 2 h, the dichlo- equilibrium with the ionic liquid solution and to reduce the error
romethane layer was removed, washed three times with 2 mL in composition due to differential evaporation, the glass vial
of water and dried over anhydrous Na2SO4. Ethanol (1 mL) and was almost completely filled with the mixture. The uncertainty
diethyl ether (2 mL) were slowly added to complete crystal- of the mole fraction is estimated as (0.0001. The cells were
lization. The orange crystals were filtered off and dried under then immersed in a thermostatic water bath whose temperature
reduced pressure. Yield: 200 mg (99%). 1H NMR (300 MHz, was monitored using a platinum resistance thermometer with a
tel-00708600, version 1 - 15 Jun 2012
CD2Cl2): δ ) 2.2 (m, 8H), 4.6 (s, 4H), 7.5 (m, 30H). 31P NMR precision of (0.1 K. The temperature of the bath was first
(121 MHz, CD2Cl2): δ ) 23.6 with JRh-P ) 147 Hz. 103Rh NMR increased slowly until one phase was observed. The clear
(500 MHz, CD2Cl2): δ ) -80 ppm with JRh-P ) 147 Hz. Mass homogeneous system was then cooled very slowly (5 K/h) under
spectrometry: positive mode: m/z ) 473 w Rh(COD)(PPh3)+; continuous stirring. The temperature at which the first sign of
627 w Rh(PPh3)2+. UV-vis spectroscopy (6.6 mg of turbidity (first cloudiness) appeared was considered as the
[Rh(COD)(PPh3)2]NTf2 dissolved in 2 mL of one of ionic temperature of the liquid-liquid phase transition. The overall
liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2]); spectrum re- accuracy in the measurement of cloud-point temperatures is
corded at λmax ) 450 nm. estimated to be (2 K.
[Rh(CYD)(PPh3)2]NTf2. [Rh(COD)(PPh3)2]NTf2 (100 mg, Isothermal Titration Calorimetry. The heat effects resulting
9.84 × 10-5 mol) was dissolved in 1 mL of CYD. After 10 from mixing aliquots of CYD with the ionic liquid were
min, the excess of solvent was removed under reduced measured at 303.15 K using an isothermal titration nanocalo-
pressure. Yield: 100 mg (99%). 1H NMR (300 MHz, CD2Cl2): rimeter equipped with 4 mL glass cells in a Thermal Activity
δ ) 2.1 (s, 2H), 5.3 (s, 2H), 7.5 (m, 30H). 31P NMR (121 Monitor TAM III from TA Instruments. An electrical calibration
MHz, CD2Cl2): δ ) 25.2 ppm with JRh-P ) 17 Hz. Mass was done before each experiment, and the instrument was
spectrometry: positive mode: m/z ) 707 w Rh(cyclohex- chemically calibrated 5 times by titration of a 0.01 M aqueous
adiene)(PPh3)2+; m/z ) 627 w Rh(PPh3)2+. UV-vis spec- solution of 18-crown-6 ethers with an 0.2 M aqueous solution
troscopy (6.6 mg of [Rh(CYD)(PPh3)2]NTf2 dissolved in 2 mL of of BaCl2. The enthalpies of binding of Ba2+ ions to 18-crown-6
one of ionic liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2]); were found to be slightly higher than those reported in the
spectrum recorded at λmax ) 500 nm. literature, 2.5%19 and 1.6%,20,21 respectively. No correction
Ligand Exchange. In [Rh(CYD)(PPh3)2]NTf2 was followed attributable to these differences was introduced in the raw data.
by UV-vis spectroscopy. CYD (0.15 mL, 1.6 mmol) was added Approximately 2.75 mL of degassed ionic liquid were
to a system of [Rh(COD)(PPh3)2]NTf2 (1.6 mg, 1.6 µmol) in introduced into 4 mL glass measuring and reference cells. The
one of the ionic liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2] liquid in the measuring cell was stirred by a turbine stirrer at
(1 mL). The UV-visible spectra of 1 mL of the solution were 160 rpm and volumes of 4 µL of CYD were injected during
recorded on the Perkin-Elmer LAMBDA 950 Spectrophotometer 180 s using a motor driven pump (Thermometric 3810 Syringe
in stirred and closed UV cells at λmax ) 500 nm every second. Pump) equipped with a 100 µL gastight Hamilton syringe. In
Reaction with H2. [Rh(CYD)(PPh3)2]NTf2 (8 mg, 8.1 × 10-6 all experiments, the intervals between consecutive injections
mmol) was dissolved in one of the ionic liquids, [C1C4Im][NTf2] were 35-40 min, which provided a good thermal stabilization
or [C1C1C4Im][NTf2] (5 mL) and pressurized under 1.2 bar of of the ionic liquid solution and the return to a stable baseline.
hydrogen. The UV-visible spectra of 1 mL of the solution were To minimize the undesirable effects of diffusion of the ionic
recorded at a given time at λmax ) 500 nm. liquid into the canula linking the CYD syringe to the cell, the
Hydrogenation of 1,3-Cyclohexadiene. The hydroge- canula was immersed in the sample 10 min prior the first
nation of CYD was carried out at 1.2 atm of H2 and 30 °C. injection.
CYD (0.15 mL, 1.6 mmol) was dissolved in a system of A peak with an area proportional to the resulting heat effect
[Rh(COD)(PPh3)2]NTf2 (3.2 mg, 3.2 µmol) in one of ionic Qi translates to the thermal effect due to each injection of CYD.
liquids, [C1C4Im][NTf2] or [C1C1C4Im][NTf2] (1 mL), under The integration of peaks from the recorded calorimetric plots
argon resulting in red homogeneous solutions. The reaction was performed using the TAM III Assistant software. Each
mixture was kept under a hydrogen atmosphere (1.2 atm, experiment was repeated four times to obtain reproducible values
constant pressure) until 4 mL of acetonitrile was added to the of Qi at different concentrations within the error bar of (2%.
catalytic solution. The product distribution in the reaction Density and Viscosity. The mixtures of ionic liquid and CYD
mixture and the conversion were determined by GC analyses at different compositions were prepared gravimetrically follow-
8158 J. Phys. Chem. B, Vol. 114, No. 24, 2010 Campbell et al.
ing the procedure already described, including the precautions gradients to improve line shape and trapezoidal gradients (the
to minimize vapor headspace.9 The viscosity of the mixture was pulse sequences shown in Figure S-6 in Supporting Information).
measured at 298.15 K (controlled to within (0.005 K and The diffusion evolution time was 100 ms, the constant amplitude
measured with the accuracy better than (0.05 K) using a rolling part of the gradient was 3 ms, and the cosine raising and falling
ball viscometer from Anton Paar, model AMVn, equipped with part of gradient were 150 µs. The diffusion space was sampled
capillary tubes of 3.0 and 1.8 mm in diameter. Before starting by 32 linearly spaced gradients.
the measurements, the 3.0 mm diameter capillary tube was Molecular Simulation. The microscopic structures of the IL-
calibrated as a function of temperature and angle of measure- CYD mixtures studied experimentally were also investigated
ment with a standard viscosity oil from Cannon (N35). The 1.8 by molecular simulation, using an atomistic force field that
mm diameter tube was calibrated with water by the manufac- describes interactions and conformations.23-25 CYD was rep-
turer. The overall uncertainty of the viscosity is estimated as resented by the optimized potential for the liquid simulations
(2.0%. force field in its all-atom explicit version (OPLS-AA).26 ILs
The densities of the mixtures, necessary to calculate the were represented by a specifically parametrized force field of
viscosities were measured in an Anton Paar vibrating tube the OPLS-AA family in which particular attention was paid to
densimeter model 512 P, at 298.15 K (measured by a calibrated the description of electrostatic charge distributions and torsion
PRT with an accuracy of (0.02 K). The densimeter was energy profiles. The OPLS-AA force field is known to reproduce
calibrated using n-heptane, bromobenzene, and 2,4-dichloro- H-bonds well; the electron density of aromatic systems is
toluene. The overall uncertainty of the density is estimated as represented by the values of electrostatic charges on the relevant
(0.01%. atoms that account for the molecular mutipoles, combined with
NMR Spectroscopy. 1H, 13C, and 31P solution NMR data the Lennard-Jones sites that account for dispersion interactions.
were collected at room temperature on a Bruker AC 300 MHz Explicit polarization of electron clouds is not included in the
spectrometer with the resonance frequency at 300.130 MHz for present model. Although this may be important to correctly
the 1H nucleus. 103Rh solution NMR was carried out on a Bruker reproduce dynamic properties of ILs, structural features and
tel-00708600, version 1 - 15 Jun 2012
DRX 500 instrument at 298 K (nominal) with a resonance thermodynamic quantities have been described to equivalent
frequency at 500.130 MHz. The solvent used (CD2Cl2) was levels of accuracy using fixed-charge models.27
distilled and kept in a rotaflo with molecular sieves. Chemical Molecular dynamics simulations of condensed-phase CYD-
shifts are reported in ppm (singlet ) s, doublet ) d, doublet of [C1C4Im][NTf2] and CYD-[C1C1C4Im][NTf2] mixtures were
doublet ) dd, and multiplet ) m) and were measured relative performed using the DL_POLY program.28 System sizes were
to the residual proton of the solvent to CHDCl2 for 1H, to CD2Cl2 chosen so as to contain about 10 000 atoms, and so the numbers
for 13C, and to H3PO4 for 31P spectra. of cations, anions, and CYD molecules varied according to
For 1H 1D NMR spectroscopy, the samples with molar ratio composition (e.g., 128 ion pairs and 64 molecules of CYD for
R ) 0.5). Initial low-density configurations, with ions and
R ) 0.5 (R is the molar ratio between the amount of substance
molecules placed at random in periodic cubic boxes, were
of the hydrocarbon and the amount of substance of IL) were
equilibrated to attain liquid-like densities and structures at 400
prepared. The mixtures of ionic liquids and hydrocarbon were
K and 1 bar. Temperature and pressure were maintained using
prepared in closed vials in a glovebox by adding the appropriate
a Nosé-Hoover thermostat and barostat, respectively. Produc-
amount of CYD to each ionic liquid, [C1C4Im][NTf2] or
tion runs then took 500 ps with an explicit cutoff distance of
[C1C1C4Im][NTf2]. The resulting systems were stirred for 24 h
16 Å for nonbonded interactions, and long-range corrections
at 303 K, resulting in homogeneous monophasic solutions.
applied for repulsive-dispersive interactions. Electrostatic ener-
Approximately 0.3 mL of the sample was then introduced into
gies were calculated using the Ewald summation method with
a 5 mm NMR tube. A stem coaxial capillary tube loaded with
a relative accuracy of 10-4. Structural quantities such as radial
CD2Cl2 was inserted into the 5 mm NMR tube to avoid any
and spatial distribution functions were calculated from configu-
contact between the deuterated solvent and the analyzed mixture.
rations generated during the production runs.
The deuterium in CD2Cl2 was used for the external lock of the
NMR magnetic field and the residual CHDCl2 in CD2Cl2 was
used as the 1H NMR external reference at 5.32 ppm. When 1H 3. Results and Discussion
NMR data are obtained in this way, the reference signal of Hydrogenation of 1,3-Cyclohexadiene. Selective hydroge-
CHDCl2 remains constant and is not affected by changes in nation of CYD into cyclohexene can be carried out, with good
sample concentration. conversion rates, using Osborn’s complex [Rh(NBD)(PPh3)2]PF6
For ROESY (nuclear overhauser effect spectroscopy) experi- (NBD ) norbornadiene) in IL media such as [C1C4Im][SbF6]
ments in the rotating frame, the 2D sequence was built with or [C1C4Im][PF6].29 To avoid impurities such as chloride and
the scheme proposed by Bodenhausen (the pulse sequence water, ILs based on the bis(trifluoromethylsulfonyl)imide anion,
shown in Figure S-3 in Supporting Information).22 The mixing [C1C4Im][NTf2] and [C1C1C4Im][NTf2], were chosen here since
time (200 ms) is split into two parts separated by a p-pulse. At they are hydrophobic and liquid at room temperature and their
each side of the spin lock the B1 field is ramped linearly (4.5 purification is well controlled.16 Anion exchange between the
ms) to ensure adiabatic conditions for spin lock. During the IL and the catalyst can be circumvented by replacing Osborn’s
first spin lock pulse, the frequency is shifted to O1 + Df whereas complex with [Rh(COD)(PPh3)2]NTf2 (COD ) 1,5-cycloocta-
the second pulse frequency is set to O1 - Df; O1 is the offset diene).
frequency and Df set to give a B1 field at the magic angle. In In the hydrogenation experiment the appropriate quantity of
this way, the ROESY response is roughly constant across the CYD was added to the yellow solution of [Rh(COD)(PPh3)2]-
spectra of interest. 1D sequence PFGSE (pulse field gradient NTf2 in [C1C4Im][NTf2] or [C1C1C4Im][NTf2] to reach a molar
spin echo for selective excitation) has been used, and spin lock ratio between CYD and IL R ) 0.5, corresponding to a molar
followed the same scheme as previously. ratio substrate/Rh atom r ) 500. The resulting red solution was
For the 2D DOSY (diffusion order spectroscopy) experiments stirred under 1.2 bar of hydrogen at 303 K. The formation of
a Bruker sequence ledbpgp2s was implemented for shorter pulse cyclohexene (CYE) was followed by GC analysis. Note that
Catalytic Hydrogenation of 1,3-Cyclohexadiene J. Phys. Chem. B, Vol. 114, No. 24, 2010 8159
Figure 1. Hydrogenation of CYD at 303 K under 1.2 bar H2 (R ) 0.5 Figure 2. Evolution as a function of time of the ligand exchange
and r ) n(CYD)/n(Rh) ) 500) in [C1C4Im][NTf2] ([) and reaction COD-CYD for [Rh(COD)(PPh3)2]NTf2 in [C1C4Im][NTf2]
[C1C1C4Im][NTf2] (b). (full line) and [C1C1C4Im][NTf2] (dashed line) monitored at λ ) 500
nm by UV-vis spectroscopy. UV-visible spectra of [Rh(COD)(PPh3)2]-
NTf2 and [Rh(CYD)(PPh3)2]NTf2 are provided in the Supporting
CYE is less soluble than CYD in both IL media so the system Information (Figures S-1 and S-2).
will tend to become biphasic during the course of the reaction.
tel-00708600, version 1 - 15 Jun 2012
( )
E
this experiment, mainly because of the effect of precooling and E
∂∆Hmix QCYD
kinetics of demixing, as already discussed,33 and consequently HCYD ) ≈ (1)
∂nCYD nIL,p,T ∆nCYD
the solubility was determined at only two temperatures. Deter-
mination of the solubility of CYD in ionic liquids by liquid-vapor
equilibria measurements using a static apparatus35,36 was not where nCYD and nIL denote the quantity of CYD and IL,
E
possible due to the chemical properties of CYD, such as its respectively, ∆Hmix is the excess molar enthalpy of the entire
tendency to polymerize37 and its compatibility with o’ring system (enthalpy of mixing) and ∆nCYD is the quantity of solute
materials. per injection. ∆nCYD was calculated from the injected volumes
The solubility of CYD at 303.15 K, expressed in mole fraction and the density of CYD was obtained from the literature.40 In
is 0.42 in [C1C4Im][NTf2] and 0.34 in [C1C1C4Im][NTf2] (see these calculations, heat due to evaporation of the solute from
Figure 4), which could indicate less favorable interactions IL solution is assumed to be negligible. Hence, no correction
between CYD and the [C1C1C4Im] ion in comparison with the for the vapor pressure of the solute was made.
