You are on page 1of 28

12.

17
Green Organometallic Chemistry
E G Hope, A P Abbott, D L Davies, G A Solan, and A M Stuart, University of Leicester, Leicester, UK
2007 Elsevier Ltd. All rights reserved.
12.17.1

Introduction

838

12.17.2

Alternative Solvents

839

12.17.2.1

Water

12.17.2.1.1
12.17.2.1.2

12.17.2.2

839

Biphasic hydrogenation
CC bond-forming reactions in water

839
840

Supercritical Fluids

12.17.2.2.1
12.17.2.2.2

840

Homogeneous hydrogenation
Asymmetric hydrogenation

841
843

12.17.2.3

Fluorous Media

844

12.17.2.4

Ionic Liquids

846

12.17.2.4.1
12.17.2.4.2

12.17.3

Hydrogenation
Palladium-catalyzed cross-coupling reactions

846
847

Enhanced Technologies

848

12.17.3.1

Microwave Chemistry

848

12.17.3.2

Sonochemistry

848

12.17.3.3

Microreactors

849

12.17.3.3.1
12.17.3.3.2

Heterogeneous catalysis in microreactors


Homogeneous catalysis in microreactors

849
849

12.17.3.4

Membranes for Integrated Catalysis and Separation

12.17.3.5

Thermoregulated Catalysis

850

12.17.3.6

Solid-supported Organometallic Systems

852

12.17.4
12.17.4.1

Hydroformylation: A Green Case Study I


Hydroformylation of Olefins on Supported Catalysts

850

852
852

12.17.4.2

Hydroformylation of Olefins in scCO2

853

12.17.4.3

Hydroformylation of Olefins in Biphase Systems

854

12.17.4.3.1
12.17.4.3.2
12.17.4.3.3

12.17.4.4

Hydroformylation of Olefins in Supported Liquid-phase Catalysis

12.17.4.4.1
12.17.4.4.2

12.17.4.5
12.17.5

Hydroformylation of olefins in aqueous media


Hydroformylation of olefins in fluorous media
Hydroformylation of olefins in ionic liquids
Hydroformylation of olefins in supported aqueous phase catalysis
Hydroformylation of olefins in supported ionic liquid-phase catalysis

Overview

854
854
855

855
855
855

855

Olefin Polymerization: A Green Case Study II

856

12.17.5.1

Metal-mediated Polymerizations of Olefins in Aqueous Media

856

12.17.5.2

Metal-mediated Polymerizations of Olefins in scCO2

858

12.17.5.3

Metal-mediated Polymerizations of Olefins in Fluorous Media

858

12.17.5.4

Metal-mediated Polymerizations of Olefins in Ionic Liquids

858

12.17.6

Future Perspectives

859

References

859

837

838

Green Organometallic Chemistry

12.17.1 Introduction
Green chemistry, a term only coined in the past decade, is not a new or alternative branch of science, rather an
approach, philosophical or conceptual, that can be applied throughout chemistry and chemical engineering. By its
very nature, it is extremely broad in that it can be applied to processes or procedures that may involve one or more of a
number of issues: (i) reduce or eliminate the use and generation of hazardous substances, (ii) reduce waste,
(iii) reduce energy consumption, and (iv) utilize renewable resources; these ideals are elegantly delineated in Paul
Anastas 12 principles of green chemistry.1 These concepts have not just been invented; indeed, the synthetic
chemist and chemical engineer have always sought better selectivities, higher activities, better catalysts, and
improved process conditions. Rather societal, political, and economic pressures for sustainable development have
focused the chemists attention (in both industry and academia) on providing products, valued and needed by society,
without damaging the environment. The green chemistryorganometallic chemistry interface is, therefore, very large,
and it might be argued that most, if not all, of organometallic chemistry seeks to deliver against one or more of
Anastas principles. Throughout this collection, specialist authors have highlighted, in context, developments in
green chemistry, whether it involves improved syntheses, catalysis, neoteric solvents, alternative energy sources,
renewable resources, or new technologies. In this contribution, we highlight the two principal aspects of organometallic chemistry: synthesis of organometallic compounds and the application of organometallic compounds in synthesis, as viewed from a green chemistry perspective, using hydroformylation and olefin polymerization as in-depth
illustrations. For further information, the reader is directed to the specialist contributions in this compilation, special
topics elsewhere in this compilation (Chapter 12.12), green chemistry websites,25 and textbooks.623
The comparison of the relative merits of alternative processes is probably the most fundamental problem facing the
development of green organometallic chemistry. Full life cycle assessment (LCA) is the only complete tool to
quantitatively evaluate the environmental impact of a particular process, but here the absence of data introduces
assumptions and approximations, which make it difficult to draw firm conclusions.24 In a direct comparison of the
asymmetric hydrogenation of ketoesters using enzymatic yeast, heterogeneous (Pt on alumina chinchinodine) and
homogeneous (RuBINAP) processes, the homogeneous processes offer the lowest cumulated environmental
impact.25 However, the authors here deduce that the environmental impact of a synthetic reaction depends more
on the process parameters than the individual technology, where volume of solvent, number of distillations, and
catalyst losses have the greatest impact on the LCA. Alternative approaches include the application of a variety of
metrics (Figure 1),2630 which often focus on individual components of green chemistry (waste generation, reaction
stoichiometry, amount of solvent or auxiliaries), but only seldom offer insight into the integration of technology and
chemistry to deliver overall environmental benefits.31
Atom economy is a simple metric for the initial comparison between different reactions that is receiving considerable attention.28 In an ideal scenario, all the atoms in the starting materials are incorporated into the product with
anything else needed in only catalytic amounts, and from this viewpoint, addition reactions are ideal green reactions.
Many types of addition reactions are known, catalytic hydrogenation and hydroformylation are examples that are

Figure 1 Commonly used green chemistry metrics.

Green Organometallic Chemistry

performed on very large scales, and even in complex organic synthesis, addition reactions (e.g., DielsAlder reactions)
are well established. Over the last 10 to 15 years, the use of ruthenium complexes in atom-economic transformations
has been greatly expanded from two-component to three- and even four-component coupling reactions, and the area
has been reviewed.3234 Similarly, impressive strides have been made in rhodium-catalyzed reactions, for example,
[4 2]-, [5 2]-, and [4 4]-cycloaddition reactions,35,36 redox isomerizations, for example, the conversion of
allylalcohols into ketones,3234 where aqueous techniques allow the catalyst to be recycled,37 and CH activation
and functionalization.3840 In 1993, Murai and co-workers reported the first example38 of a highly efficient, selective
ruthenium-catalyzed CH/alkene coupling reaction, which may in certain cases replace the palladium-catalyzed
processes that require stoichiometric base and halogenated or similarly functionalized aromatic starting materials, and
the process has since been reviewed.39 Catalytic methods of CH functionalization have also been reviewed
recently.39,40

12.17.2 Alternative Solvents


The replacement of relatively harmful volatile organic solvents with, alternative, less environmentally damaging
solvent systems/approaches is generating interest from a green perspective for synthetic chemistry and catalysis. A
number of these are reviewed elsewhere in this compilation. Here, we highlight some notable developments in green
applications of water, supercritical (sc) fluids, ionic liquids, and fluorous solvents in organometallic chemistry.

12.17.2.1 Water
Organometallic chemistry in water is covered elsewhere (Volume 1) and has been reviewed on several occasions.1618,4144
A wide variety of reactions have been catalyzed in water including hydrogenation of alkenes and ketones and
imines,45 conjugate addition,46 isomerization of allylic alcohols,37 and many Pd-catalyzed CC bond-forming reactions47,48 (see also Section 12.17.3.1). Various transition metal-catalyzed polymerizations have also been effected in
water, including CO/alkene polymerization,49 alkene polymerization,50 and ring-opening metathesis polymerization.51
Selected examples of organometallic catalysis in water are discussed below to illustrate the use of aqueous phase
reactions in green chemistry. These will be considered in two sections on aqueous biphasic hydrogenation catalysis,
which may be done on a larger scale where recycling of the catalyst is an important issue, and CC bond-forming
reactions, which are generally carried out on a smaller scale for which the use of water provides practical benefits in
terms of reduced hazard of the reagents involved.
The green credentials of water as a solvent are that it is nonflammable, not combustible, and nontoxic. In addition,
its polarity and density make it easy to separate from most organic compounds, and it has favorable thermal
properties. Hydrophobic effects and hydrogen bonding can each contribute to the rate enhancement of some
reactions in water.52,53 In addition, the low solubility of oxygen in water can facilitate the use of air sensitive
transition metal catalysts in air. The use of water as a solvent implies that there is no need for protectiondeprotection
processes for acidic hydrogen-containing functional groups, thereby increasing synthetic efficiency. Similarly, laborious derivatization processes are not necessary for water-soluble compounds such as carbohydrates. Potential drawbacks include the low solubility of organic substrates and reactions involving water sensitive compounds, though it is
noteworthy that larger rate increases are sometimes found for insoluble substrates.54

12.17.2.1.1

Biphasic hydrogenation

The need to retain catalysts in the aqueous phase has led to the development of specifically designed ligands. The
design principles are well illustrated by phosphine ligands.55 These fall into three groups: (i) those with anionic
substituents such as sulfonate, carboxylate, or phosphate groups, for example, 1a; (ii) those with cationic substituents
such as ammonium or guanidinium groups, for example, 2; and (iii) those with neutral hydrophilic groups such as
alcohol, phenol, or carbohydrate groups, for example, 3 (Figure 2). Water-soluble bidentate and chiral ligands have
also been synthesized based on the same substituents. Triphenylphosphine trisulfonate (TPPTS; 1a),56 used in the
industrial hydroformylation of propene, has a solubility of 1.1 kg l1.57
In a slight variation using an amine-substituted ligand, the reaction may be carried out homogeneously in an
organic solvent with addition of acid after the reaction to enable extraction of the ammonium-substituted catalyst into
an aqueous phase.58,59

839

840

Green Organometallic Chemistry

Figure 2 Water-soluble monodentate phosphine ligands.