[C1C4Im] ion. In the temperature range studied, both systems Figure 5 represents partial molar excess enthalpies of CYD
CYD-[C1C4Im][NTf2] and CYD-[C1C1C4Im][NTf2] show a in both ILs at 303.15 K as a function of composition. A larger
behavior compatible with the existence of upper critical solution dispersion of the values (up to 3%), especially at higher
temperatures and very steep liquid-liquid equilibrium lines. concentration, was observed. At times, before each injection of
With this method the solubility of the ionic liquid in CYD could CYD into the IL, an exothermic effect was detected, mainly
not be detected, as it lies below the limit of detection. From originating from insufficient mixing of both components at the
measurements and from calculations using COSMO-RS,38 it is beginning of the injection period. As already discussed in the
known that the concentration of hydrocarbons in the ionic liquid literature41 in the case of aromatic hydrocarbons, the rate of
rich phase is very low, with mole fractions on the order of 10-4. dissolution of nonpolar, low density and low viscosity CYD in
In comparison with aromatic molecules in these ILs, CYD is the IL is slow and sometimes leads to formation of a solute-
much less soluble. For instance, the solubility of toluene at rich layer on the surface of the solution. This is often
303.15 K expressed in mole fraction is 0.73 in [C1C4Im][NTf2] accompanied by partial evaporation or even polymerization of
and 0.72 in [C1C1C4Im][NTf2],9 the solubility of benzene in CYD before the mixing process is complete, hence disturbing
[C1C4Im][NTf2] being even higher (mole fraction of 0.78 at the measurements. A large number of injections were made to
303.15 K).39 The small difference in solubility of toluene with ensure reliability. [The formation of a polymer was observed
respect to benzene can be attributed to the presence of an and confirmed by NMR analysis.37 This side reaction is
additional methyl group, whereas the much lower solubility of attributed to the fact that CYD has been distilled to eliminate
CYD is explained by less favorable interactions between the the stabilizer BHT before the physical chemical measurements.]
solute and the ILs. Disruption of the aromatic system changes The dependence of partial molar excess enthalpy of CYD in
the interactions between solute and ionic liquid and thus [C1C4Im][NTf2] and [C1C1C4Im][NTf2] on the mole fraction was
decreases the solubility. approximated by a linear regression, eq 2 and Figure 5. The
The energy involved in the interactions of different solutes partial excess molar enthalpies of CYD in both ILs at infinite
E,∞
dissolved in ILs can be assessed by measuring the excess dilution, HCYD , were obtained from eq 2 by setting xCYD ) 0
Catalytic Hydrogenation of 1,3-Cyclohexadiene J. Phys. Chem. B, Vol. 114, No. 24, 2010 8161
E
ane in [C1C2Im][NTf2]. As expected, the values of ∆Hmix for
CYD obtained here are intermediate between the values for
cyclohexene and benzene. Comparison of [C1C4Im][NTf2] and
[C1C1C4Im][NTf2] shows that introduction of an additional
methyl group on the C2 carbon of the imidazolium ring strongly
affects the interaction with molecules of CYD, which becomes
less favorable. To rationalize the differences in the solvation of
CYD in both ILs, and to complement the enthalpic data with
structural information, the microscopic structure of the mixtures
was investigated using NMR and molecular simulation.
Molecular Structure of the Reaction Media. The molecular
structure of the mixture of CYD and ILs, [C1C4Im][NTf2] and
[C1C1C4Im][NTf2], and therefore the sites of specific interactions
were studied by 1H NMR and ROESY experiments. The 1H
NMR chemical shifts of the CYD-IL mixtures at R ) 0.5 were
not significantly different from those of the neat CYD and ILs,
in both cases (Table S-1 in Supporting Information).
Figure 6. Excess molar enthalpies of the systems: CYD-[C1C4Im]-
[NTf2] (s) and CYD-[C1C1C4Im][NTf2] (---) vs the mole fraction of NOESY experiments exhibited very weak or null cross peak
CYD at 303.15 K. intensities. This was probably due to the fact that the quantity
ωτc was such that the NOE intensity was close to the null point.
TABLE 1: Excess Molar Enthalpies of Mixing of Organic Rotating frame NOE experiments (ROESY) allowed us to obtain
Solutes and Ionic Liquids at Mole Fraction of Solute 0.1
positive NOEs irrespective of the long rotational correlation time
E
ionic liquid solute ∆Hmix (J/mol) T (K) due to the high viscosity of the system. In 1H-1H ROESY
tel-00708600, version 1 - 15 Jun 2012
[C1C4Im][NTf2] 1,3-cyclohexadiene +85 ( 9 303.15 techniques based on space cross relaxations, the selective
[C1C1C4Im][NTf2] 1,3-cyclohexadiene +164 ( 10 303.15 irradiation of a proton group affects the intensities of integrals
[C1C4Im][NTf2] benzene -189.842 363.15 of all proton groups that are spatially close but not necessarily
[C1C4Im][NTf2] toluene -150.440 363.15 connected by chemical bonds. The method is based on the
[C1C4Im][NTf2] methylcyclohexane +435.640 363.15 assumption of short-range intermolecular distances (4-5 Å).43
[C1C2Im][NTf2] benzene -112.734 323.15
[C1C2Im][NTf2] toluene -115.639 298.15 The strength of the ROE signal is proportional to the inverse
[C1C2Im][NTf2] cyclohexene +356.534 323.15 sixth power of the distance between the atoms, I ∝ 1/r.6 In the
[C1C2Im][NTf2] cyclohexane +394.034 323.15 liquid state, this relation is possible if the intermolecular
association is tight enough to turn the intermolecular relaxation
and were found to be 0.857 kJ mol-1 for CYD-[C1C4Im][NTf2] into “intramolecular” within the ion pair or ion-molecule
and 1.629 kJ mol-1 for CYD-[C1C1C4Im][NTf2]. association.41 Highly structured, bulk ionic liquids seem to fulfill
these requirements, thus allowing the use of the intermolecular
E
ROESY to derive lower limits for interionic distances. Conse-
HCYD (kJ/mol) ) a + bxCYD (2) quently, the intensity of the integrals (ICYD-IL) could be
considered as roughly inversely proportional to the intermo-
By integrating eq 1 and taking into account eq 2, we lecular distances (Figure S-4 and S-5, Supporting Information).
E
calculated the excess molar enthalpy of mixing ∆Hmix by Molecular mechanics calculations have been performed using
the SYBYL software with the TRIPOS force field developed
nCYD by Clark et al.44 on isolated pairs of CYD and imidazolium
E
∆Hmix (kJ/mol) ) ∆Hmix ) ∫ 0
E
HCYD dnCYD /(nCYD + nIL) cations. Intermolecular distances were fixed using results from
) (a + b)xCYD + b(1 - xCYD)ln(1 - xCYD) NMR ROESY experiments and the energy of both systems,
(3) CYD-[C1C4Im] and CYD-[C1C1C4Im], were minimized. The
average distances between molecules of CYD and the imida-
From Figure 6, it can be seen that the excess molar enthalpies zolium cations were determined from the geometry at the
of mixing for the binary mixtures of CYD and [C1C4Im][NTf2] potential energy minimum. CYD is mainly located near the butyl
and [C1C1C4Im][NTf2] are positive over the studied range of chain of [C1C4Im][NTf2], the π bonds being located closer to
compositions. The values of excess molar enthalpy of mixing the imidazolium cation, a configuration similar to that previously
of several aromatic and nonaromatic solutes in ILs together with observed for toluene (Figure 7).9
our results are presented in Table 1. To the best of our Access to the microscopic structure of the mixtures in the
knowledge the energetics of solvation of organic solutes in condensed liquid phase is possible using molecular dynamics
[C1C1C4Im][NTf2] had never been studied; therefore, we could simulation with all atoms explicitly present and periodic
establish comparisons only for 3-alkyl-1-methylimidazolium boundary conditions to represent a vitually infinite system.
bis(trifluoromethylsulfonyl)imide, [C1CnIm][NTf2]. Negative Condensed-phase simulations take into account all the two-body
E interactions from the environment of each molecule or ion. The
values of ∆Hmix for benzene and toluene in [C1C4Im][NTf2]
indicate more favorable interactions between aromatic systems molecular simulation details are given in the Experimental
and imidazolium ILs than for methylcyclohexane, for which Section. In Figure 8 are plotted the site-site radial distribution
the enthalpy of mixing is positive. A similar effect was observed functionssthe probability of finding pairs of atoms at a given
in [C1C2Im][NTf2], where the aromatic solutes have more distance, compared to the averagesbetween the hydrogens on
favorable interactions leading to negative enthalpies of mixing Cβ of CYD, Hβ, and selected atoms of the [C1C4Im] and
and higher solubilities. Cyclohexene, due to the presence of a [C1C1C4Im] cations at R ) 0.5 (atoms are labeled as indicated
double bond, also has lower enthalpy of mixing than cyclohex- in Scheme 1). In both cases, there is a higher probability (a
8162 J. Phys. Chem. B, Vol. 114, No. 24, 2010 Campbell et al.
NTf2]. The strong hydrogen bond between C2-H and the anion
in [C1C4Im][NTf2] prevents such an interaction with CYD. This
observation is consistent with the distances between the IL
cations and CYD obtained from ROESY experiments and
molecular mechanics calculations on isolated pairs, as seen in
Figure 7.
Detailed structural features are much better perceived in
3-dimensional spatial distribution functions. In Figure 10 is
represented the distribution of ions around a CYD molecule.
In blue is plotted the iso-surface corresponding to a local density
of twice the average density of C2 carbon atoms of the
imidazolium cations. Cation headgroups (the blue regions) are
located above and below the CYD ring, interacting preferentially
with the π-system of CYD. It can be seen that the terminal
Figure 7. Representation of molecular positions in saturated solutions carbons of the alkyl side chain, C9 (the gray regions), surround
of CYD-[C1C4Im][NTf2] and CYD-[C1C1C4Im][NTf2] from ROESY the CYD molecule. Oxygen atoms of the NTf2- anion (plotted
NMR extrapolation. in red) are found in the plane of CYD interacting with the
hydrogen atoms of the double bonds. Similar results have been
previously obtained for toluene in the same ILs.9 This shows
that although in CYD the π-system is smaller than that of a
fully aromatic system, it still determines the structure of the
solvation shell in ILs with cations positioned above and below
the plane of CYD and anion in the plane.
tel-00708600, version 1 - 15 Jun 2012
) 1.9. The availability of H2 for the hydrogenation reaction is comparable within the same order of magnitude. This is
also dependent on the diffusivity of the gas in both ILs. No commensurate with the expected precision of the molecular
experimental data were found on the diffusivity of hydrogen in simulation results.27 All these results indicate that the difference
tel-00708600, version 1 - 15 Jun 2012
ILs, but it is expected that it will also be inversely proportional in reactivity of both systems at R ) 0.5 can be mainly assigned
to the viscosity of the liquid medium since H2 interacts weakly to the difference in viscosity and therefore in mobility of the
with ILs (its solubility is very low45,46). Molecular simulation molecules, namely, CYD in the exchange reaction and both H2
results for the diffusivity of H2 in [C6C1Im][NTf2] at different and CYD in the hydrogenation reaction.
temperatures have recently been published.47 We have calculated
that the increase of diffusivity of hydrogen and the decrease in 4. Conclusion
viscosity of [C6C1Im][NTf2] at the different temperatures48 are
In this work, we studied the impact of two ionic liquid
solvents on the rates of two reaction steps of the catalytic
hydrogenation of 1,3-cyclohexadiene (CYD) with [Rh(COD)-
(PPh3)2]NTf2: the ligand exchange [Rh(COD)(PPh3)2]NTf2 to
[Rh(CYD)(PPh3)2]NTf2, and the catalytic hydrogenation of
CYD itself (COD ) 1,5-cyclooctadiene). It was found that
the iso-surface corresponding to a local density of 4 times the average GC. UV-vis spectra of [Rh(COD)(PPh3)2]NTf2 and
density of oxygen atoms from the NTf2- anion. [Rh(CYD)(PPh3)2]NTf2, ROESY and DOSY NMR pulse se-
quences, 1H NMR shifts and ROESY NMR spectra of the
TABLE 2: Density and Viscosity, at 298.15 K and
mixture of CYD in [C1C4Im][NTf2] and [C1C1C4Im][NTf2]. This
Atmospheric Pressure, of Mixtures of CYD-[C1C4Im][NTf2]
and CYD-[C1C1C4Im][NTf2] at R ) 0.5 material is available free of charge via the Internet at http://
pubs.acs.org.
xIL R F (g cm-3) η (mPa s)
CYD-[C1C4Im][NTf2] References and Notes
1.000 0.000 1.4375 48.45
0.667 0.498 1.3597 24.87 (1) Wasserscheid, P.; Welton, T. Ionic Liquids in Synthesis; Wiley
VCH: Weinheim, Germany, 2008.
CYD-[C1C1C4Im][NTf2] (2) Parvulescu, V. I.; Hardacre, C. Chem. ReV. 2007, 107, 2615.
1.000 0.000 1.4177 105.00 (3) Dyson, P. J.; Zhao, D. Hydrogenation. Multiphase Homogeneous
0.662 0.510 1.3442 47.13 Catalysis; Wiley-VCH: Weinheim, 2005; Vol. 2, pp 494.
(4) Olivier-Bourbigou, H.; Vallee, C. Multiphase Homogeneous Ca-
talysis; Wiley-VCH: Weinheim, 2005; Vol. 2, pp 413-431.
both steps are twice as fast in 1-butyl-3-methylimidazolium (5) Hintermair, U.; Gutel, T.; Slawin, A. M. Z.; Cole-Hamilton, D. J.;
bis(trifluoromethylsulfonyl)imide, [C1C4Im][NTf2], than in 1-bu- Santini, C. C.; Chauvin, Y. J. Organomet. Chem. 2008, 693, 2407.
(6) Dupont, J.; Suarez, P. A. Z. Phys. Chem. Chem. Phys. 2006, 8,
tyl-2,3-dimethylimidazolium bis(trifluoromethylsulfonyl)imide, 2441.
[C1C1C4Im][NTf2]. The rate-determining step in both ionic (7) Dupont, J.; Suarez, P. A. Z.; De Souza, R. F.; Burrow, R. A.;
liquids is the hydrogenation of CYD step. Kintzinger, J.-P. Chem.sEur. J. 2000, 6, 2377.
(8) Lachwa, J.; Bento, I.; Duarte, M. T.; Canongia Lopes, J. N.; Rebelo,
Molecular dynamics simulations and NMR experiments L. P. N. Chem. Commun. 2006, 2445.
indicate that in both ILs, for a molar ratio CYD/IL equal to (9) Gutel, T.; Santini, C. C.; Padua, A. A. H.; Fenet, B.; Chauvin, Y.;
0.5, CYD is solvated preferentially in lipophilic regions (in close Canongia Lopes, J. N.; Bayard, F.; Costa Gomes, M. F.; Pensado, A. S. J.
proximity to the alkyl side chains of the cations) as already Phys. Chem. B 2009, 113, 170.
(10) Mele, A.; Romano, G.; Giannone, M.; Ragg, E.; Fronza, G.; Raos,
observed for saturated hydrocarbons and the methyl group of G.; Marcon, V. Angew. Chem., Int. Ed. 2006, 45, 1123.
toluene. In addition, in [C1C1C4Im][NTf2] the probability of (11) Padua, A. A. H.; Costa Gomes, M. F.; Canongia Lopes, J. N. A.
finding CYD near the C2 of the cation is higher than in Acc. Chem. Res. 2007, 40, 1087.
(12) Triolo, A.; Russina, O.; Bleif, H.-J.; Di Cola, E. J. Phys. Chem. B
[C1C4Im][NTf2], in agreement with the shorter CYD- 2007, 111, 4641.
[C1C1C4Im] distances determined by ROESY NMR. On the (13) Canongia Lopes, J. N.; Padua, A. A. H. J. Phys. Chem. B 2006,
other hand, a higher solubility of CYD in [C1C4Im][NTf2] and 110, 3330.
(14) Canongia Lopes, J. N.; Costa Gomes, M. F.; Padua, A. A. H. J.
smaller positive enthalpies of mixing for the CYD-[C1C4Im]- Phys. Chem. B 2006, 110, 16816.
[NTf2] system in comparison with CYD-[C1C1C4Im][NTf2] (15) Hunt, P. A. J. Phys. Chem. B 2007, 111, 4844.
indicate more favorable interactions between CYD and the (16) Magna, L.; Chauvin, Y.; Niccolai, G. P.; Basset, J.-M. Organo-
[C1C4Im] cation than with the [C1C1C4Im] cation. These metallics 2003, 22, 4418.
(17) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976, 98, 4450.
thermodynamic factors concur with the differences in catalytic (18) Blesic, M.; Canongia Lopes, J. N.; Padua, A. A. H.; Shimizu, K.;
activity, although they cannot fully explain the differences Costa Gomes, M. F.; Rebelo, L. P. N. J. Phys. Chem. B 2009, 113, 7631.
observed. Association between macroscopic thermodynamic (19) Izatt, R. M.; Terry, R. E.; Haymore, B. L.; Hansen, L. D.; Dalley,
N. K.; Avondet, A. G.; Christensen, J. J. J. Am. Chem. Soc. 1976, 98, 7620.
information, thermophysical data, and microscopic structural (20) Liu, Y. F.; Sturtevant, J. M. Protein Sci. 1995, 4, 2559.
information is necessary to fully explain the differences in (21) Briggner, L. E.; Wadso, I. J. Biochem. Biophys. Methods 1991,
reactivity found in both IL media. 22, 101–118.