In many cases, a co-solvent or surfactants are added to increase the solubility of the substrate in the aqueous phase.
In hydrogenation, the activity and selectivity is often decreased in water, but addition of a micelle forming
amphiphile leads to a significant increase in reaction rate and enantioselectivity. Using a rhodium complex of a
neutral bisphosphine, these effects were found with charged and non-ionic surfactants, with the biggest change in
activity and selectivity occurring near the critical micelle concentration.60 Indeed, enantioselectivity could be
observed using an achiral rhodium catalyst in the presence of carbohydrate amphiphiles.61 Chen et al. carried out a
systematic study of the use of surfactants on hydroformylation of higher alkenes with RhTPPTS complexes; in this
case, only cationic amphiphiles had a significant effect on yield and regioselectivity.62 The effect of amphiphiles has
led to the development of amphiphilic ligands.63 However, the use of amphiphiles leads to problems with separation
of the products, amphiphile, and catalyst after the reaction. Attempts to overcome this problem have included the use
of amphiphilic polymers.64 An alternative approach to catalysis in water that is experiencing a revival is coupling
organometallic catalysts with proteins. The huge advance in protein engineering in recent years means that such
catalysts are amenable to combinatorial optimization through genetic modification of the protein.65 Examples of this
approach applied to hydroformylation66 and asymmetric hydrogenation67 have been reported recently.

12.17.2.1.2

CC bond-forming reactions in water

Conventionally, organometallic reagents require dry solvents and inert atmosphere techniques, particularly for
organomagnesium and organolithium reagents that often also require low temperatures. Recently, alternatives
have been developed based on late transition and main group metals, for example, Pd, Rh, In, Sn, Zn.46,68,69
Examples include BarbierGrignard-type reactions of allyl halides with carbonyl compounds and imines,70
rhodium-catalyzed addition of aryl or alkenylboronic acids to enones71 or aldehydes,72 and palladium catalyzed
CC47,48 and CN73 bond-forming reactions, some of which are also performed with microwave irradiation.74
However, a drawback of all these reactions is the need for a halide starting material and the generation of
stoichiometric quantities of halide waste, and stoichiometric metal waste in several cases. Thus, direct CH
activation in water would provide a more green solution. Such processes have been demonstrated in the synthesis
of propargyl alcohols from catalytic CH activation of an alkyne and in situ reaction with an aldehyde.75 This
methodology has been extended to a three-component coupling of aldehyde, amine, and alkyne with a gold
catalyst in water.76 The importance of water is evidenced by low yields in organic solvents but almost quantitative conversions in water.
It is clear that apart from the green benefits in terms of reduced hazards and toxicity, the special properties of water
as a solvent can have dramatic effects on reactions. Further research in this area is very likely to produce new
reactions of general applicability.

12.17.2.2 Supercritical Fluids


Substances in the sc state have a unique set of physical properties that make them attractive alternatives as reaction
solvents. They have high miscibility with gases, liquid-like solvating power, and better-than-liquid transport properties, which invariably provide improved reaction rates. By far, the most commonly used fluid is CO2 because it is
inexpensive, nontoxic, nonflammable, environmentally benign, and has low critical constants (Tc 304.2 K; Pc 72.8
bar). Accordingly, it has been lauded as a replacement for volatile organic solvents. The sc fluids also offer the
potential to tune the solvent properties and affect yield, rate, and selectivity with pressure. In addition, the
morphology of the product can be controlled by rapid expansion of sc solutions, and selective extraction of products
from complex mixtures can be achieved by careful choice of solution density.

Green Organometallic Chemistry

The sc solvents have already found significant application for the extraction of natural products,77 purification of
lubricating oils,77 pharmaceutical processing, spray painting,78 and dry cleaning. The key advantage that has been
exploited with these processes is the lack of solvent residues following processing which is particularly important for
consumer products. The decreased emission of organic solvents associated particularly with the last two applications
could have a significant environmental impact. Multi-tonne processes such as the decaffeination of coffee or the
extraction of hops77 have led to industrial confidence in the application of sc fluids. Reactor systems capable of
operating up to 300 bar are readily available; hence, these solvents need not be considered as exotic media. While the
application of sc fluids is not advantageous for all reactions, there are specific applications where their use is clearly
desirable.
Examples of the industrial use of sc fluids for synthesis are less extensive with only fluorinated polymer production79 and heterogeneous hydrogenation that have been commercialized.80 Academic studies have, however, covered
the vast majority of reaction types. By far, the most comprehensive review is that by Leitner and Jessop,12 which
details not only all of the reaction types studied up to that point (1999) but also explains the underlying physical
principles of the fluids. The book includes stoichiometric organic reactions, photochemical processes, polymerization
reactions, heterogeneous catalysis, as well as metal complex-catalyzed reactions. A special issue of Chemical Reviews in
1999 (issue 2) covered a wide range of organometallic reactions in sc fluids.81 A condensed overview of CO2 as a green
reaction medium is presented by Leitner.82 Numerous other reviews detail specific reaction types and compare CO2
with other solvents.8387
One of the main difficulties experienced with the design of catalysts for sc fluids is the low polarity, particularly, of
CO2 (the dielectric constant ranges from 1.5 to 1.7 depending on pressure).12 This means that most organometallic
complexes traditionally used in conventional liquid solvents are insoluble in scCO2. This has led to most studies
being carried out using heterogeneous catalysts. Comprehensive reviews of heterogeneous hydrogenations in sc fluids
have been published.12,88
The design of catalysts for use in scCO2 involves the same principles as those required for fluorous biphase
catalysis (vide infra), that is, fluorous- or trialkyl-phosphine groups are used as ligands.89,90 An alternative approach is
to use conventional catalysts and change the sc fluid. Hope and co-workers have used scCF2H2 and scCF3CFH2 for a
variety of reaction types and showed that hydrogenation catalysts, such as Wilkinsons catalyst, are soluble without
modification.91 A few detailed cases are given below using hydrogenation as a specific reaction. The solubility of
substrates and catalysts has been studied for most reactions and the majority of the data is brought together in two
studies. Depressurization of the solution at the end of the reaction leads to a complex mixture of product and catalyst,
and a number of methodologies have been devised to circumvent this. The solvent polarity can be tuned using
changes in temperature or pressure, and this has been shown to be effective at separating the catalyst from the
substrate.90 Work by Leitner and co-workers also showed that facile solute separation could be achieved when the
substrate acted as a co-solvent, but the product was a catalyst antisolvent.92 A membrane reactor was used by Broeke
and co-workers to carry out the continuous, homogeneous hydrogenation of 1-butene using a fluorous derivative of
Wilkinsons catalyst in scCO2.93 The silica membrane in the reactor had a pore size of 0.50.8 nm, whereas the
catalyst was 24 nm in size. The substrate and product were able to diffuse through the membrane and, therefore,
efficient catalyst recovery was facilitated. The catalyst was prepared in situ and conversions of around 40% were
obtained during the investigation. The sc fluids have also been used in biphasic catalysis, and the use of sc fluids in
conjunction with ionic liquids is seen as being particularly desirable, as CO2 can act as both a medium to extract the
product and also as an expansion medium to decrease solvent viscosity.9496 An overview of catalyst immobilization is
given in a recent review.97

12.17.2.2.1

Homogeneous hydrogenation

The rates of many hydrogenation reactions in liquids are proportional to hydrogen concentration and are sometimes
limited by the rate of diffusion of hydrogen from the gas to the liquid phase.98 These problems are overcome by the
use of sc fluids as reaction media which offer a significant potential rate advantage for hydrogenation reactions over
more conventional processes. It should be noted that in scCO2, the insertion of CO2 into the metalhydride bond of
catalytic species to produce formate complexes is also possible.99,100 This has the potential to inhibit hydrogenation
reactions in scCO2, depending on the ability of the formate to revert back to the hydride and CO2. One of the first
examples of hydrogenation in scCO2 (Equation (1)) was the hydrogenation of 3,3-dimethyl1,2-diphenylcyclopropene 4 by [MnH(CO)5] via a radical mechanism.101 Depending on the alkene and the solvent, either hydrogenation
or hydroformylation products were observed.101,102

841

842

Green Organometallic Chemistry

The selectivity for hydrogenation over hydroformylation has been used as a measure of the strength of the solvent
cage, with stronger cages favoring hydroformylation where the selectivities obtained from radical hydrogenations in
scCO2 were similar to those obtained in liquid organic solvents.101,103
The use of CO2 as a starting material for the synthesis of organic compounds has been of great interest to synthetic
chemists. The hydrogenation of CO2 has been particularly attractive but remains difficult under conventional gasphase conditions. In 1994, Noyori and co-workers reported the use of ruthenium(II) phosphine catalysts 5 or 6
(Figure 3) for the hydrogenation of CO2 to formic acid 7 (Scheme 1) employing scCO2 as both solvent and
substrate.99 The catalysts were chosen for their known solubility in non-polar solvents such as hexane and were
shown to be more active than [RuH2{P(C6H5)3}4], which was considered to be the most active catalyst in benzene
solution. Other reactions include the one-pot synthesis of dimethylformamide (DMF) and methyl formate from
scCO2.100 Catalyst 5 showed a catalytic activity which was two orders of magnitude greater than any reported in liquid
solvents, giving 99% selectivity to DMF production and a conversion of 99% for the dimethylamine.104
It is only over the last few years that the mechanistic aspects of homogeneous hydrogenation in sc fluids have been
investigated. One of the first such investigations, Equation (2), was carried out by Leitner and co-workers using
dimethyl itaconate 8 and a rhodium catalyst 9 as a model reaction in both hexane and scCO2.105 They concluded that
the mechanism was the same in both solvents and that CO2 did not interfere with the activation and transfer of the
hydrogen molecule at the catalyst.

Figure 3 Representative examples of catalysts used for hydrogenation in sc CO2.