Pure [C1C1C4Im][NTf2] has a higher viscosity than pure (22) Cutting, B.; Ghose, R.; Bodenhausen, G. J. Magn. Reson. 1999,
138, 326.
[C1C4Im][NTf2], and the same relative values are observed in (23) Canongia Lopes, J. N.; Deschamps, J.; Padua, A. A. H. J. Phys.
the mixtures with CYD. These differences in viscosity induce Chem. B 2004, 108, 2038.
Catalytic Hydrogenation of 1,3-Cyclohexadiene J. Phys. Chem. B, Vol. 114, No. 24, 2010 8165
(24) Canongia Lopes, J. N.; Padua, A. A. H. J. Phys. Chem. B 2004, (37) Natori, I.; Imaizumi, K.; Yamagishi, H.; Kazunori, M. J. Polym.
108, 16893. Sci., Part B 1998, 36, 1657.
(25) Canongia Lopes, J. N.; Shimizu, K.; Padua, A. A. H.; Umebayashi, (38) Domanska, U.; Pobudkowska, A.; Eckert, F. Green Chem. 2006,
Y.; Fukuda, S.; Fujii, K.; Ishiguro, S. I. J. Phys. Chem. B 2008, 112, 9449. 8, 268.
(26) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. J. Am. Chem. (39) Lachwa, J.; Bento, I.; Duarte, M. T.; Canongia Lopes, J. N.; Rebelo,
Soc. 1996, 118, 11225. L. P. N. Chem. Commun. 2006, 2445.
(27) Maginn, E. J. Acc. Chem. Res. 2007, 40, 1200. (40) Letcher, T. M.; Marsicano, F. J. Chem. Thermodyn. 1974, 6, 509.
(41) Marczak, W.; Verevkin, S. P.; Heintz, A. J. Solution Chem. 2003,
(28) Smith, W.; Forester, T. R.; Todorov, I. T.; The DL_POLY molecular 32, 519.
simulation package, 2.20 ed.; T. D. P. m. s., Ed.; STFC Daresbury (42) Nebig, S.; Bolts, R.; Gmehling, J. Fluid Phase Equilib. 2007, 258,
Laboratorys, Warrington, U.K., 2007. 168.
(29) Chauvin, Y.; Mussmann, L.; Olivier, H. Angew. Chem., Int. Ed. (43) Frezzato, D.; Rastrelli, F.; Bagno, A. J. Phys. Chem.B 2006, 110,
1996, 34, 2698. 5676.
(30) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976, 98, 2134. (44) Clark, M.; Cramer, R. D., III; Van Opdenbosch, N. J. Comput.
(31) Costa Gomes, M. F. Unpublished results. Chem. 1989, 10, 982–1012. Sybil program version 7.0, Tripos Inc., 1699
(32) Camper, D.; Becker, C.; Koval, C.; Noble, R. Ind. Eng. Chem. Res. South Hanley Rd., St. Louis, MO 63144, U.S.A.
2006, 45, 445. (45) Jacquemin, J.; Costa Gomes, M. F.; Husson, P.; Majer, V. J. Chem.
(33) Cui, Y.; Biondi, I.; Chaubey, M.; Yang, X.; Fei, Z.; Scopelliti, R.; Thermodyn. 2006, 38, 490. Jacquemin, J.; Husson, P.; Majer, V.; Costa
Hartinger, C. G.; Li, Y.; Chiappe, C.; Dyson, P. J. Phys. Chem. Chem. Gomes, M. F. Fluid Phase Equilib. 2006, 240, 87.
Phys. 2010, 12, 1834. (46) Costa Gomes, M. F. J. Chem. Eng. Data 2007, 52, 472.
(34) Domañska, U.; Żołek-Tryznowska, Z.; Królikowski, M. J. Chem. (47) Shi, W.; Sorescu, D. C.; Luebke, D. R.; Keller, M. J.; Wickra-
Eng. Data 2007, 52, 1872. manyake, S. J. Phys. Chem. B 2010, 114, 6431.
(48) Marsh, K. N.; Brennecke, J. F.; Chirico, R. D.; Frenkel, M.; Heintz,
(35) Kato, R.; Krummen, M.; Gmehling, J. Fluid Phase Equilib. 2004, A.; Magee, J. W.; Peters, C. J.; Rebelo, L. P. N.; Seddon, K. R. Pure Appl.
224, 47. Chem. 2009, 81, 781.
(36) Husson, P.; Pison, L.; Jacquemin, J.; Costa Gomes, M. F. Fluid
Phase Equilib. 2010, DOI: 10.1016/j.fluid.2010.02.021. JP102941N
tel-00708600, version 1 - 15 Jun 2012
Journal of Catalysis 275 (2010) 99–107
Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat
a r t i c l e i n f o a b s t r a c t
Article history: Tailor-made and size-controlled ruthenium nanoparticles, RuNPs, of three distinct sizes between 1 and
Received 1 June 2010 3 nm are generated from the decomposition of (g4-1,5-cyclooctadiene)(g6-1,3,5-cyclooctatriene)
Revised 19 July 2010 ruthenium(0) [Ru(COD)(COT)], under H2 in 1-butyl-3-methylimidazolium bis(trifluoromethanesulpho-
Accepted 21 July 2010
nyl)imide, C1C4ImNTf2, by simply varying experimental conditions. Catalytic hydrogenation of 1,3-cyclo-
hexadiene, CYD, and cyclohexene, CYE, in C1C4ImNTf2, has been used as a probe for the relationship
between size and catalytic performance (activity and selectivity) of RuNPs. To allow comparison between
Keywords:
different reactions, all catalytic reaction mixtures were diligently prepared in order that the parameters
Nanoparticles
Ruthenium
such as substrate/catalyst and substrate/ionic liquid ratio, and therefore, viscosity and mass transport
Ionic liquids factors remained constant. It was found that the catalytic activity increases with the NP size, while high
Size effect selectivity is only observed with the smaller NPs. In addition, the studied RuNPs exhibit a high level of
Selective hydrogenation recyclability with neither loss of activity nor significant agglomeration.
Ó 2010 Elsevier Inc. All rights reserved.
0021-9517/$ - see front matter Ó 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcat.2010.07.018
100 P.S. Campbell et al. / Journal of Catalysis 275 (2010) 99–107
of 1,3-cyclohexadiene, CYD, and cyclohexene, CYE, as probes for carbon). Binding energies were determined with an accuracy of
the relationship between size and catalytic performance of tailor- ±0.2 eV. The data were analysed using Casa-XPS (v 2.3.13) employ-
made and size-controlled RuNPs, generated in the ionic liquid, all ing a Shirley background subtraction prior to fitting and a peak
other physicochemical variables being constant. shape with a combination of Gaussian and Lorentzian (30%
Lorentzian). High-resolution spectra were acquired in the region
2. Experimental of Ru 3p as the Ru 3d region overlaps with the C 2p region of the
residual ionic liquid.
2.1. Materials and methods
2.5. Preparation of catalytic experiments
All operations were performed in the strict absence of oxygen
and water under a purified argon atmosphere using glovebox (Jac- In Table 1 are collected all data concerning the studied hydroge-
omex or MBraun) or vacuum-line (Schlenk) techniques. The ionic nation reaction. In columns 1–3, size and dispersion values of
liquid, C1C4ImNTf2 [30], and the complex, [Ru(COD)(COT)] [31], (Ru0), (Ru25) and (Ru50) are reported.
were synthesised as reported. The halide content of the ionic liquid Each solution of NP was produced from the decomposition of
was under 200 ppm (E.A.) and water under 5 ppm (limit of Karl 43.0 mmol Lÿ1 solution of Ru(COD)(COT) as described in Section
Fischer titration). Elemental analyses were performed at the CNRS 2.2 and shown in column 4.
Central Analysis Department of Solaize. For each different temperature of decomposition, a different
1-Methylimidazole (>99%) was purchased from Aldrich and dis- size of NP is obtained: at 0 °C, 1.1 nm (Ru0); at 25 °C, 2.3 nm
tilled prior to use. Chlorobutane (>99%, Aldrich) and lithium bis(tri- (Ru25); and at 50 °C, 2.9 nm (Ru50), shown in columns 1 and 2.
fluoromethanesulfonyl)imide (Solvionic) were used without The value of dispersion, D, describing the ratio of surface sites
further purification. Rus to total number of ruthenium atoms, varies with the NP size
and is given for each size of NP in column 3. Using this and the
tel-00708600, version 1 - 15 Jun 2012
Table 1
Calculations for the composition of the catalytic systems. Column 1 – name of catalyst, column 2 – average RuNP diameter measured by TEM, column 3 – calculated dispersion,
column 4 – initial Ru(COD)(COT) concentration, column 5 – consequent Rus concentration, column 6 – volume of IL/RuNP solution, column 7 – volume of pure IL added, column 8
– consequent number of moles of Rus in the 5 mL mixture, column 9 – mass of IL, column 10 – mass of substrate, column 11 – substrate/catalyst ratio, column 12 – substrate/IL
ratio.
1 2 3 4 5 6 7 8 9 10 11 12
d D [Ru] (mmol Lÿ1) [Rus] (mmol Lÿ1) Vol. IL–RuNP Vol. IL pure (mL) Rus/10ÿ5 (mol) m (IL/g) m (CYD/g) CYD/Rus CYD/IL
(nm) (%) solution (mL)
Ru0 1.1 82 43.0 35.2 2.42 2.58 9.52 7.0 0.78 105 0.59
Ru25 2.3 53 43.0 22.9 3.74 1.26 9.52 7.0 0.78 105 0.59
Ru50 2.9 43 43.0 18.5 5.00 – 9.52 7.0 0.78 105 0.59
The composition of the mixture was determined by gas-phase CYD may be partially hydrogenated with high selectivity by molec-
chromatography using toluene as the internal standard. ular catalysts due to the reduced miscibility of cyclohexene (CYE)
in the medium [27,32,33]. Full hydrogenation would lead to cyclo-
2.7. Product quantification hexane (CYA), Scheme 1.
volume was 1 lL. The programme was as follows: initial temper- has been performed by Dupont’s group in ionic liquids due to the
ature 70 °C for 13.5 min; ramp 40 °C/min to 250 °C, hold 2 min. difference in solubility of the partially hydrogenated product
[34]. The same group has also described the possibility of extract-
ing cyclohexene during benzene hydrogenation using this solubil-
2.8. Density and viscosity
ity difference [35]. For this reason, solubilities of CYD, CYE and CYA
are measured. It is found that the solubility of the hydrogenated
The mixtures of IL and CYD at different compositions were pre-
products (6 ± 1% wt – CYE, 4 ± 1% wt – CYA) is much lower than
pared gravimetrically following the procedure already described
that of CYD (12 ± 2% wt); therefore, the medium may tend to a
[27]. The viscosity of the mixture was measured at 298.15 K
biphasic system during the course of the reaction. As a result, the
(controlled to within ±0.005 K and measured with the accuracy
collection of aliquots from a single batch would render inaccurate
better than ±0.05 K) using a rolling-ball viscometer from Anton
results. Consequently, each point recorded in this work corre-
Paar, model AMVn [27]. The overall uncertainty of the viscosity
sponds to a separate experiment, quenched after time t by opening
is estimated as ±2.0%. The densities of the mixtures, necessary to
the reaction vessel to air, thus releasing the hydrogen and dissolv-
calculate the viscosities, were measured in an Anton Paar vibrating
ing the catalytic system entirely in a 1 M solution of toluene in ace-
tube densimeter model 512 P, at 298.15 K (measured by a cali-
tonitrile for gas-phase chromatography.
brated PRT with an accuracy of ±0.02 K). The overall uncertainty
of the density is estimated as ±0.01%.
3.3. Viscosity
Table 2
Density, q, and viscosity, g, of CYD–IL mixtures of different compositions. xIL = molar
fraction of IL, R = molar ratio CYD/IL.
H2 R xIL q (g cmÿ3) g (m Pa s)
+ 0.000 1.000 1.4376 ± 0.0001 48.5 ± 0.4
0.100 0.909 1.4202 ± 0.0001 44 ± 1
0.200 0.833 1.3999 ± 0.0001 37.0 ± 0.4
0.300 0.769 1.3874 ± 0.0003 33.3 ± 0.3
CYD CYE CYA 0.397 0.716 1.3718 ± 0.0001 31.0 ± 0.3
0.498 0.667 1.3597 ± 0.0001 24.8 ± 0.3
Scheme 1. 1,3-Cyclohexadiene and its hydrogenation products.
102 P.S. Campbell et al. / Journal of Catalysis 275 (2010) 99–107
Frequency
3.4. Catalyst characterisation
80
It has been previously demonstrated that the size of RuNPs
60
generated from the decomposition of [Ru(COD)(COT)] under H2
[36], may be governed by the degree of self-organisation of the 40
imidazolium-based ionic liquid in which they are formed: the
more structured the ionic liquid, the smaller the size [16]. Follow- 20
ing previously described methods, RuNPs are synthesised at 0 °C,
0
25 °C, 50 °C and 75 °C in an attempt to obtain a selection of mono-
1 2 3 4 5
disperse sizes of RuNP in the same IL.
Size/ nm
3.4.1. TEM Fig. 2. Comparative size distribution histograms for RuNPs prepared in C1C4Im NTf2
Analysis of the suspensions obtained by transition electron at different temperatures.
microscopy allows the determination of the sizes generated:
1.1 ± 0.2 nm, 2.3 ± 0.3 nm, 2.9 ± 0.4 nm and 3.1 ± 0.7 nm, for RuNPs interplanar distances match with the hcp crystalline phase of Ru
generated at 0 °C (Ru0), 25 °C (Ru25), 50 °C (Ru50) and 75 °C (Ru75), (see Supplementary information). For Ru0, only a larger NP of
2 nm is observed by HREM, probably due to the difficulty in
tel-00708600, version 1 - 15 Jun 2012
respectively, Fig. 1. As can be seen from the TEM image of Ru75 and
the consequent size distribution histogram Fig. 2, the size of these observing the smallest NPs with limited contrast although may
NPs does not vary significantly compared to those of Ru50 although be indicative of a lower degree of crystallinity in very small RuNPs,
a poorer size control (wider distribution) is apparent. For this rea- as already observed by reverse Monte Carlo simulations.[37].
son, these NPs are not used in catalytic tests. High-resolution elec-
tron microscopy reveals the crystalline nature of the RuNPs formed 3.4.2. XPS
through elucidation of the crystal planes. The Fourier transform In order to establish the oxidation state of the RuNPs, X-ray
images of the HREM have been exploited and indicate that the photoelectron spectroscopy (XPS) is performed. Due to the weak
Fig. 1. Transition electron micrograph of RuNPs and high-resolution electron micrograph examples showing crystallinity for Ru0 (top left), Ru25 (top right), Ru50 (bottom left)
and Ru75 (bottom right).
P.S. Campbell et al. / Journal of Catalysis 275 (2010) 99–107 103
Ru 3p1/2
490 480 470 460 450 490 480 470 460 450 490 480 470 460 450
Binding Energy / eV Binding Energy / eV Binding Energy / eV
Fig. 3. XPS spectra of the Ru 3p region after NP filtration onto SiO2, experimental data and fitted peaks. Left Ru0, middle Ru25 and right Ru50.
100%
concentration of the solution and the penetration limit of the
90%
X-rays in the solution (maximum depth of 10 nm), no peaks corre-
sponding to Ru binding energies are observed when the analyses 80%
are carried out directly on the RuNP/IL solutions. Samples are 70%
Dispersion
therefore prepared by filtering the RuNPs onto silica under inert 60%
atmosphere and eliminating as much IL as possible. The resulting
tel-00708600, version 1 - 15 Jun 2012
50%
spectra of the Ru 3p region are depicted in Fig. 3. It is clear that
in each case, fine peaks are observed, indicating the presence of 40%
only one Ru species. The low 3p3/2 binding energy observed in each 30%
case, 460.3 eV, and doublet separation of 22.2 eV correspond clo- 20%
sely to metallic zero-valent ruthenium, often reported with a
10%
3p3/2 binding energy of around 461 eV [38]. The small difference
may be attributed to the presence of small crystallites, which tend 0%
0 0.5 1 1.5 2 2.5 3 3.5
to exhibit lower binding energies than bulk metal. Indeed, as re-
cently shown for AuNPs [39], the d band narrows with decreasing Diameter / nm
particle size and shifts towards the Fermi level. Fig. 4. Curve of dispersion D against mean diameter of crystalline hcp RuNPs.
Table 3
Conversion, selectivity, turnover number and turnover frequency for the hydrogenation of CYD at 30 °C. Experiments 1–3 using 1.2 bar H2. Experiments 4 and 5 using 4 bar H2.
Experiment number Catalyst (nm) Pressure H2 (bar) Conversion at 90 min (%) Selectivity CYE (%) TON TOF (hÿ1) Size after catalysis (nm)
1 Ru0 (1.1) 1.2 66 ± 5 97 70 ± 5 46 ± 3 1.3 ± 0.4
2 Ru25 (2.3) 1.2 75 ± 5 86 79 ± 5 53 ± 3 2.1 ± 0.5
3 Ru50 (2.9) 1.2 83 ± 5 80 87 ± 5 58 ± 3 2.7 ± 0.5
4 Ru0 (1.1) 4.0 57 ± 5 92 59 ± 5 40 ± 3 –
5 Ru50 (2.9) 4.0 73 ± 5 80 77 ± 5 51 ± 3 –
CYD, the selectivity in CYE versus CYA drops from 97% to 80% when
the RuNP size increases from 1.1 to 2.9 nm.
3.8. Catalytic selectivity Our hypothesis is based on idealised particle shape, which is not
likely to exist in reality. Nonetheless, it is widely accepted that
In the hydrogenation of CYD, CYE is obtained as the major prod- large NPs are more likely to present larger open facets where pla-
uct. Interestingly, the selectivity for CYE diminishes with increas- nar p-coordination of diene substrates can occur, whereas small
ing NP size. Indeed, for Ru0, selectivity for CYE is 100% at low NPs are often reported to be amorphous, therefore presenting no
conversion and only slightly diminishes at high conversion (97%). open facets, making this planar coordination even less likely [41].
In contrast, for Ru50, the hydrogenation is unselective even at In studies of CO hydrogenation on RhNPs, the difference in reac-
low conversion, Fig. 5. tivity with size was related to the increasing probability of finding
Assuming highly crystalline particles with hcp structure, a par- step sites with increasing NP size [49–51]. However, for RuNPs of
ticle of diameter 1.1 nm would have the vast majority of catalytic less than 3 nm, as reported here, calculations have shown that such
surface atoms occupying vertex or edge positions. Such vertex step sites are not likely to exist [52].
ruthenium atoms Ruv, which under H2 atmosphere are ligated by The highly selective hydrogenation of 1,3-cyclohexadiene to
hydrides, may coordinate one C@C double bond of CYD. The prod- cyclohexene has also been performed with PdNPs in organic and
uct of the subsequent hydrogenation is CYE, which must undergo a IL media [33,53,54]. The high selectivity results from the intrinsic
second coordination to give the fully hydrogenated CYA. Similarly, properties of Pd metal and its small NP size [55].
for a larger particle of average diameter 2.9 nm assuming high
crystallinity and an hcp structure, it is evident that most of the cat- 3.9. Hydrogen effect
alytically active surface ruthenium atoms are found in facial posi-
tions, Ruf. Indeed, here, such crystallinity has already been According to the literature, the hydrogenation of olefins in ILs is
observed by HREM, Fig. 1. Ruf may hydrogenate the olefin using often biphasic in its nature [56–59], due to the poor solubility of
the mechanism previously discussed, but due to the planar hydrogen and olefins in these media [60–63]. Therefore, the diffu-
arrangement of Ruf, another mechanism may be envisaged involv- sion process of the substrate or H2 may limit the rate of
ing the double coordination of the diene, as generally found during hydrogenation.
To find out whether H2 is in fact a rate-limiting reagent, exper-
iments varying the H2 pressure are performed. Increasing H2 pres-
sure to 4 bars in the case of Ru0 and Ru50 is seen to affect neither
the activity nor selectivity in a substantial manner, Table 3, Exper-
iments 4 and 5. This is similar to results reported by Dupont’s
group who observed that the reaction rate does not depend on
the H2 pressures in C1C4ImBF4 [64].
In parallel, it is generally reported that a higher H2 concentra-
tion should influence the activity of the large rather than small par-
ticles, as the H2 storage capacity is related to the particle volume;
therefore, large particles may experience an increase in the avail-
ability of subsurface hydrogen [44].
In reality, little difference in activity is observed in either case.
This proves that the rate is not dictated by the availability and
adsorption of H2, in agreement with observations of labile surface
hydrides on the NP surface [17], but by the mobility and absorption
of the substrate [26]. Indeed, it is highly likely that the surface of
Fig. 5. Selectivity for cyclohexene as a function of conversion for the three different the NPs is already saturated with adsorbed H2 at the temperature
sizes of RuNPs. of the reaction [17].
P.S. Campbell et al. / Journal of Catalysis 275 (2010) 99–107 105
tel-00708600, version 1 - 15 Jun 2012
Fig. 6. SYBYL representations of CYD coordinated to the face of highly crystalline RuNPs of mean diameter 1.3 nm (left) and 2.8 nm (right).
Fig. 7. Transition electron micrographs of RuNPs after CYD hydrogenation for Ru0 (left), Ru25 (middle) and Ru50 (right).
Table 5
Recycling of the catalyst Ru0. Conversion, selectivity, turnover number and turnover frequency for the hydrogenation of CYD at 30 °C under 1.2 bar H2. Products removed under
vacuum after each run and analysed by GC.
Experiment number Cycle Conversion at 90 min (%) Selectivity CYE (%) TON TOF (hÿ1) Size after catalysis (nm)
1 1st 66 97 70 46 1.3 ± 0.4
9 2nd 73 95 76 51 –
10 3rd 69 94 73 49 –
11 4th 68 89 71 47 –
12 5th 64 86 67 45 –
13 6th 65 86 68 45 1.8 ± 0.5
3.10. Recycling the most selective catalyst, Ru0, recycling experiments are per-
formed, by extracting in vacuo and quantifying the volatiles after
TEM images of the reaction medium after hydrogenation, Fig. 7, each 90-min run. More CYD is then added for hydrogenation. From
show that the average size measured (Table 3) does not differ the results, Table 5 and Fig. 8, it can be seen that both the activity
greatly from the original size, in accordance with the stability of and selectivity remain high after five recycles, diminishing only
RuNPs in ILs under molecular hydrogen [17]; however, the size dis- slightly with each run. This small decrease of course is attributable
tribution is larger, probably as an effect of stirring [65]. to the gradual coalescence of the NPs, leading to a diminution in
The apparent resistance to coalescence of the NPs means that the number of active surface sites and larger, less selective NPs.
they may be tested for their recyclability. Consequently, using Indeed, TEM images obtained of the NPs after all recycling
106 P.S. Campbell et al. / Journal of Catalysis 275 (2010) 99–107
55 100 References
80
[5] A.Z. Moshfegh, J. Phys. D: Appl. Phys. 42 (2009) 233001. 233032.
40 [6] G.A. Somorjai, J.Y. Park, Top. Catal. 49 (2008) 126–135.
[7] H. Boennemann, K.S. Nagabhushana, in: B. Corain, G. Schmid, N. Toshima
70 (Eds.), Metal Nanoclusters in Catalysis and Materials Science: The Issue of Size
35 Control, Elsevier B.V., Amsterdam, pp. 21–48.
[8] H. Bönnemann, K.S. Nagabhushana, R.M. Richards, in: D. Astruc (Ed.),
60 Nanoparticles and Catalysis, Wiley-VCH, Weinheim, 2008, pp. 49–92.
30
[9] H. Tada, T. Kiyonaga, S. Naya, Chem. Soc. Rev. 38 (2009) 1849.
[10] D. Astruc, Nanoparticles and Catalysis, Wiley-VCH, Weinheim, 2008.
25 50 [11] J. Dupont, J.D. Scholten, Chem. Soc. Rev. 39 (2010) 1780–1804.
1 2 3 4 5 6 [12] J. Dupont, Nanoparticles and Catalysis, Wiley-VCH, Weinheim.
[13] T. Gutel, C.C. Santini, K. Philippot, A. Padua, K. Pelzer, B. Chaudret, Y. Chauvin,
Cycle J.-M. Basset, J. Mater. Chem. 19 (2009) 3624–3631.
[14] A.A.H. Padua, M.F. Costa Gomes, J.N.A. Canongia Lopes, Acc. Chem. Res. 40
Fig. 8. Evolution of TOF and selectivity with catalyst recycling. (2007) 1087–1096.
[15] J.N. Canongia Lopes, M.F. Costa Gomes, A.A.H. Padua, J. Phys. Chem. B 110
(2006) 16816–16818.
experiments showed that the NPs undergo coalescence to attain an
[16] T. Gutel, J. Garcia-Anton, K. Pelzer, K. Philippot, C.C. Santini, Y. Chauvin, B.
average diameter of 1.8 ± 0.5 nm, approaching the size of Ru25 and Chaudret, J.-M. Basset, J. Mater. Chem. 17 (2007) 3290–3292.
of course presenting a similar selectivity. [17] P.S. Campbell, C.C. Santini, D. Bouchu, B. Fenet, K. Philippot, B. Chaudret, A.A.H.
Padua, Y. Chauvin, Phys. Chem. Chem. Phys. 12 (2010) 4217–4223.
[18] C. Vollmer, E. Redel, K. Abu-Shandi, R. Thomann, H. Manyar, C. Hardacre, C.
tel-00708600, version 1 - 15 Jun 2012
[54] J. Huang, T. Jiang, H. Gao, B. Han, Z. Liu, W. Wu, Y. Chang, G. Zhao, Angew. [60] P.J. Dyson, G. Laurenczy, C.A. Ohlin, J. Vallance, T. Welton, Chem. Commun.
Chem. Int. Ed. 43 (2004) 1397–1399. (2003) 2418–2419.
[55] G. Ertl, H. Knozinger, J. Weitkamp (Eds.), Handbook of Heterogeneous [61] J. Jacquemin, M.F. Costa Gomes, P. Husson, V. Majer, J. Chem. Thermodyn. 38
Catalysis, Wiley-VCH, Weinheim, 1997. (2006) 490–502.
[56] J. Dupont, G.S. Fonseca, A.P. Umpierre, P.F.P. Fichtner, S.R. Teixeira, J. Am. [62] J. Jacquemin, P. Husson, V. Majer, M.F. Costa Gomes, Fluid Phase Equilib. 240
Chem. Soc. 124 (2002) 4228–4229. (2006) 87.
[57] P.J. Dyson, Appl. Organomet. Chem. 16 (2002) 495–500. [63] J. Jacquemin, P. Husson, V. Majer, M.F. Costa Gomes, J. Solution Chem. 36
[58] P.J. Dyson, in: a.M.T.J. Mc Cleverty (Ed.), Comprehensive Coordination (2007) 967–979.
Chemistry II, Elsevier, Amsterdam, pp. 557–566. [64] P. Migowski, D. Zanchet, G. Machado, M.A. Gelesky, S.r.R. Teixeira, J. Dupont,
[59] P.J. Dyson, D. Zhao, Multiphase Homogeneous Catalysis, Wiley-VCH, Phys. Chem. Chem. Phys. 12 (2010) 6826–6833.
Weinheim, pp. 494–511. [65] D. Li, R.B. Kaner, J. Am. Chem. Soc. 128 (2006) 968–975.
tel-00708600, version 1 - 15 Jun 2012
View Online
cyclooctadiene) dissolved in imidazolium ionic liquids (ILs) undergo reduction and decomposition,
respectively, to afford stable ruthenium and nickel metal(0) nanoparticles (Ru(0)-NPs and Ni(0)-NPs)
in the absence of classical reducing agents. Depending on the case, the reduction/auto-decomposition is
promoted by either the cation and/or anion of the neat imidazolium ILs.
NPs via chemical routes. This can be controlled simply by The formation of Ru(0)-NPs using standard protocol with
08 October 2010
incorporating coordinating groups,5–7 varying the coordination [Ru(COD)(2-methylallyl)2] in IL (i.e. under 4 bar hydrogen at
50 C)7,14 was accompanied by the evolution of small amounts of
Published onversion
of reducing reagents such as molecular hydrogen gas, complex The volatiles were thus analysed online in the gas phase with
hydrides (NaBH4 and LiAlH4), hydrazine,1–4 alcohols,16 and a mass gas-analyser. Indeed, compounds were detected with
thiols.17 In some cases, the IL itself can carry the reducing agent, masses that could be assigned to small molecules formed due to
e.g. hydroxylated imidazolium salts,18,19 and depending on the the fragmentation of the imidazolium ring, such as acetonitrile/
redox potential of the metal precursor, the imidazolium cation isocyanomethane (M+ ¼ 41), methylamine (M+ ¼ 29, MeN),
may even undergo oxidation. For example, in the case of Au(III), ethylene (M+ ¼ 28) and hydrogen cyanide (M+ ¼ 27). Moreover,
the imidazolium cation itself can act as a reducing agent to yield the observed C4-fragment signals (M+ ¼ 56–58) provide strong
prismatic gold particles in BMI$PF6.20 evidence for the decomposition of the ruthenium complex
The preparation of M-NPs in ILs by simple decomposition of [Ru(COD)(2-methylallyl)2] involving the formation of isobutene/
organometallic compounds in their formal zero oxidation states isobutane from the 2-methylallyl-ligand in the Ru(II) complex.
is invariably performed in the presence of hydrogen11,21,22 or Interestingly, the fragmentation of the IL seems to occur only
under thermal8,15,23 or photolytic9 conditions. during the reduction of Ru complex, but it is not promoted by
In this work, we report a novel approach for the synthesis of the Ru(0)-NPs. The gaseous by-products were exclusively
Ru(0)- and Ni(0)-NPs in imidazolium ILs, which act as detected during the NP synthesis, but not when Ru(0)-NPs were
incommensurably mild reducing/decomposing reagents for the stirred in BMI$NTf2 and HM2I$NTf2 for a prolonged duration
organometallic complexes [Ru(COD)(2-methylallyl)2] (several days) under identical conditions. A similar observation
for the imidazolium ring fragmentation has previously been
made, during the ultrasonic irradiation of imidazolium chloride
a
Laboratory of Molecular Catalysis, Institute of Chemistry, Universidade
Federal do Rio Grande do Sul, Av. Bento Gonçalves 9500, 91501-970
Porto Alegre, RS, Brazil. E-mail: jairton.dupont@ufrgs.br; Fax: +55 51
33087304; Tel: +55 51 33086321
b
Universit
e de Lyon, Institut de Chimie de Lyon, C2P2, UMR 5265
CNRS—ESCPE Lyon, 43 bd du 11 Novembre 1918, F-69626
Villeurbanne Cedex, France. E-mail: santini@cpe.fr
c
Centro de Tecnologias Estrategicas do Nordeste—CETENE, 50740-540
Recife, PE, Brazil
† Present address: Humboldt University Berlin, Brook-Taylor-Strasse 2,
12489 Berlin, Germany. E-mail: . E-mail: martin.prechtl@hu-berlin.de Fig. 1 Structure of the ILs used in this study.
at 135 C, causing degradation of ILs.24–26 However, herein the were similar to those of Ru(0)-NPs previously generated
observed decomposition of the imidazolium ring is not clearly using hydrogen gas as a reducing agent for the reduction
understood. of [Ru(COD)(2-methylallyl)2] or decomposition of
More interestingly, further investigations of the Ru(II)/IL [Ru(COD)(COT)] (COT ¼ 1,3,5-cyclooctatriene).7,11,14,22
reaction system revealed a more important result: the presence of The Ru(0)-NP formation led us to propose a reaction pathway
hydrogen gas as a reducing reagent for the Ru complex seems to in which the imidazolium ILs BMI$NTf2 and HM2I$NTf2 might
be obsolete. Stirring a mixture of the complex in HM2I$NTf2 act as reducing agents for [Ru(COD)(2-methylallyl)2]. Here the
under argon atmosphere at 50 C for a prolonged period (2 days) NTf2 anion would act as a nucleophile27,28 and attack the allylic-
resulted in a dark brown/black colloidal solution. Samples for ligand of the complex (Schemes 1 and 2).