O
H

OH

7
Scheme 1

6
scCO2
373 K
NEt3
99%

CO2
120 bar

H2
85 bar

5
scCO2
373 K
HNMe2
99%

NMe2

Green Organometallic Chemistry

12.17.2.2.2

Asymmetric hydrogenation

The enantioselectivity of asymmetric hydrogenation of a prochiral olefin depends strongly on the hydrogen
concentration, and higher concentrations can lead to higher, lower, or reversed enantioselectivity.12 It has
been suggested by Burk et al. that the miscibility of hydrogen gas in sc fluids may lead to better enantioselectivity.106 They studied the asymmetric hydrogenation of -disubstituted enamide 10 in scCO2, which gave the
valine derivative (R)-11 in 85% ee (Equation (3)).106 The previous highest reported enantioselectivity for this
substrate was 55% ee.107 The same reaction was carried out in hexane pressurized to 345 bar with nitrogen. This
additional experiment indicated that this enhancement of enantioselectivity was not a pressure effect but was
associated with the use of scCO2. In the asymmetric hydrogenation of several -enamides in scCO2 using the cationic
Rh complexes 12a or 12b (Equation (4)), the ees obtained were comparable to those obtained in methanol and
hexane.106

Catalyst

Yield (%)

ee (%)

12a
12b
12a
12a
12b
12b
12b
12b

H
C2H5
C6H5
3,5-(CF3)2C6H3
H
C2H5
C6H5
3,5-(CF3)2C6H3

100
100
100
100
100
100
100
100

99.5
98.8
99.5
99.5
99.1
98.8
99.1
99.1

While fluoroalkylated monodentate phosphite, phosphonite, and phosphoramidite ligands offer good selectivity
and enantioselectivity in the Rh-catalyzed asymmetric hydrogenation of dimethylitaconate in CH2Cl2, the levels of
reactivity and selectivity are not retained in scCO2.108 In contrast, Xiao et al. found that ,-unsaturated carboxylic
acids, such as tiglic acid 13, could be hydrogenated in scCO2 by [Ru(OCOCH3)2{(S)-H8-BINAP}] 15 to give (S)-2methylbutanoic acid 14 (Equation (5)).109 The product was obtained with an 81% ee, which is comparable to that in
methanol (82%) and greater than that in hexane (73%). The partially hydrogenated BINAP catalyst 15 (which gave
99% yield) showed increased solubility in scCO2 over the fully aromatic derivative, which gave poorer yield (50%)
and selectivity (37% ee). Addition of the perfluorinated alcohol CF3(CF2)6CH2OH to 15 offered solubility enhancement and produced an increase in enantioselectivity (89% ee). However, Xiao and co-workers add a note of caution
regarding the apparent enhancements observed when polar co-solvents are introduced into sc catalytic systems. In a
series of experiments on the RuBINAP-catalyzed asymmetric hydrogenation of dimethyl itaconate in scCO2, the
high reaction rates and enantioselectivities obtained in the presence of MeOH could arise from pre-reaction in a
CO2-saturated liquid solvent.110

843

844

Green Organometallic Chemistry

Leitner and co-workers have used scCO2 for the homogeneous iridium-catalyzed hydrogenation of prochiral
imines.92 Cationic iridium(I) complexes 16, modified with perfluoroalkyl groups for increased CO2-philicity, were
synthesized and used in the hydrogenation of N-(1-phenylethylidene)aniline 17 to give (R)-N-phenyl1-phenyl
ethylamine 18 (Equation (6)). In 6 h with 0.078 mol% catalyst, 81% ee was obtained in scCO2, while, in dichloromethane, more than 22 h and 5 times the amount of catalyst were required to give comparable results.

In conclusion, it can be seen that organometallic catalysis in sc fluids has moved on significantly during the last
decade. Many of the issues associated with catalyst solubility and separation have been addressed. Also, since many of
the technical difficulties concerning high-pressure chemistry have been solved, it should be anticipated that over the
next 10 years, some processes could be commercialized and new ligand chemistry could provide opportunities to
extend the types of reactions studied in these solvents.

12.17.2.3 Fluorous Media


In 1994, Horvath and Rabai introduced a new approach to the heterogenization of homogeneous catalysts called
fluorous biphase catalysis.111 In essence, the approach is very similar to that in the aqueous biphase system, but the
aqueous phase is replaced by perfluorocarbon solvents that are immiscible with most common organic solvents at
ambient temperature. By modifying ligands with long perfluoroalkyl groups, often nicknamed fluorous ponytails, the
catalyst can be anchored in the fluorous solvent while the substrate(s) and product(s) are preferentially soluble in the
upper organic phase. Upon heating or under pressure, the fluorous biphase system (FBS) becomes monophasic,
allowing genuine homogeneous catalysis to occur, and on cooling, the transition is reversed so that the two discrete
phases are observed again, enabling straightforward separation and recycling by simple decantation. For example,
Schneider and Bannwarth synthesized four different perfluoroalkylated analogs of triphenylphosphine and investigated their applications in palladium-catalyzed Suzuki couplings (Scheme 2).112 The FBS was nearly homogeneous
at 75  C enabling the catalysis to occur, while at room temperature, the fluorous catalysts partitioned into the fluorous
phase and were recycled 6 times without significant decrease in the yield. Although a quantitative conversion could
be achieved when the catalyst loading was reduced from 1.5 mol% to 0.1 mol%, the yields dropped dramatically on
recycling (89%, 78%, 50% for 19a).
This approach has now been applied to a broad range of catalytic reactions, such as hydrogenation, hydroformylation, hydroboration, and many palladium-catalyzed cross-coupling reactions, and excellent catalyst recovery and reuse
have been reported.113119 Although there are many comprehensive reviews of fluorous biphase catalysis,113119 the
Handbook of Fluorous Chemistry provides the most detailed accounts of all aspects of fluorous chemistry and includes
an excellent experimental section on optimized procedures from the literature.13 As it stands, fluorous biphase

B(OH)2
+
Br

OMe

PdCl2L2 (1.5 mol%)


CF3C6F11, (MeOCH2)2
2M Na2CO3, 75 C, 2 h

L=P
R

Scheme 2

OMe

19a: R = 4-C2H4C8F17 87, 90, 92, 89, 86, 90%


19b: R = 3-C2H4C8F17 95, 95, 93, 92, 94, 89%
19c: R = 4-C8F17 95, 93, 91, 90, 91, 88%
19d: R = 3-C8F17 99, 96, 90, 86, 83, 88%

Green Organometallic Chemistry

Figure 4 Representative examples of perfluoroalkyl supports and catalysts.

catalysis is not a viable option for industry because of the prohibitive expense and the environmental persistence of
the perfluorocarbon solvents. Fluorous reverse-phase (FRP) silica gel 20 (Figure 4) was developed originally by
Curran for high-throughput organic synthesis because it can be used to separate fluorous-labelled compounds from
conventional organic molecules on a preparative scale.120 Recently, it has been used to develop two new methods for
recycling fluorous catalysts without using perfluorocarbon solvents: (i) fluorous solid-phase extraction, (ii) solidsupported catalysis.
In the first approach, a fluorous SCS pincer palladium complex 21 was used to promote the Heck reaction
(Equation (7)) under microwave or thermal heating.121 At the end of the reaction, the crude reaction mixture was
charged directly onto a short column of fluorous silica and was eluted with 90% MeOHH2O to give the organic
fraction which was then filtered through conventional silica gel to give the pure product. By changing the solvent to
diethyl ether, the fluorous catalyst was recovered quantitatively from the fluorous silica gel and was reused twice
without loss of activity. The major advantages of this approach are that: (i) fewer and shorter fluorous ponytails are
required (<40% fluorine content by molecular weight) compared to fluorous biphase catalysis (>60% fluorine content
by molecular weight) and (ii) fluorous solvents are no longer required for either the reaction or separation step. The
only other successful organometallic example of this technique is the recycling of light fluorous GrubbsHoveyda
catalysts 22 and 23 after alkene metathesis.122
(3 mol%)

R1

R2

Yield (%)

OMe
CN
COMe
COMe

Ph
Ph
Ph
Ph

l
Br
l
Br

76
78
89
92

In the second approach, Bannwarth and co-workers123 immobilized fluorous-tagged bis(triphenylphosphine)


palladium catalysts 19a19d, on FRP silica gel for Suzuki and Sonogashira reactions in conventional organic solvents.
Although the supported catalysts showed activities comparable to those found in the FBS, the recycling results were
only good for highly active substrates. Recent detailed studies have demonstrated that the activity of the solidsupported catalysts does actually decrease in organic solvents, and this was attributed to catalyst decomposition rather
than catalyst leaching.124 On the other hand, their behavior in water is completely different and the activity remains
high upon recycling.124,125 In related work, Hope and Stuart have been investigating the applications of alternative
fluorinated solid supports, such as perfluoroalkylated polystyrene beads and amorphous perfluoroalkylated zirconium
phosphonates, in the Heck reaction, hydrogenation, and cyclopropanation.126
Besides rising to the challenge of developing new methods for recycling fluorous catalysts, there are also many reports
on the synthesis and applications of new perfluoroalkylated chiral catalysts.127129 It is difficult to envisage the successful
application of the FBS for asymmetric catalysis, but these new lightly fluorinated approaches offer significant potential.

845

846

Green Organometallic Chemistry

12.17.2.4 Ionic Liquids


Ionic liquids offer a number of potential advantages over organic solvents from a green perspective. Loss of solvent by
evaporation is effectively zero. Reactions may be more selective in an ionic liquid, thereby reducing separation costs
both from an economic and environmental perspective. For catalytic reactions as well as increasing selectivity, an
ionic liquid may stabilize a catalyst and so increase lifetime and turnover number. The use of ionic liquids in catalysis
has been the subject of several reviews in recent years,130135 and organometallic chemistry in ionic liquids is
reviewed in Volume 1.
Most studies of catalysis in ionic liquids have focused on issues of increased selectivity and particularly the easy
separation of product from the catalyst and catalyst recycling via use of a biphase. In some cases, the reaction may
occur in a biphase; in others, the biphase is only used for product separation. In some special cases, the second phase
is exclusively product, due to insolubility of the organic products in the ionic liquid, and is easily separated by
decantation, allowing the recovered ionic catalytic solution to be reused. Of course, use of an organic solvent for
extraction does reduce some of the potential green benefits of the ionic liquid approach. More recently, scCO2 has
been used to extract the products. Alternatively, volatile products can be separated from the ionic liquid and catalyst
by distillation.
An advantage of ionic liquids compared to other catalyst immobilization strategies is that many catalytically active
transition metal complexes can be immobilized in ionic liquid solvents without the need for specially modified
ligands. However, leaching to an organic phase may occur; hence, charged ligands may be preferable to minimize
leaching. Hence, many ligands, such as sulfonated phosphines, developed for aqueous biphase reactions, work well in
ionic liquids. In the early 1990s, Chauvin et al. demonstrated the oligomerization of alkenes catalyzed by a nickel
complex in an organoaluminate ionic liquid in the presence of AlEtCl2.136 In this case, the liquid products formed are
immiscible with the ionic liquid, and the reaction could be run in a batch or semicontinuous process. The discovery of
air and moisture stable ambient-temperature ionic liquids,137,138 such as 1-n-butyl3-methylimidazolium [BMIM]
salts with [PF6], [BF4], [NTf2], [OTf], or Br as anion, has led to a huge increase in the use of ionic liquids, and a wide
range of reactions have now been catalyzed in ionic liquids including hydrodimerizations,139 alkoxycarbonylations,140
hydroformylation,138,141 epoxidations,142,143 alkene metathesis,144,145 hydrogenations, and a wide variety of palladium coupling reactions. For further details of these and other reactions, readers are directed to the reviews
mentioned above. Hydrogenation and palladium-catalyzed cross-coupling reactions are discussed here as illustrations
of the potential beneficial effects and complications arising from the use of ionic liquids as green solvents.