TEM analysis were prepared by placing a small amount of the This stoichiometric reaction would cause the concomitant
Ru(0)-NPs dispersed in HM2I$NTf2 onto a holey carbon film reduction of the Ru(II) complex and subsequent decomposition
supported by a copper grid. The diameters of the particles in the of the IL. Consequently, the ruthenium would lose its ligand-
micrographs were measured using the software Sigma Scan Pro sphere and be reduced to ruthenium(0) atoms that coalesce
Jun 2012 | doi:10.1039/C0NR00574F
5. Size distribution histograms of the NPs were obtained by generating the Ru(0)-NPs. It is proposed that during the reduc-
measuring the diameter of randomly selected particles, resulting tion process of the Ru(II) complex, the first step involves a ligand
in the particle size of 2.0 0.3 nm (see Fig. 2) with a monomodal exchange between the COD and a stronger coordinating ligand,
distribution. occurring readily under the given reaction conditions.29 A
The crystalline structure of the particles was confirmed by HR- stronger coordinating ligand is easily provided in neat imidazo-
TEM micrographs, analysed using Gatan Digital Micrograph lium ionic liquid as solvent, where classical and abnormal
Software. By means of HR-TEM measurements it was possible N-heterocyclic carbenes (NHCs) formed in situ might act as
to obtain the Fourier transform images from which lattice a ligand.30,31 As a consequence, the NTf2 anion would ‘‘lose’’ its
15http://pubs.rsc.org
spacings of 2.04 Å and 2.13 Å were measured. These lattice counterion, thus enhancing its nucleophilicity. And this would
Downloaded on 10 November 2010
spacings correspond to the interplanar distances (1 0 1) and favour its subsequent attack on the allylic species of the Ru(II)
(0 0 2), respectively, of hcp Ru(0). Isolation of the Ru(0)-NPs for complex. The resulting ruthenium(0) metal atoms in the IL go on
analysis by XRD was not possible, corroborating previous to generate the M-NPs. The isobutene by-product was detected
reports.14,22 The size and size distribution of the ruthenium NPs by MS analysis of the gas phase (see above). However, attempts
1 - on
Fig. 2 Selected TEM image of Ru(0)-NPs (2.0 0.3 nm) in HM2I$NTf2 Scheme 2 Proposed reductive elimination of the allyl-ligand induced by
and the histogram of the NPs size distribution. the NTf2 anion (L ¼ ligand/solvent-IL).
we must find a way of inhibiting this auto-decomposition in the case of [Ru(COD)(2-methylallyl)2] the decomposition of
imidazolium ILs with short alkyl chains. Another strategy [Ni(COD)2] in BMI$BF4 occurred not spontaneously but very
attempted in order to circumvent the problem was to use an IL slowly at 25 C only under 4 bar of H2, once again highlighting
which does not contain the most acidic C2–H proton such as the importance of the NTf2-anion. Therefore, as the auto-
BM2I$NTf2. Surprisingly, the [Ni(COD)2] still decomposed on decomposition of [Ni(COD)2] to Ni(0)-NPs was also observed
stirring, but this time afforded well dispersed Ni(0)-NPs (7.0 uniquely in NTf2-ILs, it is possible that the NTf2 anion inter-
2.0 nm, Fig. 4). This can only be explained by attack on the two venes in the Ni(0)-NP formation through interaction with the
less acidic protons C4–H, C5–H of the imidazolium ring and COD ligands, similarly to the case of the Ru(II) complex.
generation of transient non-classical NHC ligands.31
In another attempt to avoid auto-decomposition the
Conclusions
NTf2-anion was exchanged for the more strongly coordinating
BF4-anion, yielding important results. In this case, similarly to In summary, IL decomposition and simultaneous Ru(0)-NP
formation are limited to imidazolium salts containing the NTf2
anion. These results may explain how the Ru(0)-NPs formation is
induced by imidazolium N-triflate ILs, but cannot explain the
exact mechanism for the imidazolium ring fragmentation in small
quantities. Most importantly, the formation of small-sized active
ruthenium nanoscale hydrogenation catalysts is possible at low
temperature and atmospheric pressure in the absence of classical
and potentially dangerous reducing agents such as hydrogen gas
(under elevated pressure), pyrophoric LiAlH4 or hazardous
hydrazine. Moreover, an unexpected spontaneous decomposition
of [Ni(COD)2] occurred without the addition of hydrogen for
imidazolium ILs with short alkyl chains: ILs BMI$NTf2,
EMI$NTf2 and BM2I$NTf2. In the case of ILs BMI$NTf2 and
Scheme 4 Reactions occurring during the auto-decomposition of EMI$NTf2, TEM micrographs showed that Ni(0)-NPs of fairly
[Ni(COD)2] in imidazolium ILs. large diameter were formed as well as sponge-like super
agglomerates. However, for IL BM2I$NTf2 well dispersed Ni(0)- Then, the IL HM2I$NTf2 (1.5 mL) was added via syringe under
NPs are formed. An explanation for the activation of the acidic an argon flow. The mixture was stirred at room temperature for
protons on the imidazolium ring and the subsequent NHC 60 min, resulting in a turbid dispersion. The system was heated to
formation that led to a rapid decomposition of the complex has 50 C, and stirred under argon for two days resulting in a black
been proposed. The additional interaction between the NTf2 suspension. The Fischer–Porter bottle was connected to a mass
anion and the COD ligands on the Ni complex, which also may gas-analyser in order to evaluate the gas phase. After analysis the
support the Ni(0)-NPs formation, cannot be excluded nonethe- Fischer–Porter bottle was then kept under reduced pressure to
less. It is conceivable that this novel approach may be extended as eliminate the organic volatiles formed. An aliquot of the Ru(0)-
a general access to M-NPs in imidazolium NTf2-ILs, starting from NPs embedded in the IL was analysed by TEM.
organometallic complexes bearing only COD, allylic and/or
olefinic ligands, such as [Co(COD)2], [Ru(COD)(COT)],
Synthesis of Ni(0)-NPs
[Rh(allyl)3], [Pd(COD)2], [Pt(COD)2] or [Pt(norbornene)3].
[Ni(COD)2] (50 mg, 0.14 mmol) was stirred under argon in the
Jun 2012 | doi:10.1039/C0NR00574F
performed in a JEOL-JEM 2010 microscope operating at 200 kV A 2 mL sample of the Ni(0)-NPs solution in the IL was intro-
(UFRGS-CME, Brazil) and EDS, and used for catalytic experi- duced under argon into a Schlenk of known volume. The sample
ments as previously described.14 The HR-TEM was performed at was treated under flow of argon for 18 h. The argon atmosphere
the ‘‘Centro de Tecnologias Estrat egicas do Nordeste’’ was replaced with a known pressure of ethylene using a vacuum
1 - on
(CETENE), Recife/Brazil. The ILs were synthesised as previ- line and the system was stirred. A decrease of the internal pres-
08 October 2010
ously reported,49 and ILs EMI$B(CN)4 and BMI$N(CN)2 were sure was observed, and the composition of the gas phase was
monitored by gas chromatography. After 12 h, the atmosphere
Published onversion
10 P. S. Campbell, C. C. Santini, F. Bayard, Y. Chauvin, V. Colliere, 29 J. P. Genet, S. Mallart, C. Pinel, S. Juge and J. A. Laffitte,
A. Podgorsek, M. F. Costa Gomes and J. S a, J. Catal., 2010, 257, Tetrahedron: Asymmetry, 1991, 2, 43–46.
99–107. 30 J. D. Scholten and J. Dupont, Organometallics, 2008, 27, 4439–4442.
11 T. Gutel, C. C. Santini, K. Philippot, A. Padua, K. Pelzer, 31 S. Grundemann, A. Kovacevic, M. Albrecht, J. W. Faller and
B. Chaudret, Y. Chauvin and J. M. Basset, J. Mater. Chem., 2009, R. H. Crabtree, Chem. Commun., 2001, 2274–2275.
19, 3624–3631. 32 N. Cordente, C. Amiens, B. Chaudret, M. Respaud, F. Senocq and
12 P. Migowski, G. Machado, S. R. Texeira, M. C. M. Alves, J. Morais, M. J. Casanove, J. Appl. Phys., 2003, 94, 6358–6365.
A. Traverse and J. Dupont, Phys. Chem. Chem. Phys., 2007, 9, 4814– 33 N. Cordente, M. Respaud, F. Senocq, M. J. Casanove, C. Amiens and
4821. B. Chaudret, Nano Lett., 2001, 1, 565–568.
13 P. Migowski, D. Zanchet, G. Machado, M. A. Gelesky, S. R. Teixeira 34 P. Migowski, S. R. Teixeira, G. Machado, M. C. M. Alves, J. Geshev
and J. Dupont, Phys. Chem. Chem. Phys., 2010, 12, 6826–6833. and J. Dupont, J. Electron Spectrosc. Relat. Phenom., 2007, 156, 195–
14 M. H. G. Prechtl, M. Scariot, J. D. Scholten, G. Machado, 199.
S. R. Teixeira and J. Dupont, Inorg. Chem., 2008, 47, 8995–9001. 35 P. S. Campbell, C. C. Santini, D. Bouchu, B. Fenet, K. Philippot,
15 M. Scariot, D. O. Silva, J. D. Scholten, G. Machado, S. R. Teixeira, B. Chaudret, A. A. H. Padua and Y. Chauvin, Phys. Chem. Chem.
M. A. Novak, G. Ebeling and J. Dupont, Angew. Chem., Int. Ed., Phys., 2010, 12, 4217–4223.
2008, 47, 9075–9078. 36 L. S. Ott, S. Campbell, K. R. Seddon and R. G. Finke, Inorg. Chem.,
Jun 2012 | doi:10.1039/C0NR00574F
ChemSusChem, 2008, 1, 291–294. C. C. Santini and Y. Chauvin, J. Organomet. Chem., 2008, 693,
24 D. J. Flannigan, S. D. Hopkins and K. S. Suslick, J. Organomet. 2407–2414.
Published onversion
27 S. Chowdhury, R. S. Mohan and J. L. Scott, Tetrahedron, 2007, 63, 48 D. C. Graham, K. J. Cavell and B. F. Yates, Dalton Trans., 2007,
2363–2389. 4650–4658.
28 R. Bini, C. Chiappe, E. Marmugi and D. Pieraccini, Chem. Commun., 49 C. C. Cassol, G. Ebeling, B. Ferrera and J. Dupont, Adv. Synth.
2006, 897–899. Catal., 2006, 348, 243–248.
Appendix 3
235
tel-00708600, version 1 - 15 Jun 2012
236
Appendix 3 Separation of Hafnium and Zirconium in Ionic Liquids
The issue
Zirconium and hafnium occur naturally together, the most common ore being “zircon”. As
their chemical properties are very similar, extracting and separating highly pure metals is a
difficult challenge. In nuclear applications zirconium has a low neutron absorption cross-section
and is therefore a useful cladding material (“zircaloy”). Hafnium, on the other hand, has a high
neutron absorption cross-section (nearly 600 times that of zirconium). Therefore, for nuclear
applications, zirconium metal must contain less than 100 ppm hafnium. Three main processes are
in current use for the separation of zirconium from hafnium: multiple crystallization of potassium
zirconium fluoride; solvent extraction of the chlorides using methyl isobutyl ketone and water;
tel-00708600, version 1 - 15 Jun 2012
extractive distillation of the chlorides using a KCl/AlCl3 molten salt bath at high temperature.1-11
These three processes present several severe drawbacks, e.g. recovery of oxide products,
corrosion problems, high running temperatures, etc. Hf/Zr separation is still a economic and
technical challenge (more than 60 papers published during the last decade, of which 25 % as
patents and from more than 20 different countries).12
Strategies
As part of this work, two strategies were attempted using ionic liquids with the ultimate
goal of achieving Zr/Hf separation with reduced hazards and process costs, as ILs are considered
as “green solvents”, due to some specific features such as non-flammability, high thermal
stability and non-volatility.13
1) Firstly, a novel range of imidazolium chlorometallate ionic liquids based on hafnium
and zirconium were synthesised and fully characterised. A difference in chemical or
physical properties of the resulting materials may have been exploited in their
extraction. The full characterisation of the new materials obtained is described in the
Dalton Transactions paper attached.
2) Secondly, it is reported in the literature that ZrCl4 and HfCl4 display a different
reactivity towards methyl-substituted aromatics; ZrCl4 produces a molecular complex,14
while reaction with HfCl4 results in an ionic species.15, 16 Thus the approach was to use
this difference in reactivity to undertake a subsequent liquid-liquid extraction, based on
237
Appendix 3 Separation of Hafnium and Zirconium in Ionic Liquids
the potential high solubility of the ionic complex in an ionic liquid compared to an
organic solvent. The preliminary positive results, for which a patent application has
been submitted, are described in the Chimica Oggi paper, also included in this
appendix.
References
[1] B. Prakash, C. V. Sundaram, Met. Mater. Proc. 2009, 21, 21.
[2] V. Ogarev, A. Skotnicki, B. Ninham, ,, Patent AU. 20080527. 2009, 12pp.
[3] L. A. Niselson, E. A. Egorov, E. L. Chuvilina, O. A. Arzhatkina, V. D. Fedorov, J. Chem.
Eng., 2009, 54, 726.
[4] M. Smolik, A. Jakobik-Kolon, M. Poranski, Hydrometallurgy, 2009, 95, 350.
tel-00708600, version 1 - 15 Jun 2012
[5] K. Saberyan, A. H. Meysami, F. Rashchi, E. Zolfonoun, Chin. J. Chem., 2008, 26, 2067.
[6] M. Taghizadeh, R. Ghasemzadeh, S. N. Ashrafizadeh, K. Saberyan, M. G. Maragheh,
Hydrometallurgy, 2008, 90, 115.
[7] L. Delons, G. Picard, D. Tigreat, Compagnie Europeenne Du Zirconium Cezus, WO
20020412. 2002, 20 pp.
[8] L. Delons, S. Lagarde, A. Favre Reguillon, S. Pellet Rostaing, M. Lemaire, L. Poriel,
Compagnie Europeenne du Zirconium-Cezus, Fr. Fr 2004-7721 2006, 40.
[9] J. A. Sommers, J. G. Perrine, US, ATI Properties, Inc., USA, 20020129. 2003, 8 pp.
[10] N. Ozanne, M. L. Lemaire, A. Guy, J. Foos, S. Pellet-Rostaing, F. Chitry, WO,
Compagnie Europeenne du Zirconium Cezus, Fr., 20010910. 2002, 23 pp.
[11] A. Da Silva, E. El-Ammouri, P. A. Distin, Can. Metall. Q. 2000, 39, 37.
[12] R. H. Nielsen, in KIRK OTHMER ENCYCLOPEDIA OF CHEMICAL TECHNOLOGY
John Wiley & Sons, New York, 2004.
[13] P. Wasserscheid, T. Welton, Ionic liquids in synthesis, Wiley-VCH, Weinheim, 2008.
[14] F. Musso, E. Solari, C. Floriani, K. Schenk, Organometallics 1997, 16, 4889.
[15] F. Calderazzo, I. Ferri, G. Pampaloni, S. Troyanov, J. Organomet. Chem., 1996, 518,
189.
[16] F. Calderazzo, P. Pallavicini, G. Pampaloni, P. F. Zanazzi, J. Chem. Soc., Dalton Trans.,
1990, 2743.
238
PAPER www.rsc.org/dalton | Dalton Transactions
attempted of [C1 C4 Im]2 [HfCl6 ], and [C1 C4 Im]2 [ZrCl6 ] monitored by 35 Cl and 91 Zr solid NMR (high
temperature up to 551 K).
This journal is © The Royal Society of Chemistry 2010 Dalton Trans., 2010, 39, 1379–1388 | 1379
Scheme 1 The preparation of imidazolium chlorometallates.
1380 | Dalton Trans., 2010, 39, 1379–1388 This journal is © The Royal Society of Chemistry 2010
became clear that a reaction had occurred as the products were for Zr. In the case of Hf only the band [3(C1 C4 Im)MCl6 )]+ at m/z =
coloured solids, (pale green Hf, and red Zr). Deliberate formation 808.7 was present.
of this red Zr product by heating first to 473 K for 30 min gave DSC Negatively charged clusters due to the presence of Cl-
profiles with distinct endothermic peaks at 351 K and 374 K. One and Mx Cly - anions appear in the ESI- spectra. For R = 0.33,
hypothesis is that at room temperature there is formation of a the isotopic patterns corresponding to MCl5 - (m/z = 266.9 (Zr)
mixture of [C1 C4 Im][M2 Cl9 ] (liquid) and [C1 C4 Im]2 [MCl6 ] (solid), and 356.96 (Hf)) and the cluster [5(C1 C4 Im)3 (HfCl6 )]- , m/z =
the “kinetic products” as wet crystals, which on heating to 340 K 1867 were present. For R = 0.5 only the isotopic patterns
undergo a partial comproportionation reaction to give a eutectic corresponding to [MCl5 ]- {(m/z = 266.9 (Zr) and 356.9 (Hf)}
mixture, possibly containing M2 Cl10 2- anions already reported.35,36 were observed, Fig. 4. These data for R = 0.5 agree with the
DSC results showed firstly a variation in the melting point of structure [C1 C4 Im][MCl5 ] and not [C1 C4 Im]2 [M2 Cl10 ], as previ-
the salts produced, that was almost identical in the case of Hf ously reported.35,36
and Zr. Also interestingly, the complexes [C1 C4 Im][M2 Cl9 ] at R = For R = 0.67, as well as the isotopic pattern corresponding to
0.67 were still liquid at sub zero temperatures. The appearance [MCl5 ]- (m/z = 266.9 (Zr) and 356.9 (Hf), the isotopic pattern
of the DSC profiles themselves indicated that pure complexes corresponding to [M2 Cl9 ]- m/z = 500.3 for Zr, and m/z = 676.5
[C1 C4 Im]2 [MCl6 ], were achieved in the case of R = 0.33, as for Hf, and the cluster corresponding to [C1 C4 Im][Zr2 Cl9 ] at
these gave the sharpest most precise peak for the phase change. 1141.0 were apparent, Fig. 5. These results are consistent with
For R = 0.5, the DSC data would support the formation of findings by FTICR-MS, IR, and Raman spectra proving that
[M2 Cl10 ][C1 C4 Im]2 during the acquisition. [(C4 C4 ImCl)2(TiCl4 )] contained a large amount of [Ti2 Cl9 ]- ions
and a small amount of [TiCl6 ]2- ions.33
NMR experiments
Electrospray ionisation mass spectrometry is used to investigate
the nature of ILs in both the gas and solution phase and to The cation [C1 C4 Im]+ can be used as a probe to study the nature of
determine the nature of the ionic species present within.37-39 the ionic liquid. For this purpose, solutions of the different ionic
In the positive-ion mode mass spectra of the ionic liquids studied liquids in the same solvent (CD3 CN) and at the same concentration
at concentrations of ~10-6 M in acetonitrile, for all ionic liquids, and temperature (0.1 mol L-1 , 278 K), were analysed by 1 H, 13 C
the peak at m/z = 139.1 corresponding to the [C1 C4 Im]+ ion is NMR. Also, solid state 1 H, 13 C, NMR analyses have been run
present. The peak of the ion [2(C1 C4 Im)Cl]+ at m/z ~ 313, which on neat derivatives. The assignment of the nuclei is depicted in
decreased when R increased, was absent in each case when R = Scheme 2.
0.67. Larger clusters of [n(C1 C4 Im)(n-1)Cl]+ were not observed.
When R = 0.33 larger metal containing clusters were observed as
follows : [3(C1 C4 Im) MCl6 ]+ and m/z = 720.7 for Zr, and m/z =
808.7 for Hf, Fig. 3. At R = 0.5 clusters of [2(C1 C4 Im)MCl5 ]+ at
544.5 and 693.8 were observed for Zr and Hf, respectively. For R =
0.67 clusters of [2(C1 C4 Im)MCl5 ]+ at 544.5, [3(C1 C4 Im) MCl6 ]+ at
720.7 (major) and of [2(C1 C4 Im)M2 Cl9 )]+ at 776.1 were observed Scheme 2 Assignments of the C and H in the Imidazolium cation.
Fig. 3 (a) Full ESI+ mass spectra of products for R = 0.33 [C1 C4 Im]2 [HfCl6 ] (upper) and [C1 C4 Im]2 [ZrCl6 ] (lower), (b) zoom on the [(C1 C4 Im)3 MCl6 ]+
region showing isotopic peaks, (c) theoretical simulation of [C1 C4 Im]3 [MCl6 ]+ isotopic peaks for comparison.
This journal is © The Royal Society of Chemistry 2010 Dalton Trans., 2010, 39, 1379–1388 | 1381
tel-00708600, version 1 - 15 Jun 2012
Fig. 4 (a) Zoom ESI- mass spectra of products for R = 0.5; Hf (upper) and Zr (lower), in [MCl5 ]- region showing isotopic peaks, (b) theoretical
simulation of [MCl5 ]- isotopic peaks for comparison.
Fig. 5 (a) Full ESI-mass spectra of R = 0.67 i.e. [C1 C4 Im][Hf2 Cl9 ] (upper) and [C1 C4 Im][Zr2 Cl9 ] (lower), (b) zoom on the [M2 Cl9 ]- region showing
isotopic peaks, (c) theoretical simulation of [M2 Cl9 ]- isotopic peaks for comparison.
1 13
H solution NMR C solution NMR
In 1 H NMR spectra of [C1 C4 Im]+ , the chemical shift of C2 – Besides the chemical shift of carbon C 2 (dC 2 ) which shifts towards
H proton (d H2 ) as well as ethylenic hydrogens C4 –H (d H 4 ) and high field with increasing R, Fig. 6c, the evolution of the 13 C NMR
C5 –H (d H5 ), although less markedly, exhibit a monotonous spectra is not significant (Dd < 0.1, Experimental part).
decrease with increasing R, Fig. 6a and 6b. This evolution could be The solution NMR analyses of both 1 H and 13 C of the
due to the decreasing hydrogen bond interactions between Cn –H imidazolium moiety, show clearly a linear displacement of chem-
and the anion as R increases. ical shifts with increasing MCl4 concentration. The addition
1382 | Dalton Trans., 2010, 39, 1379–1388 This journal is © The Royal Society of Chemistry 2010
tel-00708600, version 1 - 15 Jun 2012
Fig. 6 (Upper) Variation of the chemical shift in 1 H NMR (CD3 CN, 298 K, 0.1 mol L-1 , d ppm) of (a) d H 2 as a function of R (b) d H 4 and d H 5 as a
function of R. (Lower) Variation of the chemical shift of d C 2 in (c) 13 C solution NMR (CD3 CN, 0.1 mol L-1 , d ppm). (d) 13 C Solid state NMR, as a
function of R.
Table 1 Evolution of the chemical shift d C 2 –H in 1 H solution NMR and and R unlike for dC 2 , which is shifted monotonously towards
of the melting point for the ratios 0.33, 0.5 and 0.67 high field with increasing R, Fig. 6d. NMR analysis performed
in neat IL can reveal information, on the internal structure and
TiCl4 m.p. or Zr Cl4 m.p. or Hf Cl4 m.p. or AlCl3 m.p or
T g /K (d C 2 –H T g /K (d C 2 –H T g /K (d C 2 –H T g /K (d C 2 –H ionic association of these media. It could be proposed that the
R ppm)33 ppm)a ppm)a ppm)40,41 C2 H–anion H-bond is thus a significant factor in the tuning of the
physical properties of the resulting salt, Table 2.
0.33 377.96 (n.d) 363 (8.68) 391 (8.74) 213 (6.72)
0.5 337.5 (n.d) ~ 357 (8.49) ~ 370 (8.52) (6.5)
0.67 233.65 (n.d) 228 (8.40) 228 (8.40) 177 (6.2.5)
35
Cl liquid NMR (variable temperature)
a
This work.
In the 35 Cl NMR spectra of [C1 C4 Im]2 [HfCl6 ], and [C1 C4 Im]2 -
of MCl4 clearly has an effect on the electron density of the [ZrCl6 ], at 398 K, there is only one broad peak centred at 300 ppm,
ring. Ionic interactions in the MCl4 : ImCl system led to the which persists upon cooling to 378 K.
formation of hydrogen-bonded aggregates of anions and cations, The variable temperature 35 Cl solution NMR (CD3 CN) spectra
mainly through a C2 –H–anion interaction. The nature and extent [C1 C4 Im] Zr2 Cl9 showed two peaks at 450 and 350 ppm in a ratio of
of these depend on the susceptibility of the anion population around 1 : 2 which became sharper but did not shift when the tem-
towards hydrogen bond formation with Im+ . In general, the perature varied from 278 to 333 K. Similarly for [C1 C4 Im] Hf2 Cl9 ,
stronger C2 –H–anion hydrogen-bond induced, the more d C2 – peaks at 350 and 300 ppm are observed although the smaller is a
H is shifted to lower field, and the higher the melting point, shoulder possibly due to their proximity, Fig. 7. In the literature,
Table 1. the anion [M2 Cl9 ]- in complexes [M2 Cl9 ]- [C]+ M = Zr, Hf presents
the same structure of Cl3 M(m2 -Cl3 )MCl3 , i.e. six terminal chloride
13
C solid state NMR of IL in pure form at 298 K atoms and three bridging chloride atoms.42-45 Consequently, by
analogy with the structure of published [M2 Cl9 ]- [C]+ , the peaks at
In the 13 C solid state NMR (of neat IL), when the R value is 350 ppm (Hf) and 450 ppm (Zr) were assigned to the three bridging
increased all peaks are clearly affected, Table 2, particularly d C7 , chloride atoms and the peak at 300 ppm (Hf) and 350 ppm (Zr) to
d C 8 and d C10 , which are eventually completely replaced by three the six terminal chloride atoms. The 35 Cl NMR and ESI- findings
new peaks. However, there is no clear trend between their d values are consistent regarding the structure of these anions.
This journal is © The Royal Society of Chemistry 2010 Dalton Trans., 2010, 39, 1379–1388 | 1383
Table 2 13 C Solid-state NMR of neat compounds obtained by reaction of MCl4 with [C1 C4 Im][Cl] in ratios R = nMCl4 /nMCl4 + n[C1C4IM][Cl] equal to 0, 0.33,
0.5 and 0.67
Fig. 7 Variable temperature NMR spectra; (a) 35 Cl of [C1 C4 Im][Hf2 Cl9 ], (b) 35 Cl of [C1 C4 Im][Zr2 Cl9 ].
Thermal dissociation by 35 Cl and 91 Zr solid NMR (high M’2 [ZrCl6 ] → 2M’Cl + ZrCl4 .
temperature) of [C1 C4 Im]2 HfCl6 , and [C1 C4 Im]2 ZrCl6
In the case of 1-butyl-2,3-dimethylimidazolium trichlorozincate Due to the presence of a strong hydrogen bond B–H ◊ ◊ ◊ Cl the
[C1 C1 C4 Im][ZnCl3 ] and 1-butyl-3-methylimidazolium trichloro- thermal dissociation, with BHCl, is total only in the presence of
zincate [C1 C4 Im][ZnCl3 ], the results of variable temperature 35 Cl oxygen affording ZrO2 instead of ZrCl4 which is economically
and 13 C NMR analyses indicated that the species are dissociated unavailable for the Zr/Hf separation process.
to give ImCl and ZnCl2 , at temperatures superior to 373 K and The thermogravimetric analysis (TGA) of [C1 C4 Im]2 [ZrCl6 ], and
338 K, respectively.46 The reverse process was observed on cooling. [C1 C4 Im]2 [HfCl6 ], indicated also that the weight losses began at
This dissociation could be used for the recovery and/or separation 528 K and 538 K, respectively, instead of 493 K for [C1 C4 Im]Cl.
of a metal chloride from a mixture. Such dissociation in the This shows firstly, the presence of the chlorometallate increases
case of zirconium and/or hafnium chloride would be of crucial significantly the thermal stability, and secondly no significant
interest for their separation which is still an economic challenge.34 difference is apparent between the Zr and Hf compounds.
Furthermore, in the literature MCl4 compounds are described as Due to these findings, 35 Cl and 91 Zr solid NMR spectra could
typical Lewis acids that readily react with chloride ions to produce be run at high temperatures ranging from 513 K to 551 K,
the hexachlorometallate ion, either with alkali metals: Fig. 8. The spectra were recorded both with increasing and
decreasing temperature. For 35 Cl, the peak width decreases with
2M’Cl + ZrCl4 → M’2 [ZrCl6 ]47,48 temperature corresponding to a decrease in viscosity. A small
displacement combined with a decrease of the peak width is
or with bases such as pyridine or quinoline (B) in the presence of
observed at low temperature. This may be due to an increase of
HCl: 35
Cl mobility towards the C1 C4 Im cation. No such phenomenon
2B + 2HCl + ZrCl4 → (BH)2 [ZrCl6 ].49 is observed in the 91 Zr spectrum, whose peak remains sharp at all
the temperatures. Consequently, no dissociation was observed and
All these compounds undergo thermal dissociation at 622 K for so no decomposition occurred. This result for [C1 C4 Im]2 [ZrCl6 ],
KCl and above 522 K for (BHCl): could be explained by the presence of a close interaction between
1384 | Dalton Trans., 2010, 39, 1379–1388 This journal is © The Royal Society of Chemistry 2010
liquids whose catalytic properties are currently under study. The
high temperature 35 Cl and 91 Zr solid state NMR experiments show
that there is no dissociation of [C1 C4 Im]2 MCl6 in [C1 C4 Im][Cl] and
MCl4 up to 551 K. This result could be explained by the presence
of a close interaction between the acidic ring protons and the
chlorine of the anion.
Experimental
All operations were performed in the strict absence of oxygen and
water under a purified argon atmosphere using glovebox (Jacomex
or MBraun) or vacuum-line techniques. Under argon, solvents
were distilled using the appropriate drying agents: pyridine
from CaH2 , acetonitrile and acetonitrile-d 3 (99% de deuterium)
from P2 O5 and stored over molecular sieves. 1-chlorobutane and
1,2-dimethylimidazole were obtained from Aldrich, and were
freshly distilled before use. 1-Butyl-3-methylimidazolium chloride
[C1 C4 Im][Cl] was synthesised as described in the literature.1
1
H and 13 C and 35 Cl liquid NMR
tel-00708600, version 1 - 15 Jun 2012
1
H and 13 C solid NMR
This journal is © The Royal Society of Chemistry 2010 Dalton Trans., 2010, 39, 1379–1388 | 1385
contained in vacuum-sealed PyrexTM cells (7 mm diameter and necessary with use of a heat gun, in order to obtain a homogenous
15 mm long). In order to avoid compound decomposition during solution. The products are left stirring for 2 h and then cooled
sealing under reduced pressure of argon, the ampoule containing down to room temperature. A solid is obtained for R = 0.2, 0.33,
the compound was cooled down to liquid nitrogen temperature a semi-solid for R = 0.5, and liquids obtained for R = 0.1, 0.7.
just before sealing. The side walls and bottom of the ampoules were Example of the synthesis of 1-butyl-3-methylimidazolium zirco-
grounded in order to make these cells fit snugly into the heat flow nium pentachloride: {[C1 C4 Im][ZrCl5 ], with R = 0.5}: Anhydrous
detector. This consists of two thermopiles, connected in electrical zirconium tetrachloride (0.667 g, 2.86 mmol) was mixed under
opposition, accommodating both the previous experimental cell argon with solid 1-butyl-3-methylimidazolium chloride (0.5 g,
with the sample and another empty and identical cell, which acts as 2.86 mmol). The mixture was heated to 70 ◦ C and stirred over
a reference. This experimental DSC set up is characterised both by 2 h until a homogenous mixture formed.