12.17.2.4.1

Hydrogenation

Hydrogenations which occur by oxidative addition,137,138 heterolytic activation,146 and homolytic cleavage146 of
hydrogen can all occur in ionic liquids. Alkane products are easily recovered by decantation and the recovered catalyst
solution can be reused several times. It is noteworthy that isomerization products were not observed in rutheniumcatalyzed hydrogenation of 1-hexene performed in [BMIM][BF4]146 though they are formed in aqueous biphase
reactions. However, Chauvin and co-workers reported that isomerization did occur for rhodium-catalyzed hydrogenation of 1-pentene; when the ionic liquids contained trace amounts of chloride ions, the rate of hydrogenation dropped
while the selectivity to the isomerization product pent-2-ene increased.138
The catalytic hydrogenation of arenes with an arene ruthenium catalyst precursor has also been performed in ionic
liquids and the products separated from the ionic solution by distillation under high vacuum, allowing the same batch
of ionic liquid to be used repeatedly for the catalytic hydrogenation of several different arenes.147 Interestingly,
colloidal rhodium catalysts can hydrogenate arenes in aqueous/sc fluid biphasic media whereas the reactions do not
work in [BMIM][BF4].148
Hydrogenation has also been extended to asymmetric versions using an RuBINAP complex dissolved in
[BMIM][BF4] with isopropanol as a co-solvent to solubilize the substrates.149 The enantioselectivity is comparable
to those obtained in conventional organic solvents though the ionic liquid did increase the catalyst stability. Similarly,
an RhDUPHOS catalyst was found to be much more air stable in [BMIM][PF6] than in an organic solvent and could
be reused five times with very little drop in enantioselectivity.150 The effect of hydrogen pressure on enantioselectivity is not simple, and apparently conflicting results have been found using [BMIM][BF4] and [BMIM][PF6]. These
results were explained on the basis of differing concentrations of molecular hydrogen in the ionic liquids;151 however,
a more recent study suggests that there is little difference in the solubility of H2 in the pure ionic liquids.152 More
recently, asymmetric hydrogenation of tiglic acid has been performed in [BMIM][PF6] with some added water, with
extraction of the products by scCO2.153

Green Organometallic Chemistry

12.17.2.4.2

Palladium-catalyzed cross-coupling reactions

The first reported example of palladium-catalyzed CC bond-forming reactions in an ionic liquid was the Heck
reaction in ammonium and phosphonium halide ionic liquids using simple catalyst precursors such as [PdCl2] or
[Pd(OAc)2]3 without phosphine ligands.154 The products could be distilled from the ionic liquid and the
remaining catalyst solution still showed catalytic activity. Seddon and co-workers reported that Heck reactions
proceed better in hexylpyridinium rather than dialkylimidazolium-based ionic liquids. Addition of phosphine
inhibited the reaction in [C6py]Cl but increased the reactivity in imidazolium-based ionic liquids.155 Herrmann
and co-workers also noted poorer reactivity in imidazolium-based ionic liquids compared to tetraalkylammonium
salts.156,157 The first evidence that palladium N-heterocyclic carbene (NHC) complexes were being formed came
from the observation that the Heck reaction proceeds more efficiently in [BMIM]Br than [BMIM][BF4] and that
palladium black precipitated in the latter solvent but not the former.158 In addition, NHC complexes, 24 and 25,
(Figure 5) could be isolated from the reaction in [BMIM]Br. While NHC complexes are implicated in some cases,
catalysis by palladium nanoparticles is an alternative proposition, particularly where the formation of NHC
complexes is not possible.159
The Suzuki coupling160 of 4-bromotoluene with phenylboronic acid catalyzed by [Pd(PPh3)4] has been carried out
in [BMIM][BF4].161 The best results were achieved by preheating the aryl halide in the ionic liquid with the Pd
complex before adding the arylboronic acid and Na2CO3. The use of [BMIM][BF4] has several advantages over
toluene as a solvent; a significant increase in reactivity (turnover frequency (TOF) 455 h1 in [BMIM][BF4], in
comparison to 5 h1 in toluene); the formation of homocoupling aryl byproducts is eliminated; the reaction can be
performed in air; and the catalyst can be reused 3 times without loss of its activity.161 In these cases, the products
were extracted with diethyl ether, and the inorganic byproducts, which modify the miscibility of water with
[BMIM][BF4] such that two phases are formed at ambient temperature, were removed by washing with water,
leaving the clean ionic liquid catalyst solution. Alternatively, the product could be isolated from the [BMIM][BF4]
reaction mixture by sublimation or precipitation by addition of water, all without any apparent leaching of palladium
species into the product. Welton and co-workers have now shown that for Suzuki reactions, in imidazolium-based
ionic liquids, NHC complexes are always found when the reaction is successful and not when the reaction is
unsuccessful.162
In conclusion, it was initially assumed in many reactions that ionic liquids were acting as entirely innocent noncoordinating solvents; this is now known not to be the case.163 In several examples of palladium-catalyzed reactions,
formation of NHC complexes is established;162 however, even in these cases, the precise nature of the active catalyst
is not known since palladium nanoparticles are known to be stabilized in ionic liquids.159 Other problems may arise
through the presence of halide impurities from the initial preparation of the ionic liquids which can poison many
metal catalysts, particularly cationic complexes.138,164 Although ionic liquids might be thought to be ideal to stabilize
ionic species, Daguenet and Dyson have suggested that dissociation of a halide from a transition metal complex can
be thermodynamically disfavored in an ionic liquid.165 Also, particularly in [BMIM][PF6], traces of water can lead to
hydrolysis of the anion, generating HF and phosphate and subsequently metal fluorides.166 On a more general note,
as solvents, imidazolium-based ionic liquids are rather expensive, certainly when compared to traditional organic
solvents, let alone water. In this regard, the use of a supported ionic liquid phase, analogous to the well-established
supported aqueous phase catalysis,167 has been reported in a hydroformylation reaction168 (Section 12.17.4.4.2). In
addition, the toxicity of ionic liquids has not yet been extensively investigated.169,170 Notwithstanding these issues,
the enhanced stability and selectivity of many catalysts in ionic liquids mean that further exploration of the biphase
applications of ionic liquids is likely in the next few years.

Figure 5 NHCpalladium complexes identified in [BMIM]Br/Pd(II) systems.

847

848

Green Organometallic Chemistry

12.17.3 Enhanced Technologies


A variety of new and old technologies are generating interest from a green perspective for synthetic chemistry and
catalysis, including alternative energy sources, process intensification, and catalyst handling. A number of these have
found relatively few applications in organometallic chemistry and others are reviewed elsewhere in this compilation.
Here, we highlight some notable developments in green applications of microwaves, sonochemistry, microreactors,
membranes, thermoregulated and supported catalysts in organometallic chemistry.

12.17.3.1 Microwave Chemistry


Microwaves can provide a rapid and efficient method of heating, and as such microwave chemistry is attracting
interest from the point of view of green chemistry; microwave heating, microwave-assisted synthesis and scale-up
have been reviewed.171175 Microwave heating is not only rapid but also occurs directly in the solution rather than via
the walls; hence, side-reactions, particularly catalyst decomposition, may be slower, giving longer catalyst lifetimes
and/or lower catalyst loadings.
The recent advent of microwave reactors specifically designed for chemistry has led to an increase in reproducibility in conditions and, hence, reaction outcomes, leading to a much wider use of microwave chemistry. However,
most reactions have been investigated on rather small scales and, therefore, the greatest impact of microwave
chemistry to date has been in drug discovery programs in the pharmaceutical industry.171,176,177 Here, the increased
speed and yield of reactions, particularly on polymeric supports, have led to a rapid acceptance of this methodology.
In terms of organometallic chemistry, microwave chemistry has been used in the preparation of several simple organometallic complexes, for example, [ML(CO)4] (M Cr, Mo, W; L en, bipy, dppm, dppe),178,179 which might be routinely
prepared in undergraduate laboratories. Other examples include the preparation of cyclometallated iridium complexes that
may have application as OLEDs.180 Recently, microwave acceleration of homogeneous catalysis has been investigated and
the topic has been reviewed.181 Perhaps, the greatest impact of microwaves on organometallic catalysis has been in the area
of palladium-catalyzed reactions such as the Heck, Sonogashira, Suzuki, and Stille cross-couplings. Many such reactions
have the additional green benefit of being performed in water,182 since this is reasonably effective at absorbing microwaves,
or sometimes with no solvent at all. Moreover, there have been reports that Suzuki and Sonogashira-type coupling reactions
can be performed in water with microwave irradiation with no transition metal catalyst.183185 Leadbetter has subsequently
shown that small impurities (50 ppb) of palladium in the Na2CO3 used as base were actually responsible for the catalysis
effectively at very low loading.186 The same group has since published similar methodology for performing the Heck
reaction using ultralow metal catalyst concentrations.187 The microwave-assisted Suzuki reaction has been scaled up in a
continuous flow capillary reactor using a supported palladium catalyst,188 and in an automated batch process.189
Other examples of microwave-assisted catalysis include allylic alkylation, both palladium catalyzed190 and molybdenum catalyzed.191 In the latter case, air stable precursor complexes could be used under non-inert conditions.
Microwave-enhanced PausonKhand reactions have also been reported,192 as have hydroamination of alkynes,193 and
metathesis of functionalized alkynes.194 Recently, microwave enhancement has been applied to CH activation
reactions, for example, for the formation of functionalized heterocycles,195 allowing the reaction to be performed with
no solvent purification and minimal precautions to exclude air. A solvent-free chelation-assisted hydroacylation
reaction has also been reported.196

12.17.3.2 Sonochemistry
Volume 1 includes a thorough review of the applications of sonochemistry in organometallic chemistry, and here we
present selected examples to illustrate the green advantages arising from the extreme localized conditions generated
by cavitational collapse under acoustic radiation; rate enhancements, catalyst regeneration, non-classical reagents,
unusual selectivities.197 In heterogeneous hydrogenation over Pd black or Raney-Ni, up to 20-fold improvements in
reaction rates under ultrasound conditions are reported, with better recycle of the catalyst, which appear to be
ascribed not only to the extreme reaction conditions but also to decontamination of the surface active sites and
creation of fresh active sites as a direct consequence of the mechanical stress in the catalyst induced by the ultrasonic
field.198,199 The extension to the enantioselective hydrogenation of 1-phenyl-1,2-propanedione over 5% Pt/silica
fibers modified with cinchonidine afforded increases in initial reaction rate and enantioselectivity of 4 and 2 times,
respectively, under sonication.200 In Suzuki coupling reactions, up to fourfold improvements in the reaction rate are
reported with ultrasound.201,202 Issues of scale-up of sonochemical reactors are now being contemplated.203

Green Organometallic Chemistry

12.17.3.3 Microreactors
Traditional chemical manufacturing, which is heavily dependent on large batch processes and associated transport
and storage of raw materials, brings with it risk, in terms of health and safety considerations, for both the operators and
the local communities. The miniaturization of chemical reactors offers some fundamental and practical advantages,
including the minimization of chemical storage and inventory, improved in situ reaction monitoring and control,
solvent-free mixing, in situ reagent formation and integrated separation, often alongside increased reaction rates,
improved reaction yields and selectivity. Furthermore, advocates of this new technology identify scale-out as opposed
to scale-up (i.e., increasing the number of parallel reaction channels as opposed to building larger reactors) as an
intrinsically safer philosophy when converting laboratory-based reactions into full-scale plant operation.204207
Essentially, microreactors consist of a series of channels (10300 mm) etched, by substrate-dependent techniques,
for example, photolithography, laser microforming, injection molding, or hot embossing,208 into a solid substrate
(silicon, quartz, glass, metals, polymers).209,210 For typical organometallic chemistry or catalytic processes, the
channel networks are linked to a series of reservoirs, containing reagents, catalysts, or products, to form the complete
device, which is typically no more than a few cm3 in volume.211 Under process conditions, reagents are brought
together, in specific sequences, mixed, and allowed to react in a carefully controlled region of the device using
electrokinetic (electroosmotic or electrophoretic) or hydrodynamic pumping. In this way, reaction profiling (temperature, pressure, small reaction inventories, flow rate, reagent concentrations, real time product separation)
provides a level of reaction control that is unimaginable in classical bulk stirred-tank reactors. Although this
technology is still in its infancy, organic synthesis research has shown that microreactor methodology is compatible
with both gas-phase and liquid-phase chemistry, offering considerable potential for safer and more efficient processes. The range of organometallic synthesis and catalysis is currently limited and, here, representative illustrations
of the application of microreactors in these areas are highlighted.

12.17.3.3.1

Heterogeneous catalysis in microreactors

The microreactor technology is particularly suited to heterogeneous catalysis, where either the walls of the reaction
channels can be used as the catalyst support or a supported catalyst can be placed in the reaction channels.
Impregnation of a solution of [Pd(acac)2] into an aluminum wafer microreactor with mechanical and anodical etching
generated microchannels with 0.18% palladium metal in layers 18 mm thick.212 This device was used in the gas-phase
hydrogenation of benzene, cyclodienes, and cyclotrienes, with controlled selectivity to the partially or fully reduced
products. Similarly, palladium-coated silica channels, deposited from [Pd(PPh3)4], have been used for the rapid roomtemperature hydrogenation of a variety of alkenes and alkynes under triphasic conditions, with no detectable metal
leaching to the organic product.213 Yields comparable to those reported under homogenous conditions, together with
negligible metal leaching and the absence of an external base, were obtained in palladium-catalyzed Suzuki reactions
using 1.8% palladium on silica sandwiched between microporous silica frits within microreactor reaction channels.214

12.17.3.3.2

Homogeneous catalysis in microreactors

Increased reaction rates, over those in conventional reactors, were reported for the KumadaCorriu cross-coupling
reaction using an Nisalen catalyst anchored onto a Merrifield resin as a catalytic plug in a polytetrafluoroethylene
(PTFE) microreactor.215 Issues of channel blocking caused by the polymer swelling under process conditions were
overcome using the same Nisalen catalyst, tethered, with a C11H22 linker, to a functionalized silica support.216
Although reaction rates remained high in the microreactor, a gradual reduction in performance was apparent over
extended reaction runs (>5 h). This was ascribed to blocking of active sites on the supported catalyst by reaction
byproducts (MgBr2, MgBrCl). Increased reaction rates (20) are also reported for a Pdsalen catalyst supported on a
Merrifield resin in the SuzukiMiyaura reaction in a microreactor.217 Here, the supported catalyst is solvent
expanded to completely fill the reaction channel such that the liquid path is through the microchannels of the
macroporous resin structure. In the asymmetric hydrogenation of a series of substrates with RhJosiphos/Diop
catalysts, enantioselectivities obtained in a single channel falling-film microreactor were comparable to those
obtained in conventional batch reactors.218 However, the key advantage is the extremely low catalyst loading,
typically 0.1 mg, allowing catalyst screening to be accomplished on substantially reduced catalyst inventories. In a
combination of microreactor and aqueous biphase technologies, the RuTPPTS-catalyzed hydrogenation of
,-unsaturated aldehydes under plug flow, alternating organic and aqueous slugs, has been reported.219

849

850

Green Organometallic Chemistry

12.17.3.4 Membranes for Integrated Catalysis and Separation


Membranes are either permeable or semi-permeable thin films, which offer advantages, such as lower energy
requirements and reductions in capital investment, over conventional separation processes, and are finding increasing
applications in the petrochemical and pharmaceutical industries. Essentially, membranes function by controlling,
either by sieving or by influencing relative rates of transport, the exchange of materials between two adjacent fluids,
and current membrane processes include microfiltration, ultrafiltration, nanofiltration, and gas/vapor separation.
Membrane materials include porous inorganic solids (silica, alumina, zeolites, zirconia), organic polymers (polypropylene, fluoropolymers, polyamides, polysulfones, silicones), and microporous carbons. However, the major step
forward in green chemical technology is the combination of the separation step with the synthetic step using
membranes in the membrane reactor concept, which has been made possible with the development of organic
solvent-resistant membranes. This coupling of the two key processes results in the continuous separation/feed of
reactants/products, which can result in enhancements in selectivity and/or yield.220,221 The application of this
technology in organometallic chemistry can be illustrated by hydrogenation and CC bond-coupling processes. In
the reduction of acetophenone in isopropanol, a polysiloxane-bound ruthenium(II)-transfer hydrogenation catalyst
afforded excellent enantioselectivity (97%) with slightly reduced activity (as compared to the homogeneous catalyst)
in a continuous flow system equipped with an MPF-50 membrane (Koch Membrane Systems, Germany).222
Membrane technology can also be combined with catalysis in scCO2.93 Using a perfluoroalkylated Wilkinsons
catalyst analog, van Koten and co-workers have reported the continuous hydrogenation of 1-butene with a tubular
microporous silica membrane (ECN Petten, the Netherlands), where high reaction rates were coupled with low
(<0.1%) rhodium leaching levels.93 Metal leaching levels were poorer in a palladium-mediated cross-coupling
synthesis of trans-stilbene following post-reaction nanofiltration with a polyimide Starmem 122 membrane
(W.R. Grace & Co., USA), but this was ascribed to catalyst decomposition rather than difficulties with the membrane
technology.223

12.17.3.5 Thermoregulated Catalysis


Bergbreiter224 synthesized smart phosphine ligands 26 and 27 (Figure 6) from an ethylene oxidepropylene oxide
ethylene oxide triblock co-polymer that possesses a property termed inverse temperature-dependent solubility.225 This
means that at low temperature, the phosphines and their rhodium complexes are soluble in water through the formation
of hydrogen bonds between oxygen atoms and water molecules, but, on heating above the critical-temperature cloud

Figure 6 Representative examples of ligands used in thermomorphic catalysis.