1
a high sensitivity and an excellent integration of the thermal heat H NMR (CD3 CN, d) : 8.49 (1H, s, C2 H), 7.38 and 7.34 (2H,
flow since the sensing thermocouples in each thermopile are evenly d, C4,5 H), 4.20 (2H, t, C7 H 2 ), 3.83 (3H, s, C6 H 3 ), 1.80 (2H,
distributed all over the cell surface. Experiments were conducted qn, C8 H 2 ), 1.31 (2H, sx, C9 H 2 ), 0.95 (3H, t, C10 H 3 ).13 C NMR
at heating and cooling rates ranging up to 5 K min-1 . (CD3 CN, d) : 136.83 (s, C2 ), 124.59 and 123.21 (d, C4,5 ), 50.30
As the differential scanning calorimeter works in a wide temper- (s, C7 ), 36.98 (s, C6 ), 32.56 (s, C8 ), 19.91 (s, C9 ), 13.65 (s, C10 ).13 C
ature range, knowledge of the calorimeter constant dependence Solid-state NMR (d) : 135.82 (C2 ), 124.46 and 123.15 (C4,5 ), 49.68
on temperature is crucial in measurements with this apparatus. (C7 ), 39.11 and 37.52 (C6 ), 32.50 (C8 ), 20.08 and 19.02 (C9 ),
Such a dependence for Setaram DSC 121 was determined during 13.47 and 10.83 (C10 ). ESI+ : 139.1 [C1 C4 Im+ ], 312.8 [2(C1 C4 Im+ )
calorimeter calibration by the “ Joule effect”. This calibration Cl- ], 544.5 [2(C1 C4 Im+ )ZrCl5 - ], 721.0 [3(C1 C4 Im+ ) ZrCl6 2- ], 1302.8
was carried out at defined temperatures by the so-called “step [5(C1 C4 Im+ ) 2ZrCl6 2- ] ESI- : 266.9 [ZrCl5 - ]
method” (DT = 5 K) over the entire temperature range of the
tel-00708600, version 1 - 15 Jun 2012
calorimeter work. As a result the calibration curve, i.e. calorimeter [C1 C4 Im+ ][Cl- ] (R = 0). 1 H NMR (CD3 CN, d) : 9.70 (1H, s,
constant dependence on temperature, K(mV mW-1 ) = f (T), was C2 H), 7.49 and 7.46 (2H, d, C4,5 H), 4.20 (2H, t, C7 H 2 ), 3.88 (3H, s,
obtained. This dependence was automatically used during data C6 H 3 ), 1.81 (2H, qn, C8 H 2 ), 1.30 (2H, sx, C9 H 2 ), 0.91 (3H, t,
treatment by the original Setaram software. The measurement C10 H 3 ). 13 C NMR (CD3 CN, d) : 138.35 (s, C2 ), 124.38 and 123.13
of sample temperature was performed through a platinum probe (d, C4,5 ), 49.80 (s, C7 ), 36.56 (s, C6 ), 32.58 (s, C8 ), 19.83 (s, C9 ), 13.63
located in the calorimetric block. The fusion of standard materials (s, C10 ). 13 C Solid-state NMR (d) : 137.89 (C2 ), 125.23 and 124.28
was studied at various scanning rates and temperature correction (C4,5 ), 48.34 (C7 ), 38.57 (C6 ), 35.70 (C8 ), 20.17 (C9 ), 12.96 (C10 ).
coefficients were determined and introduced into the calorimeter M = Zr R = 0.1 ; [C1 C4 Im][Cl] (0.5 g, 2.86 mmol); ZrCl4
software. The maximum relative error of enthalpy of phase (7.40 ¥ 10-2 g, 3.17 ¥ 10-4 mol) 1 H NMR (CD3 CN, d) : 9.52
transition determination did not exceed 1%. It was checked by (1H, s, C2 H), 7.47 and 7.43 (2H, d, C4,5 H), 4.19 (2H, t, C7 H 2 ),
measurements of temperature and enthalpy of phase transitions 3.88 (3H, s, C6 H 3 ), 1.81 (2H, qn, C8 H 2 ), 1.30 (2H, sx, C9 H 2 ), 0.91
of standard substances. Results obtained (differences in fusion (3H, t, C10 H 3 ). 13 C NMR (CD3 CN, d) : 138.16 (s, C2 ), 124.41
temperatures less than 1 K, differences in enthalpies of fusion less and 123.10 (d, C4,5 ), 49.98 (s, C7 ), 36.67 (s, C6 ), 32.60 (s, C8 ),
than 0.5%) confirmed the correct working of the calorimeter. 19.88 (s, C9 ), 13.63 (s, C10 ). 13 C Solid-state NMR (d) : 138.06 (C2 ),
125.45 and 124.21 (C4,5 ), 48.49 (C7 ), 38.60 (C6 ), 35.88 (C8 ), 20.04
Mass spectrometry (C9 ), 13.36 (C10 ). ESI+ : 139.1 (C1 C4 Im+ ), 312.8 ([2C1 C4 Im,35 Cl]+ )
ESI- : 384.8 ([2C1 C4 Im+ ,335 Cl- ]), 556.8 ([3C1 C4 Im+ ,435 Cl- ]), 734.8
Mass spectra were acquired on a ThermoFinnigan LCQ Advan- ([4C1 C4 Im+ ,535 Cl- ])
tage ion trap instrument, detecting positive (+) and negative (-) R = 0.2 ; [C1 C4 Im][Cl] (0.5 g, 2.86 mmol); ZrCl4 (0.167 g, 7.15 ¥
ions in the ESI mode. Samples (1 to 10 mg ml-1 in acetonitrile) were 10-4 mol) 1 H NMR (CD3 CN, d) : 9.15 (1H, s, C2 H), 7.42 and 7.39
infused directly into the source (5 ml min-1 ) using a syringe pump. (2H, d, C4,5 H), 4.20 (2H, t, C7 H 2 ), 3.87 (3H, s, C6 H 3 ), 1.81 (2H,
The following source parameters were applied: spray voltage qn, C8 H 2 ), 1.30 (2H, sx, C9 H 2 ), 0.92 (3H, t, C10 H 3 ). 13 C NMR
3.0–3.5 kV, nitrogen sheath gas flow 5–20 arbitrary units. The (CD3 CN, d) : 137.81 (s, C2 ), 124.42 and 123.09 (d, C4,5 ), 49.99
heated capillary was held at 473 K. MS spectra were obtained by (s, C7 ), 36.74 (s, C6 ), 32.59 (s, C8 ), 19.85 (s, C9 ), 13.62 (s, C10 ).
13
applying a relative collision energy of 25 to 40% of the instrumental C Solid-state NMR (d) : 137.97 (C2 ), 125.16 and 124.18 (C4,5 ),
maximum. 50.75 and 48.49 (C7 ), 38.59 (C6 ), 35.68 (C8 ), 20.35 (C9 ), 13.95
(C10 ). ESI+ : 139.1 (C1 C4 Im+ ), 312.8 ([2C1 C4 Im,35 Cl]+ ) ESI- : 268.8
Preparation of ionic liquid (ZrCl5 - ), 384.8 ([2C1 C4 Im+ ,335 Cl- ]), 556.8 ([3C1 C4 Im+ ,435 Cl- ]),
734.8 ([4C1 C4 Im+ ,535 Cl- ]).
Under argon in glovebox, the solid reactants are weighed and R = 0.33 ; [C1 C4 Im][Cl] (0.5 g, 2.86 mmol); ZrCl4 (0.333 g, 1.43 ¥
mixed together thoroughly using a mortar and pestle. For each 10-3 mol) 1 H NMR (CD3 CN, d) : 8.68 (1H, s, C2 H), 7.38 and 7.36
experiment, 0.5 ± 0.005 g (2.86 ¥ 10-3 mol) of the ionic liquid (2H, d, C4,5 H), 4.20 (2H, t, C7 H 2 ), 3.86 (3H, s, C6 H 3 ), 1.81 (2H,
C1 C4 ImCl (relative molecular mass 174.7 g mol-1 , white solid at qn, C8 H 2 ), 1.32 (2H, sx, C9 H 2 ), 0.94 (3H, t, C10 H 3 ). 13 C NMR
room temperature) is weighed into each tube and the according (CD3 CN, d) : 137.20 (s, C2 ), 124.54 and 123.13 (d, C4,5 ), 50.24
amounts of MCl4 are added. (s, C7 ), 36.97 (s, C6 ), 32.66 (s, C8 ), 19.92 (s, C9 ), 13.65 (s, C10 ). 13 C
This is then heated to 343 K, (m.p. C1 C4 ImCl = 339 K), under Solid-state NMR (d) : 136.31 (C2 ), 124.19 and 122.71 (C4,5 ), 49.97
constant stirring. This produces a homogenous solution except (C7 ), 38.10 and 36.86 (C6 ), 31.91 (C8 ), 19.28 (C9 ), 15.09, 13.36 and
for in the cases of R = 0.33 and R = 0.5, where extra heating is 11.63 (C10 ). ESI+ : 139.1 [C1 C4 Im+ ], 313.0 [2(C1 C4 Im+ ) Cl- ], 720.7
1386 | Dalton Trans., 2010, 39, 1379–1388 This journal is © The Royal Society of Chemistry 2010
[3(C1 C4 Im+ ) ZrCl6 2- ], 1302.8 [5(C1 C4 Im+ ) 2ZrCl6 2- ] ESI- : 266.9 13
C Solid-state NMR (d) : 134.52 (C2 ), 124.61 and 123.37 (C4,5 ),
[ZrCl5 - ] 51.03 (C7 ), 37.16 (C6 ), 32.20 (C8 ), 19.57 (C9 ), 13.87 (C10 ). ESI+ :
R = 0.67; [C1 C4 Im][Cl] (0.5 g, 2.86 mmol); ZrCl4 (1.56 g, 139.1 [C1 C4 Im+ ], 808.7 [3(C1 C4 Im+ ) HfCl6 2- ] ESI- : 356.9 [HfCl5 - ],
6.67 mmol) 1 H NMR (CD3 CN, d) : 8.40 (1H, s, C2 H), 7.35 and 676.5 [Hf2 Cl9 - ]
7.31 (2H, d, C4,5 H), 4.12 (2H, t, C7 H 2 ), 3.80 (3H, s, C6 H 3 ), 1.78
(2H, qn, C8 H 2 ), 1.27 (2H, sx, C9 H 2 ), 0.88 (3H, t, C10 H 3 ). 13 C NMR
(CD3 CN, d) : 136.61 (s, C2 ), 124.57 and 123.20 (d, C4,5 ), 50.31 (s,
Acknowledgements
C7 ), 37.01 (s, C6 ), 32.49 (s, C8 ), 19.90 (s, C9 ), 13.69(s, C10 ). 13 C We would like to thank Areva Cezus for financing this work, Anne
Solid-state NMR (d) : 134.76 (C2 ), 124.84 and 123.39 (C4,5 ), 51.05 Baudouin for solid state NMR and also undergraduate students
(C7 ), 37.50 (C6 ), 32.18 (C8 ), 19.58 (C9 ), 14.03 (C10 ). ESI+ : 139.1 Céline Chong, Thibault Alphazan and Sandra Garcia for their
[C1 C4 Im+ ], 544.5 [2(C1 C4 Im+ )ZrCl4 - ], 720.7 [3(C1 C4 Im+ )ZrCl6 2- ], participation.
776.1 [2(C1 C4 Im+ )Zr2 Cl9 - ]ESI- : 266.9 [ZrCl5 - ], 500.3 [Zr2 Cl9 - ],
1141.0 [C1 C4 Im+ 2Zr2 Cl9 - ]
M = Hf R = 0.1; [C1 C4 Im][Cl] (0.5 g, 2.86 mmol); HfCl4 (0.102 g, Notes and references
3.17 ¥ 10-4 mol) 1 H NMR (CD3 CN, d): 9.44 (1H, s, C2 H), 7.46 1 P. Wasserscheid and T. Welton, Ionic Liquids in Synthesis, Wiley-VCH,
and 7.43 (2H, d, C4,5 H), 4.19 (2H, t, C7 H 2 ), 3.87 (3H, s, C6 H 3 ), Weinheim, 2003.
1.78 (2H, qn, C8 H 2 ), 1.32 (2H, sx, C9 H 2 ), 0.91 (3H, t, C10 H 3 ). 13 C 2 A. P. Abbott, G. Capper, D. L. Davies and R. K. Rasheed, Chem.–
Eur. J., 2004, 10, 3769–3774.
NMR (CD3 CN, d) : 138.10 (s, C2 ), 124.37 and 123.07 (d, C4,5 ), 3 R. T. Carlin and R. A. Osteryoung, J. Mol. Catal., 1990, 63, 125–
49.89 (s, C7 ), 36.64 (s, C6 ), 32.57 (s, C8 ), 19.86 (s, C9 ), 13.60 (s, C10 ). 129.
13
C Solid-state NMR (d) : 137.82 (C2 ), 125.44 and 124.21 (C4,5 ), 4 Y. Chauvin, S. Einloft and B. H. Olivier, Ind. Eng. Chem. Res., 1995,
34, 1149–1155.
48.50 (C7 ), 38.84 (C6 ), 35.88 (C8 ), 20.28 (C9 ), 13.11 (C10 ). ESI+ :
tel-00708600, version 1 - 15 Jun 2012
This journal is © The Royal Society of Chemistry 2010 Dalton Trans., 2010, 39, 1379–1388 | 1387
32 P. B. Hitchcock, R. J. Lewis and T. Welton, Polyhedron, 1993, 12, 43 F. Musso, E. Solari, C. Floriani and K. Schenk, Organometallics, 1997,
2039. 16, 4889–4895.
33 L. Gao, L. Wang, T. Qi, J. Chu and J. Qua, J. Electrochem. Soc., 2009, 44 M. Schormann, S. Garratt, D. L. Hughes, J. C. Green and M.
156, 49–55. Bochmann, J. Am. Chem. Soc., 2002, 124, 11266–11267.
34 R. H. Nielsen, in Kirk Othmer Encyclopedia of Chemical Technology, 45 M. Schormann, S. Garratt and M. Bochmann, Organometallics, 2005,
Wiley, New York, 2004. 24, 1718–1724.
35 B. Neumueller and K. Dehnicke, Z. Anorg. Allg. Chem., 2004, 630, 46 C. C. Cassol, G. Ebeling, B. Ferrera and J. Dupont, Adv. Synth. Catal.,
2576–2578. 2006, 348, 243–248.
36 E. Robe, S. Maria, P. Richard and R. Poli, Eur. J. Inorg. Chem., 2007, 47 J. E. Drake and G. W. A. Fowles, J. Inorg. Nucl. Chem., 1961, 18, 136.
2434–2442. 48 P. Gelbman and A. D. Westland, J. Chem. Soc., Dalton Trans., 1975,
37 F. C. Gozzo, L. S. Santos, R. Augusti, C. S. Consorti, J. Dupont and 1598.
M. N. Eberlin, Chem.–Eur. J., 2004, 10, 6187–6193. 49 H. V. Thanh, L. Gruzdiewa, J. Rak and J. Blazejowski, Thermochim.
38 P. J. Dyson, I. Khalaila, S. Luettgen, J. S. McIndoe and D. Zhao, Chem. Acta, 1993, 230, 269–292.
Commun., 2004, (19), 2204–2205. 50 V. Lacassagne, C. Bessada, B. Ollivier, D. Massiot and J.-P. Coutures,
39 D. F. Kennedy and C. J. Drummond, J. Phys. Chem. B, 2009, 113, C. R. Acad. Sci. IIb, 1997, 325, 91–98.
5690–5693. 51 D. Massiot, C. Bessada, P. Echegut, J. P. Coutures and F. Taullele, Solid
40 C. Scordilis-Kelley, K. D. Robinson, K. A. Belmore, J. L. Atwood and State Ionics, 1990, 37, 223.
R. T. Carlin, J. Crystallogr. Spectrosc. Res., 1993, 23, 601–606. 52 (a) M. Gaune-Escard, Calorimetric methods, in Molten Salt Tech-
41 D. Bankmann and R. Giernoth, Prog. Nucl. Magn. Reson. Spectrosc., niques, Eds. R.J. Gale and D. G. Lovering, Vol. 4, Plenum Press, New
2007, 51, 63–90. York, pp. 151-192 (1991); (b) E. Ingier-Stocka, L. Rycerz, M. Berkani
42 F. Calderazzo, I. Ferri, G. Pampaloni and S. Troyanov, J. Organomet. and M. Gaune-Escard, J. Mol. Liq., 2009, 148, 40–44 and references
Chem., 1996, 518, 189–196. therein.
tel-00708600, version 1 - 15 Jun 2012
1388 | Dalton Trans., 2010, 39, 1379–1388 This journal is © The Royal Society of Chemistry 2010
34536 73893:;
Z?.*CA 2J
(%"(-',!%"
CAI#"B$ DC#I E@$#C?GA $?F*#C/ _%(`J P#^A"ACN 0@A @?.@ I?/E?X?$?0+ O@A/A >CA$?I?'BC+ CA/*$0/ GAI#'/0CB0A 0@B0 ^@A' CABE0AG ^?0@
#D 0#$*A'A ?' ?#'?E $?F*?G/ _^0J m (1 >ACEA'0` _%U!%f` EB' XA IA0@+$ /*X/0?0*0AG BCA'A/N G?DDACA'0 CABE0?"?0+ >CA/A'0AG X+
/@BC>$+ GAECAB/AG ?' 0@A >CA/A'EA #D B$?>@B0?E @+GC#EBCX#' [C6$ ( B'G PD6$ ( IB+ XA A=>$#?0AG ?' 0@A?C /A>BCB0?#'J O@?/
_%&N %Y`J 6#'/AF*A'0$+N D#C 0@A A=0CBE0?#'/ B I?=0*CA #D 0#$*A'A >C#EA// IB+ XA D*C0@AC #>0?I?HAG X+ 0A/0?'. "BC?#*/ ?#'?E
tel-00708600, version 1 - 15 Jun 2012
B'G >A'0B'A ^B/ E@#/A'J O@?/ /0*G+ @B/ #>A'AG 0@A ^B+ 0# B $?F*?G/ _B$0AC?'. 0@A EB0?#'`N 0@A /#$"A'0/ ?' 0@A #C.B'?E >@B/A
>#0A'0?B$ 'A^ >C#EA// D#C 0@A /A>BCB0?#' #D [C B'G PDN *'GAC #C 0@A $?.B'G B'G 0@A GA"A$#>IA'0 #D B I*$0?/0B.A >C#EA// D#C
I?$G E#'G?0?#'/ */?'. B' ?#'?E $?F*?G,#C.B'?E /#$"A'0 /+/0AIJ 0@A /A>BCB0?#' B'G CAE#"AC+ #D >*CA [C6$( IB+ 0@ACAD#CA XA
A'"?/B.AGJ
$+,'-#,)."&)&!,(',,!%" +/0123415678)91:63;5
-$$ #>ACB0?#'/ ^ACA EBCC?AG #*0 *'GAC B' B0I#/>@ACA #D
O@A E#I>$A=A/ [C_ V!LAV6V`6$%_g!6$`W[C6$W B'G hPD_ V!LAV6V` >*C?D?AG BC.#' ?' B .$#"A X#= #C */?'. /0B'GBCG ;E@$A'c
6$WijhPD%6$Yi! ^ACA /+'0@A/?HAG B/ CA>#C0AG ?' 0@A $?0ACB0*CA _%1! 0AE@'?F*A/J ;#$"A'0/ ^ACA GC?AG B'G G?/0?$$AG X+ E#'"A'0?#'B$
%%` _Z?.*CA 2`J IA0@#G/J) ;+'0@A/?/ #D 0@A ?#'?E $?F*?G 6 2 6 ( 3I-$6$ ( N PD_ V!