Green Organometallic Chemistry

point (Cp), they phase separate from the aqueous systems on account of the disruption of the hydrogen bonds.
Bergbreiter elegantly illustrated this concept by demonstrating that the cationic rhodium complex of 27 catalyzed
the hydrogenation of allyl alcohol in an aqueous system at 0  C, but on heating to 4050  C, the reaction stops. What is
most striking is the reversibility of this process, and by cooling the reaction mixture back down to 0  C, the catalyst is
rehydrated and the hydrogenation of allyl alcohol continues.
The inverse temperature-dependent solubility in aqueous media of polymer-bound palladium(0)phosphine
catalysts, based on the water-soluble polymer poly(N-isopropyl)acrylamide (PNIPAM) 28, was also used to recycle
and reuse these catalysts in nucleophilic allylic substitutions (Equation (8)) and cross-coupling reactions between aryl
iodides and terminal alkynes (Equation (9)). The catalyst was highly active in both reactions, and it was recycled 10
times with an average yield of 93% in the allylic nucleophilic substitution by precipitation with hexane.226

8
R1

R2

NuH

Temp (C)

Time (h)

Yield (%)

H
Ph
Ph

Me
Me
OEt

4-MeC6H4SO2Na
4-MeC6H4SO2Na
4-MeC6H4SO2Na

50
50
25

6
6
8

89
82
74

(0.5 mol%)

R1

R2

Yield (%)

H
OMe
CO2Me

C(CH3)2OH
C(CH3)2OH
C(CH3)2OH

92
96
95

Jin and co-workers introduced the strategy of thermoregulated phase-transfer catalysis in order to solve the mass-transfer
limitations associated with aqueous biphase catalysis.227,228 They initially investigated the inverse temperature-dependent
solubility of poly(ethylene oxide)-substituted triphenylphosphines (PETPPs) 29b (Figure 6) in the hydroformylation of
long-chain alkenes in the aqueous biphase system.229 Before the reaction, the catalyst resides in the aqueous phase while
the substrate is in the organic phase, but at high temperatures above the critical-temperature Cp (5295  C), the catalyst is
transferred into the organic phase enabling homogeneous catalysis to occur. On cooling the reaction mixture, the catalyst
becomes water-soluble again, and the aqueous catalyst solution can be readily separated from the products, recycled, and
reused. The catalyst activity, however, was only retained for the first four recycles, because the solubility of the catalyst in
the organic phase was increased by the presence of the polar aldehyde product. More recently, the group introduced the
water-soluble phosphine, N,N-dipolyoxyethylene-substituted2-(diphenylphosphino)phenylamine (PEODPPPA) 30,
and demonstrated that the high reactivity of the RhPEODPPPA complex in the aqueous-organic biphasic hydroformylation of 1-decene can be maintained over 20 catalyst recycles.230
In order to broaden the scope of the thermoregulated phase-transfer catalysis principle, the hydrogenation of styrene
using an [Ru3(CO)12]/PETPP complex in toluene was investigated (Equation (10)).231 In this system, the catalyst is
insoluble in toluene at room temperature and the organic phase remains colorless. On heating above the criticaltemperature Cp, the catalyst became soluble, enabling the reaction to proceed homogeneously, and, on cooling to room
temperature, the catalyst precipitated back out of solution as a viscous membrane at the bottom of the reactor. The
products were conveniently removed by syringe and the catalyst was recycled 10 times without loss of activity.

851

852

Green Organometallic Chemistry

10

12.17.3.6 Solid-supported Organometallic Systems


Any review on green organometallic chemistry would not be complete without reference to the heterogenization of
homogeneous catalysts on solid supports for recycle, recovery, and reuse, particularly since, despite the wealth of
research on homogeneously catalyzed reactions that occur in high yields and selectivities under mild conditions, the
industrial use of homogeneously catalyzed processes is relatively limited.232 However, it would require a whole
volume of this compilation to do justice to the exponential expansion of work in this area in the last 10 years:
Organometallic complexes from across the periodic table interacting via covalent bonds, adsorption, ion-pair formation, encapsulation or entrapment with support materials such as soluble and insoluble organic and inorganic
polymers, dendrimers, mesoporous and microporous inorganic solids, including glasses, ceramics, zeolites, and
clays, possibly with a thin film of a solvent (water, organic, fluorous, ionic liquid), and solgels. This area has been
extensively reviewed, and the reader is directed to these authoritative articles.18,233252

12.17.4 Hydroformylation: A Green Case Study I


Metal-catalyzed, homogeneous hydroformylation of alkenes, an archetypal atom-economic addition reaction, has
been extensively investigated and can be used as an exemplar of green organometallic chemistry in practice.
Originally exploited commercially in the 1950s using cobalt carbonyl catalysts,253 enhanced hydrogenation to alcohols
coupled with improved hydroformylation selectivity have been achieved following the incorporation of phosphine
ligands,254 and both of these technologies are still in use for the hydroformylation of long-chain alkenes. More
selective, rhodiumtriphenylphosphine catalysts, operating at lower temperatures and pressures, have replaced the
cobalt-based catalysts in the hydroformylation of propene in continuous processes operating at 3.5 million tonnes per
annum.255 Here, commercialization of these homogeneous processes relies upon the straightforward isolation of
either the volatile organic product (n-butanal) or the catalyst with thermally stable co-ligands, (cobalt carbonyls).
Further commercialization of the homogeneous hydroformylation reaction for product/catalyst combinations that do
not satisfy these limited criteria hinges upon new technologies, and four major green strategies have been developed
and investigated toward this goal.

12.17.4.1 Hydroformylation of Olefins on Supported Catalysts


Many systems using rhodium hydroformylation catalysts chemically bound to either insoluble (e.g., inorganic oxides
or polymers)20 or soluble (e.g., dendrimers250 or polymers224,244) supports have been reported. The high physical
strength and chemical inertness of inorganic materials make them particularly suited to act as heterogeneous catalyst
support materials. The principal problems are their non-uniform and partly unknown structures, mass-transport
limitations due to slow diffusion, and metal leaching. However, acceptable catalyst leaching (<100 ppb rhodium)
coupled with reasonable activity and excellent regioselectivity have been demonstrated for a solgel solution
incorporating a triethoxysilyl-functionalized Xantphos-type ligand 31256 (Figure 7). The mass-transport limitations
of insoluble supported catalysts can be overcome by using soluble supports, which lead to recyclable catalyst systems
with activities similar to their monomeric analogs. Dendrimers, as solid supports, have recently received considerable
attention since their well-defined macromolecular structures enable catalyst structures to be accurately and precisely
controlled. Tulchinsky and Miller have patented rhodium hydroformylation catalysts on phosphite-functionalized
polyamidoamine (PAMAM) dendrimers 32.257 Following application in the hydroformylation of propene (good
reactivities and poor regioselectivities), nanofiltration afforded excellent rhodium and dendrimer retention, and the
catalyst could be recycled with virtually no drop in activity.

Green Organometallic Chemistry

Figure 7 Representative ligand systems used in hydroformylation catalysis.

12.17.4.2 Hydroformylation of Olefins in scCO2


Although organometallic complexes, particularly rhodium-based hydroformylation catalysts incorporating phosphine
ligands, are virtually insoluble in apolar scCO2, solubility can be enhanced by the incorporation of long perfluoroalkyl
groups in the ligands 33 and 34. Using these derivatized catalysts in the hydroformylation of 1-octene, reaction rates
and regioselectivities in scCO2 comparable to those observed in conventional media have been reported,90,258 while
significantly enhanced reaction rates have been reported for the rhodium-catalyzed hydroformylation of acrylic esters
in scCO2.259,260 However, although removal of the solvent by decompression back to the gaseous state is attractive,
this does not per se deal with the issue of product/catalyst separation. The combination of scCO2 with an ionic liquid,
35a and 35b,93 or a supported catalyst,261 are potentially powerful alternative approaches to catalyst/product
separation, but in both cases, the reported hydroformylation reaction rates are relatively low.
An alternative twist exemplified in a recent report uses scCO2 not as the solvent for the reaction, but rather the
antisolvent for post-reaction catalyst precipitation.262 The rhodium catalyst, which incorporates a

853

854

Green Organometallic Chemistry

poly(ethyleneglycol) (PEG)-modified phosphine ligand 29a for hydroformylation in neat 1-octene, precipitates on
pressurization with CO2, allowing the product to be removed by sc extraction and the catalyst to remain in situ for
subsequent catalytic cycles.

12.17.4.3 Hydroformylation of Olefins in Biphase Systems


Alkene hydroformylation has been evaluated in three types of biphasic system where the catalyst is designed to be
soluble in a solvent that, under some conditions, is immiscible with that containing the product.

12.17.4.3.1

Hydroformylation of olefins in aqueous media

The thermal instability of rhodium-based hydroformylation catalysts has already been overcome commercially in the

Ruhrchemie/RhonePoulenc
process for propene hydroformylation in which the sodium salt of a sulfonated triphenylphosphine ligand (TPPTS, 1a) is used to solubilize the catalyst in the aqueous phase. In this process, the second
phase is toluene and the reaction is carried out as a batch process with rapid stirring to intimately mix the two
immiscible phases. After reaction, the system is allowed to separate and the organic phase is simply decanted from the
aqueous catalyst phase.263 Both water-soluble polymers264,265 and PAMAM dendrimers32,266 have been reported as
supports for rhodium-catalyzed hydroformylation under aqueous biphase conditions, but reactivities and regioselectivities were only comparable to or worse than those obtained with the reference TPPTS ligand. The aqueous
biphase approach has found limited application for the hydroformylation of longer-chain alkenes, because of their
very low solubility in water leading to prohibitively slow reaction rates, but there have been a variety of approaches
directed at the solution of this problem.
(i) Amphiphilic phosphines,63 particularly those based on the Xantphos ligand 36,267 offer high activities and
regioselectivities for hydroformylation under aqueous biphase conditions. Alternatively, cationic surfactants dramatically increase the rate and regioselectivity for hydroformylation of 1-octene by the RhTPPTS catalyst system as a
result of micelle formation.62
(ii) Catalytic amounts of water-soluble receptors, such as cyclodextrins or calixarenes, can be used to promote the
solubilization of water-insoluble substrates in the aqueous phase in the presence of the water-soluble organometallic
catalyst. Partially methylated -cyclodextrins with rhodium complexes of TPPTS or TPPMS 1b have given good
reactivities and modest regioselectivities in the hydroformylation of 1-decene under aqueous biphase conditions, with
<0.5 ppm rhodium leaching to the product phase.268 Alternatively, bonding a bidentate phosphine unit to the
cyclodextrin via a spacer unit seeks to combine aqueous biphase organometallic catalysis, phase-transfer catalysis,
and molecular recognition. Unfortunately, although activities and regioselectivities are high, the catalytic systems
offered modest recycle (50% activity on the second run) due to some loss of the catalyst to the product phase.269 In
contrast, catalyst recycle following hydroformylation was excellent for phosphine-modified, sulfonated calixarenes,
but the chemoselectivity was poor with yields of the aldehyde product ranging from 52% to 86%.270 Most recently, it
has been reported that a functionalized cationic -cyclodextrin is an efficient mass-transfer promoter affording both
enhanced reactivity and regioselectivity in the rhodium-catalyzed hydroformylation of 1-decene, possibly via
the in situ formation of a new catalytic supramolecular species by ion exchange between the ligand and the
-cyclodextrin.271
(iii) In thermoregulated phase-transfer catalysis, replacement of the sulfonated ligand with phosphines incorporating PEG or other polyoxyethylene ponytails 30 generates rhodiumphosphine catalysts that are exclusively soluble
in water at room temperature, but when the ponytail undergoes a phase transition on heating, the catalysts become
preferentially soluble in the organic phase: that is, at process temperature, the catalyst acts homogeneously, while on
cooling, reversal of the phase transition returns the catalyst to the aqueous phase for separation by decantation. The
results are promising, with systems recycled up to 20 times, excellent chemoselectivity but poor regioselectivity
(linear : branched ca. 0.6).228,229,272274

12.17.4.3.2

Hydroformylation of olefins in fluorous media

Hydroformylation of long-chain alkenes under fluorous biphasic conditions, in which perfluoroalkylated phosphine/
phosphite ligands are used to solubilize the catalyst in the fluorous phase, have been extensively investigated by a
number of groups.275278 Good reaction rates, regioselectivity, catalyst stability, and rhodium leaching levels have all
been reported, but the latter can be improved substantially by omitting the organic phase. While 1-octene is fully
miscible with the fluorous phase, the hydroformylation product, 1-nonanal, is completely immiscible, forming a

Green Organometallic Chemistry

second product phase as the reaction proceeds; rhodium-leaching levels were reported as 80 ppb alongside good
regioselectivity (linear : branched 6.3). This concept has been incorporated into a continuous reactor using 33
where the fluorous catalyst underwent 15,500 turnovers at an average rate of 750 h1.279

12.17.4.3.3

Hydroformylation of olefins in ionic liquids

Chauvin et al. investigated the biphasic hydroformylation of pent-1-ene with the neutral [Rh(CO)2(acac)]/triarylphosphine as the catalyst precursor in [BMIM][PF6].138 Although high activities were observed, slight leaching of the
catalyst into the organic phase occurred. Although the use of 1b was able to suppress this completely by making the
catalyst more soluble in the ionic liquid, the activity of the system was significantly reduced. Four groups, using
different approaches, have delivered substantial improvements over this early work. Salzer and Brasse have developed and synthesized a cobaltocenium ligand280 and showed the benefits of using ligand systems that are specifically
designed for this application. Not only did they observe high activity for longer chain -olefins such as 1-octene and
good regioselectivity for the linear aldehyde, but also the rhodium leaching into the organic layer was less than 0.5%.
Dupont et al., Wasserscheid and co-workers, and van Leeuwen and co-workers have independently reported highly
active and regioselective hydroformylation of 1-octene with different Xantphos ligands 3638 in ionic liquids, with
excellent regioselectivities, very low rhodium leaching levels, and facile product separation by straightforward
decantation.281283 Jin and co-workers have reported the synthesis and application of a novel PEG-functionalized
tetraalkylammonium tosylate room-temperature ionic liquid in combination with either 1a or their
octylpolyethyleneglycolphenylenephosphite 39 for the hydroformylation of 1-tetradecene.284 Following an induction period (20 h), good conversions, but with poor regioselectivities, over three recycles in an ionic liquid : n-heptane
biphase were obtained with as little as 0.5% rhodium leaching.

12.17.4.4 Hydroformylation of Olefins in Supported Liquid-phase Catalysis


Supported liquid-phase catalysis,167,265 in which the catalyst is dissolved in a small volume of solvent, adsorbed on,
usually, a hydrophilic solid, seeks to resolve issues associated with substrate solubility in multi-phase catalysis and
performance/catalyst leaching in supported catalysis; reports on the hydroformylation of long-chain alkenes under
both supported aqueous phase and supported ionic liquid-phase regimes have been reported.

12.17.4.4.1

Hydroformylation of olefins in supported aqueous phase catalysis

In the hydroformylation of alkenes using RhTPPTS catalysts in a thin layer of water adsorbed on silica, TOFs are
independent of alkene chain length as a direct consequence of the large interfacial surface, indicating a major advance
over hydroformylation under aqueous biphase conditions.285 However, catalyst-support stability, regioselectivity, and
levels of rhodium leaching are still unacceptable. These issues have been overcome by van Leeuwen et al. using their
water-soluble sulfonated Xantphos ligand 36, where excellent regioselectivities combined with multiple recycles
without loss of activity and <1 ppm rhodium leaching levels have been reported, but reaction rates are disappointing.286 Alternatively, Naughton and Drago describe the hydroformylation of 1-hexene and 1-octene using the Rh
TPPTS catalyst dissolved in a hydrophilic, PEG film.287 Activities and regioselectivities comparable to those in a
homogeneous system are obtained, although the amount of rhodium leaching was not measured.

12.17.4.4.2

Hydroformylation of olefins in supported ionic liquid-phase catalysis

Rhodium35a and rhodium35b catalysts in [BMIM][PF6] supported on a silylimidazolium-modified silica have


been shown to have activities and regioselectivities comparable to those under ionic liquid-biphase conditions in the
hydroformylation of 1-hexene. Unfortunately, rhodium leaching levels were quite high, possibly as a consequence of
the solubility of the ionic liquid in the organic phase at high aldehyde concentrations.288,289 In contrast, negligible
amounts of rhodium were detected in the organic phase following the hydroformylation of either propene or 1-octene
using either the sulfonated Xantphos 36 or bis(m-phenylguanidinium)phenylphosphine 37 in [BMIM][PF6] in a
continuous fixed-bed reactor.290,291

12.17.4.5 Overview
In many cases, it is not possible to make direct comparisons between catalytic results from different research groups
as a result of differences in approach and reporting. However, Table 1 summarizes key catalytic data from some of

855

856

Green Organometallic Chemistry

Table 1 Rhodium-catalyzed hydroformylation of alkenesa

System

Substrate

Ligand

Pressure
(bar)

Homogeneous
Supported
Supported
scCO2
scCO2/ionic liquid
scCO2/supported
Aqueous biphase
Aqueous biphase
Aqueous/supported
Thermoregulated
Fluorous biphase
Ionic liquid
Ionic liquid
Ionic liquid
Ionic liquid/supported

1-Octene
1-Octene
Propene
1-Octene
1-Octene
1-Octene
Propene
1-Decene
1-Octene
1-Octene
1-Octene
1-Octene
1-Octene
1-Tetradecene
Propene

PPh3
31
32
34
35b
31
1a
1a
36
29b
33
37
38
1a
37

15
50
7
200
200
170
50
50
50
50
20
30
46
5
5

Temp.
( C)

TOF
(h1)

Rate (mol
dm3 h1)

95
80
70
65
100
90
120
80
80
100
70
100
100
105
120

770
287
n.r.
430
8
160
400
n.r.
15
182
4400
50
318
94
21

2.0
0.19
1.8
12.2
0.12
n.r.
1.1
n.r.
n.r.
n.r.
8.8
n.r.
1.2
n.r.
n.r.

n:i

Rh loss
(mg per
mol product)

References

8.8 : 1
40 : 1
1.4 : 1
5.5 : 1
3.1 : 1
33 : 1
19 : 1
1.9 : 1
46 : 1
n.r.
6.3 : 1
21 : 1
49 : 1
0.4 : 1
1:1

n.a.
<0.1
<0.1%b
<0.17
<0.06%b
<1.2
<0.005
<0.5 ppm
<1 ppm
n.r.b
0.08
<0.07%b
<0.005
<0.5%b
<0.7%

255
256
257
90
93
261
263
268
286
229
277
282
283
284
291

n.a. not applicable; n.r. not reported.


Reported as % of catalyst loading.

these methodologies, offering an insight into their relative values from a green chemistry perspective. The results
from the bidentate Xantphos-type ligands in SAPC generally offer the most promising results; excellent regioselectivity to the desired linear aldehydes with very low rhodium leaching levels. However, even at elevated temperatures,
the reaction rates and TOFs for these systems offer some room for improvement.

12.17.5 Olefin Polymerization: A Green Case Study II


Over the last two decades, organometallic complexes have been at the heart of many of the key advances in metalmediated alkene polymerization technology, with many examples now reaching the early stages of commercialization. While early transition metal complexes (e.g., metallocenes, constrained-geometry catalysts) have led the way,
the advent of late transition metal catalysts has presented a rich library of highly active systems that can be employed
to polymerize a variety of olefins. Comprehensive reviews covering the chronological development of both early292295
and late transition metal systems can be found elsewhere.296299 In this section, we highlight some of the key
advances that have occurred with regard to their application in alternative reaction media, aqueous, scCO2, fluorous,
and ionic liquids, for ethylene or propylene polymerization.

12.17.5.1 Metal-mediated Polymerizations of Olefins in Aqueous Media


Despite the industrial importance of free-radical-initiated emulsion polymerization,300 the use of water as a medium
for transition metal-mediated coordination polymerization of olefinic monomers has received relatively little attention.50,301310 This can be mainly attributed to the water sensitivity of commercially important early transition metal
catalysts in which hydrolysis of a metal alkyl bond can readily occur.292295 Otherwise, the water molecule can act as a
good ligand and fill the vacant site, or the water can attack coordinated monomers or other ligands. With the advent of
late transition metal catalysts and their expected tolerance for polar groups, the likelihood for aqueous polymerizations has started to show great potential. The following section highlights reports of transition metal catalysts that
have been employed either as aqueous soluble or as suspensions. General reviews on the use of organometallic
complexes as catalysts in aqueous media have been presented elsewhere,1618 while two reviews specific to aqueous
alkene polymerizations have appeared recently.300,301
In 1993, Flood and co-workers reported that the well-defined water-soluble monocationic rhodium complex
[(N^N^N)RhMe(OH2)(OH)] 40 (N^N^N trimethyl1,4,7-triazacyclononane) produced low molecular weight
polyethylene at room temperature and 60 bar ethylene after extended reaction times (ca. 90 days).302 Increasing

Green Organometallic Chemistry

Figure 8 Representative catalyst or precatalyst systems for olefin polymerization.

the temperature only led to hydrolysis of the RhC bond. Despite the low activity of this system, it signposted the
possibilities for polymerization catalysis in aqueous media (Figure 8).
More recently, the sodium salt of 41 has been reported and found to be soluble in water (stable for several hours),
and can promote the aqueous polymerization of ethylene producing low molecular weight linear polyethylene with
103 turnovers h1 under moderate conditions (70  C, 50 bar ethylene pressure).303305 A variety of other nickel
catalysts containing variations of the P^O ligand in 41 have been reported,306 including the fluorinated derivative
[Ph2PC(COOEt)C(RF)ONiPh] (RF CF3, C6F5; generated in situ), which produces low molecular weight linear
polyethylene with much higher activities in aqueous emulsion.307 In general, the catalyst activities and polymer
molecular weight are reduced in the aqueous polymerizations when compared with polymerizations in organic media
such as toluene. This observation has been attributed to insufficient local concentrations at the active centers leading
to a lower rate of chain growth.
As an alternative strategy, Mecking and co-workers reported the use of an -diimine Pd(II) complex of type 42 in a
suspension-type process where the catalyst is not solubilized since no organic solvent is added to the aqueous
phase.50,303 Amazingly, the activities observed at pressures above 20 bar are similar to the highest ones obtained
for polymerization in a soluble organic medium [900 vs. 1,100 (in CH2Cl2) mol molPd1 h1]. Notably, solubilization
of 42 by introducing sulfonate groups results in an inactive ethylene polymerization catalyst,308 and it has been
concluded that the suspension-type system operates by encapsulation of the water-insoluble catalyst in the

857

858

Green Organometallic Chemistry

hydrophobic polymer. Suspension-type polymerization also occurs using the neutral salicylaldiminatenickel systems
43, and, in contrast to the P^O systems discussed above, much higher molecular weight polymers are afforded with
moderate amounts of methyl branches evident.