O@A?C 2W6 /#$?G /0B0A 5LQ />AE0CB ?'G?EB0A 0@B0 0@A E@AI?EB$ LA V6 V`6$ Wi jhPD %6$ Yi !N B'G [C_ V!LA V6 V`6$ %_g!6$` W[C6$ W ^ACA
/@?D0/ #D @A=BIA0@+$XA'HA'A $?.B'G BCA @?.@AC 0@B' ?' 0@A DCAA /+'0@A/?HAG B/ CA>#C0AG ?' 0@A $?0ACB0*CA _2(N %1!%%`J
BC#IB0?E >C#"?'. 0@B0 0@A BCA'A ?/ E##CG?'B0AG 0# 0@A IA0B$/J <=27>) ?8@;219:15:1A) ;BI>$A/ ^ACA B'B$+HAG #' ;@?IBGH* <:n
O@A GA/@?A$G?'. ?/ @?.@AC ?' 0@A EB/A #D hPD_ V !LA V 6 V ` &11P; />AE0C#IA0ACJ K$B/0?E D?$I/ */AG D#C 0@A B'B$+/A/ BCA
6$WijhPD%6$Yi! 0@B' D#C [C_ V!LAV6V`6$%_g!6$`W[C6$Wb LA a !"_jUJ% >#$+>C#>+$A'A D#C 0@A ?#'?E $?F*?G B'G >#$+A0@+$A'A! 0ACA>@0B$B0A
>>I` "AC/*/ _j1Nf>>I` B'G 6-C a !"_j2V >>I` "AC/*/ _m 1>>I`J D#C 0@A #C.B'?E >@B/A/J
L#CA#"ACN ?' 0@A 'A.B0?"A <;3 IB// />AE0CB 0@A @?.@A/0 ,;8@63;5) "B$) 901:627 CD) 75E) CF(A ^ACA CAE#CGAG #' MC*cAC
?'0A'/?"A IB//?D D#C [C ?/ B0 VU&JU E#CCA/>#'G?'. 0# h[C %_ V! -"B'EA W11 LPH />AE0C#IA0AC D#C 2P B'G 2W6J 6@AI?EB$ /@?D0/
LAV6V`6$Yi! B'G D#C PD 0@A IB//?D B0 VfVJ( B//?.'AG 0# hPD%6$Yi!J ^ACA IAB/*CAG CA$B0?"A 0# ;?LA( B/ ?'0AC'B$ /0B'GBCG D#C 2P
-$$ 0@A /#$*X?$?0?A/ BCA A"B$*B0AG X+ n!CB+ D$*#CA/EA'EAJ O@A B'G 2W6J
IA0@#G ?/ CB>?GN /A'/?0?"AN BEE*CB0A B'G B$I#/0 *'?"AC/B$$+ 2W6 ;#$?G /0B0A 5LQ />AE0CB ^ACA CAE#CGAG #' B MC*cAC :;n!
B>>$?EBX$AJ l@A' B /BI>$A 0@B0 E#'0B?'/ A$AIA'0 - ?/ W11 />AE0C#IA0AC AF*?>>AG ^?0@ B /0B'GBCG ( II G#*X$A!
?CCBG?B0AG X+ 0@A ?'E?GA'0 n!CB+ /#*CEAN D$*#CA/EA'0 n!CB+ XABC?'. >C#XA @ABG B'G #>ACB0?'. B0 fUJ(fN B'G W11J2& LPH
CBG?B0?#' #D A$AIA'0 - ?/ .A'ACB0AGN hPD _7#` 1JU(fV B'G [C _d#` D#C 2W6N B'G 2PN CA/>AE0?"A$+J 2W6 5LQ />AE0CB ^ACA CAE#CGAG
1JWfW2i ^@ACA 0@A ?'0A'/?0+ #D 0@?/ D$*#CA/EA'EA ?/ G?CAE0$+ B/ D#$$#^/a W1q >*$/A #' 2W6 '*E$A?N B'G BEF*?/?0?#' *'GAC B'
>C#>#C0?#'B$ 0# 0@A BI#*'0 #D A$AIA'0 - ?' 0@A /BI>$AJ OBc?'. &1 dPH GAE#*>$?'. D?A$G #' >C#0#'/N ^?0@ 2/ CAE+E$A GA$B+ B'G
0@?/ ?'0# BEE#*'0N ?D 0@A D$*#CA/EA'0 n!CB+ ?'0A'/?0+ B'G U dPH />?''?'. />AAG *'$A// #0@AC^?/A />AE?D?AGJ
E#'EA'0CB0?#' #D B' A$AIA'0 E#'0B?'AG ?' B /BI>$A ?/ c'#^' B799) ,01:62;4162>A) ^B/ CAB$?HAG #' B O@ACI#Z?''?.B' 768
h[C4% B'G PD4%i 0@A' ^A EB' .# ?' CA"AC/A B'G D?'G @#^ I*E@ -G"B'0B.A ?#' 0CB> ?'/0C*IA'0N GA0AE0?'. >#/?0?"A _j` B'
A$AIA'0 - E#'0B?'AG ?' B'#0@AC /BI>$A X+ ?0/ D$*#CA/EA'0 'A.B0?"A _!` ?#'/ ?' 0@A <;3 I#GAJ ;BI>$A/ _2 0# 21 g. I$!2 ?'
n!CB+ ?'0A'/?0+JhW1i ;*?0BX$A >C#EAG*CA/ @B"A XAA' GA"A$#>AG BEA0#'?0C?$A` ^ACA ?'D*/AG G?CAE0$+ ?'0# 0@A /#*CEA _U g$ I?'!2`
D#C 0@A GA0ACI?'B0?#' #D I?EC#.CBI BI#*'0/J */?'. B /+C?'.A >*I>J O@A D#$$#^?'. /#*CEA >BCBIA0AC/ ^ACA
;#$*X?$?0+ ?/ B$/# 0A/0AG ?' 626(3I-$6$(J O@A E#I>$A= ?/ BGGAG 0# B>>$?AGa />CB+ "#$0B.A WJ1!WJU cRN '?0C#.A' /@AB0@ .B/ D$#^ !U%1
?#'?E $?F*?G B'G /0?CCAG #"AC'?.@0J O@A D#$$#^?'. F*B'0?0?A/ #D BCX?0CBC+ *'?0/J O@A @AB0AG EB>?$$BC+ ^B/ @A$G B0 (fW dJ L;
E#I>$A= BCA G?//#$"AGa U1 I. _&JUgI#$` #D [C_ V!LAV6V`6$%_g!
6$`W[C6$W ?' 626(3I-$6$( _% I$` B'G (f I. _(JY gI#$` #D #D
hPD_ V!LAV6V`6$WijhPD%6$Yi! ?' 626(3I-$6$( _2 I$ `J O@A n!CB+
?'0A'/?0+ #D X#0@ [C B'G PD ?' 0@A/A /#$*0?#'/ ^B/ IAB/*CAG
_OBX$A2`J
O@A'N 0#$*A'A ^B/ BGGAG _W I$ D#C [C /#$*0?#'b 2JU I$ D#C PD
/#$*0?#'ib 0@A I?=0*CA/ ^ACA /0?CCAG G*C?'. #'A GB+ 0# A=0CBE0 0@A
E#I>$A=J -D0AC 0@A GAEB'0B0?#'N #'$+ #'A >@B/A ^B/ #X/AC"AGJ OBX$A 2J 3'0A'/?0?A/ #D d# CB+ D#C [C B'G 7# CB+ D#C PD ?' ?#'?E $?F*?G B'G
KA'0B'A ^B/ */AG 0# A=0CBE0 0@A 0#$*A'A DC#I 0@A ?#'?E $?F*?G _% #C.B'?E >@B/A B'G A=0CBE0?#' ADD?E?A'E+ D#C [C B'G PDJ
/>AE0CB ^ACA #X0B?'AG X+ B>>$+?'. B CA$B0?"A E#$$?/?#' A'AC.+ YJ )J-J ;#IIAC/N )JkJ KACC?'AN 8ME%7/N%6-&A*-,"*+E%N1 5E%8M7E%9::9:I9O5N
#D %U 0# (1 >ACEA'0 #D 0@A ?'/0C*IA'0B$ IB=?I*IJ >J & _%11W`J
,;8@H3836>) 61969) 75E) 1/627:63;59A) -$$ /#$*X?$?0?A/ ^ACA A"B$*B0AG 21J 5J 4HB''AN LJ7J 7AIB?CA A0 B$JN FGE% )&#A$'1"*% ?2-&A**11*% 02%
C"- &1"2#%)*D2+E%K-5E%9::I:OI:N%MFN _%11%`J
X+ n!CB+ D$*#CA/EA'EAJ 3'0A'/?0+a PD _7B` 1JU(fV B'G [C _dB`
22J -J :B ;?$"BN <J <$!-II#*C? A0 B$JN )$15%4*,$335%P5N F*N >J Wf _%111`J
1JWfW2J 6#I>$A=A/ ^ACA BGGAG 0# 0@A ?#'?E $?F*?G B'G /0?CCAG
2%J QJPJ 5?A$/A'N ?' Q"-R%&,!#*-%*1 . 3&A*0"$%&S% !*#" $3%,* !1&3&'.%
#"AC'?.@0J O@A' 0#$*A'A ^B/ BGGAG B'G /0?CCAG G*C?'. #'A
>&!1%F"3*.%T%M&1+EN 5A^ S#Cc _%11(`J
GB+J 4'$+ #'A >@B/A ^B/ #X/AC"AGJ KA'0B'A ^B/ 0@A' BGGAG 2WJ 7J P*?I?'N :J ;@A'.N N&1" % U"V2"0+% "1% )!*#" $3% 71$3.+"+% _<GJa LJ
0# A=0CBE0 0@A 0#$*A'A DC#I 0@A ?#'?E $?F*?G _OBX$A 2`J d#A$`N 6Q6 KCA/// M#EB QB0#'N >J %VY _%11Y`J
2(J KJ lB//AC/E@A?GN OJ lA$0#'N N&1" % 3"V2"0+% "1% +.1,!*+"+N l?$A+!R6PN
lA?'@A?I _%11&`J
.(I"%J-+&KB+"#, 2UJ LJ7J :?A0HN M*A5%M "5%/* !&35N PCN >J %1(f _%11V`J
2VJ RJ-J 6#EB$?BN -J<J R?//AC A0 B$JN N&1" % 3"V2"0+% "1% M.1,!*+"+E% CN %G AGJ
lA 0@B'c *'GAC.CBG*B0A /0*GA'0/ 7*E?A 6B*IA00A B'G _<GJa KJ lB//AC/E@A?GN lA$0#'N OJN <G/Jb`N l?$A+ !R6PN lA?'@A?I
_%11&`J
ZBX?A' Z#?//BE D#C 0@A?C A=>AC?IA'0B$ E#'0C?X*0?#'J -''A
2fJ :J)J 6#$A!PBI?$0#'N QJKJ O##HAN )$,$3.+,% +*A$-$,"&1E% -* &W*-.% $10%
MB*G#*?'N KCJ :A'?/ M#*E@* B'G KCJ Q#XAC0 6@BCA+C#' D#C 0@A?C
-* . 3"1'X% )!*#"+,-.% $10% 6-& *++% B*+"'1N ;>C?'.ACN :#CGCAE@0
@A$> ?' 5LQN LB// ;>AE0C#IA0C+ B'G n!CB+ D$*#CA/EA'EA
_%11V`J
A=>AC?IA'0/ B'G :CJ Q#'.B _6AH*/ ;#E?A0+` D#C 0@A .?D0 #D >*CA 2&J OJ k*0A$N KJ 6BI>XA$$ A0 B$JN )!"#" $% G''"()!*#"+,-.% /&0$.N MQN
PD6$( /BI>$AJ >J (&%_%11Y`J
2YJ RJ-J 6#EB$?BN dJ<J k*0#^/c? A0 B$JN )&&-05%)!*#5%Y*W5N MNGN >J fUU
_%11V`J
$+L+$+"(+,)."&)"%#+, %1J ZJ L*//#N <J ;#$BC? A0 B$JN G-'$1&#*,$33" +N CON >J (&&Y _2YYf`J
%2J ZJ 6B$GACBHH#N 3J ZACC? A0 B$JN >5%G-'$1&#*,5%)!*#5N NC N >J 2&Y _2YYV`J
2J MJ KCBcB/@N 6JRJ ;*'GBCBIN 4*,5%4$,*-5%6-& 5N MCN >J %2%_%11Y`J %%J ZJ 6B$GACBHH#N KJ KB$$B"?E?'? A0 B$JN >5% )!*#5% M& 5E% B$3,&1% /-$1+5N
%J RJ 4.BCA"N -J ;c#0'?Ec? A0 B$JN 6$,*1,%785%9::;:<9=5%>J 2% _%11Y`J >J %f(W _2YY1`J
WJ 7J-J 5?/A$/#'N <J-J <.#C#" A0 B$JN >5%)!*#5%?1'5N U(N >J f%V _%11Y`J %WJ KJ;J 6BI>XA$$N 6J6J ;B'0?'? A0 B$JN B$3,&1%/-$1+N F*N%>J 2WfY _%121`J
(J LJ ;I#$?cN -J )Bc#X?c!d#$#' A0 B$JN @.0-&#*,$332-'.N% *NN >J WU1 %(J LJ QA.A$!Q#/#EcBN M*A5%62-"S5%/* !1&35N OON >J 2Y _%11Y`J
tel-00708600, version 1 - 15 Jun 2012
_%11Y`J %UJ PJ!OJ ;@B'.N )J!;J l* A0 B$JN >5%)!*#5%?1'5%B$,$N NCN >J 2%&V _%11V`J
UJ dJ ;BXAC+B'N -JPJ LA+/BI? A0 B$JN )!"15% >5% )!*#5N MON >J %1Vf %VJ kJ l+0HA LA?'GAC/IBN -J K#G0 A0 B$JN K2*3%6-& *++"1'%/* !1&3&'.N
_%11&`J QN >J UY _%11U`J
VJ LJ OB.@?HBGA@N QJ k@B/AIHBGA@ A0 B$JN @.0-&#*,$332-'.N *GN %fJ 9J :#IB'/cBN -J LBCE?'?BcN >5%)!*#5%/!*-#5N FQN >J Uff _%11U`J
>J 22U _%11&`J %&J )J;J O#CCAE?$$BN LJ :AA0$AD/ A0 B$JN 6!.+5% )!*#5% )!*#5% 6!.+5N CGN
fJ 7J :A$#'/N kJ K?EBCG A0 B$JN )&#A$'1"*% ?2-&A**11*% B2% C"- &1"2#% >J U22( _%11&`J
)*D2+E%FG%9::9:HI95 >J%%1 _%11%`J %YJ 7J SB'/@A'.N [J [@#'.=?' A0 B$JN 6*,-&3*2#%M "*1 *N FN >J fW _%11V`J
&J 7J :A$#'/N ;J 7B.BCGA A0 B$JN )&#A$'1"*% ?2-&A**11*% 02% W1J MJ MAEc@#DDN MJ dB''.?ArAC A0 B$JN _<G/J`N @$10Z&&R% &S% 6-$ ," $3%
C"- &1"2#J)*D2+E%K-5L%K-%9::HJ==9IN >J (1 _%11V`J [JY$.%K32&-*+ *1 *%71$3.+"+E%;>C?'.AC!RAC$B.N MAC$?' _%11V`J