12.17.5.2 Metal-mediated Polymerizations of Olefins in scCO2


In spite of the numerous reports of metallocene-mediated polymerizations in sc media such as ethylene and
propylene,311313 the use of carbon dioxide as the sc medium is considerably more scarce. The unique properties
of scCO2 as the reaction phase, such as its nontoxic and nonflammable nature along with it being inexpensive, make it
an attractive medium to carry out metal-mediated polymerizations.314316 With regard to olefin polymerization, the
paucity of examples can be attributed, in some measure, to the oxophilicity (and hence reactivity) of traditional early
transition metal catalysts. However, late transition metal catalysts for olefin polymerization are less oxophilic295298
and can, therefore, be used in scCO2. Using the -diimine Pd(II) catalysts 44,296 ethylene can be polymerized in
scCO2, affording high molecular weight polyethylene; use of 1-hexene as the monomer also forms high molecular
weight polymer.317319 The polyethylene generated is highly branched and similar to that generated in an organic
solvent (e.g., CH2Cl2) even though the polymerization in scCO2 is a precipitation polymerization. In addition, some
evidence for a new branch-on-branch structure is apparent for the polyethylene generated during this scCO2
polymerization.319

12.17.5.3 Metal-mediated Polymerizations of Olefins in Fluorous Media


Despite the interesting solvent properties exhibited by perfluorocarbon (PFC) solvents and their close relationship to
scCO2,113,319,320 examples of coordination polymerization in PFC solvents are relatively scarce.321328 The very low
catalyst loadings that can be employed in some of the more active polymerization catalysts mean that catalyst
recovery issues are not of major importance in coordination polymerization. Nevertheless, inspired by what the
potential effects a fluorous reaction medium (pure PFC solvent or FBS conditions) and, in some cases, a highly
fluorinated catalyst would have on the physical properties of the polymer, some work has started to appear in both the
academic and industrial literature.324328
For example, the zirconocene complexes 45 are hydrocarbon soluble and moderately soluble in PFC solvents. All
three complexes form active ethylene polymerization catalysts when reacted with an excess of methylaluminoxane
(MAO; Al/Zr 500/1) either in pure toluene or in fluorous biphasic solvent systems (e.g., toluene/perfluorohexanes).324,325 Increased robustness and productivity over prolonged polymerization times in comparison to the nonfluorous reference system [Zr(5-C5H4SiMe3)2Cl2] are observed in each case. This phenomenon has been ascribed to
phase segregation of the fluorous metallocene moiety from the polyethylene matrix. Remarkably, fluorous biphasic
conditions during the polymerization reaction, in general, result in higher molecular weight polymer. This phenomenon was even observed for non-fluorous [Zr(5-C5H4SiMe3)2Cl2] and is most likely caused by a suppression of the
rate of -H transfer to the metal center in these relatively non-polar solvent combinations.

12.17.5.4 Metal-mediated Polymerizations of Olefins in Ionic Liquids


Ambient-temperature ionic liquids have received much attention in both academia and industry due to their potential
as replacements for volatile organic compounds (VOCs).134,329 Recent years have witnessed an increased amount of
interest in using an ionic liquid as the medium for alkene oligomerization and polymerization.330342 The first
example of a polymerization system in an ionic liquid was reported by Carlin and Wilkes.330 Titanocene 46 in an
ionic liquid composed of acidic 1-ethyl3-methylimidazolium chloride/AlCl3 (EMIC), using AlCl3xRx (R Me, Et)
as co-catalyst, was used to polymerize ethylene with a catalytic activity up to 0.023 g of ethylene per min.
Interestingly, neither [Zr(5-Cp)2Cl2] nor [Hf(5-Cp)2Cl2] were catalytically active in the ionic liquids employed.
No recycling experiments using 46 were, however, reported.
Late transition metals have a lower oxophilicity relative to early transition metals and, therefore, a higher tolerance
for a wide range of functional groups (e.g., COOR and COOH groups).295298 Consequently, the use of late
transition metal systems for alkene polymerization in ionic liquids has been disclosed in both the academic and patent
literature.331334 For example, de Souza and co-workers disclosed the use of the nickel complex 47 for the homopolymerization of ethylene in an ambient-temperature ionic liquid [BMIM][ClAlCl3EtAlCl2] and toluene.331 The
polyethylene formed was easily isolated from the reaction mixture by decanting the upper toluene layer allowing the

Green Organometallic Chemistry

ionic liquid and 47 to be recycled for use in further polymerizations. Nevertheless, to allow reuse, trimethylaluminum
was added to overcome the loss of free alkylaluminum species into the separated organic phase. Notably, the
performance of the catalyst changes on reuse with subsequent reactions, giving a progressive shift from crystalline
to amorphous polymer, with a period that gives rise to bimodal product distributions. This change in performance has
been attributed to the changing composition of the ionic liquid with different active species being generated on
addition of trimethylaluminum.

12.17.6 Future Perspectives


The increasing legislative drivers to phase out chlorinated and carcinogenic solvents means the use of alternative
reaction media is bound to see considerable advances over the next decade. This review shows that organometallic
catalysts have now been developed for a wide range of reactions in solvents from non-polar scCO2 through to the
highly polar water and ionic liquids. Similarly, organometallic catalysis has been used as a test bed for new enhanced
reaction technologies. Over the next decade, organometallic chemistry will continue to play an increasing crucial role
in the development of more atom-efficient and selective reactions and, hence, green chemistry.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.

Anastas, P. C.; Warner, J. C. Green Chemistry Theory and Practice; Oxford University Press: New York, 1998.
Crystal www.crystalfaraday.org.
US Environmental Protection Agency www.epa.gov.
Chemistry.org www.chemistry.org.
Chemsoc www.chemsoc.org.
Nelson, W. M. Green Solvents for Chemistry; Oxford University Press: Oxford, 2003.
Lancaster, M. Green Chemistry: An Introductory Text; RSC: Cambridge, 2002.
Clark, J. H.; MacQuarrie, D. J. Handbook of Green Chemistry and Technology; Blackwell: Oxford, 2002.
Ahluwalia, V. K.; Kidwai, M. New Trends in Green Chemistry; Kluwer: Dordrecht, 2004.
Matlack, A. S. Introduction to Green Chemistry; Dekker: New York, 2001.
Adams, D. J.; Dyson, P. J.; Tavener, S. J. Chemistry in Alternative Reaction Media; Wiley: Chichester, 2004.
Jessop, P. G.; Leitner, W. Chemical Synthesis using Supercritical Fluids; Wiley-VCH: Weinheim, Germany, 1999.
Gladysz, J. A., Curran, D. P., Horvath, I. T., Eds. Handbook of Fluorous Chemistry; Wiley-VCH: Weinheim, Germany, 2004.
Wasserscheid, P., Welton, T., Eds. Ionic Liquids in Synthesis; Wiley-VCH: Weinheim, Germany, 2003.
Cornils, B., Herrmann, W. A., Eds. Applied Homogeneous Catalysis with Organometallic Compounds, 2nd ed.; Wiley-CH: Weinheim, Germany, 2002.
Grieco, P. A. Organic Synthesis in Water; Thomson Science: London, 1998.
Joo, F. Aqueous Organometallic Catalysis; Kluwer: Dordrecht, 2001.
Cornils, B.; Herrmann, W. E. Aqueous-Phase Organometallic Catalysis, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2004.
Cornils, B., Herrmann, W. E., Horvath, I. T., Leitner, W., Mecking, S., Olivier-Bourbigou, H., Vogt, D., Eds. Multiphase Homogeneous
Catalysis; Wiley-VCH: Weinheim, Germany, 2005.
Hartley, R. R. Supported Metal Complexes; Reidel: Dordrecht, 1985.
Tooze, B.; Cole-Hamilton, D. J. Recovery and Recycling in Homogeneous Catalysis; Kluwer: Dordrecht, 2005.
Wang, Y.; Holladay, J. Microreactor Technology and Process Intensification; American Chemical Society: Washington, DC, 2005.
DeSimone, J. M.; Tumas, W. Green Chemistry using Liquid and Supercritical Carbon Dioxide; Oxford University Press: New York, 2003.
Udo de Haes, H. A. J. Clean. Prod. 1993, 1, 131137.
Jodicke, G.; Zenklusen, O.; Weidenhaupt, A.; Hungerbuhler, K. J. Clean. Prod. 1999, 7, 159166.
Hudlicky, T.; Frey, D. A.; Koroniak, L.; Claeboe, C. D.; Brammer, L. E. Green Chem. 1999, 1, 5759.
Sheldon, R. A. Chem. Ind. (London) 1992, 903906.
Trost, B. M. Science 1991, 254, 14711477.
Curzons, A. D.; Constable, D. J. C.; Mortimer, D. N.; Cunningham, V. L. Green Chem. 2001, 3, 16.
Constable, D. J. C.; Curzons, A. D.; Cunningham, V. L. Green Chem. 2002, 4, 521527.
Lapkin, A.; Joyce, L.; Crittenden, B. Environ. Sci. Technol. 2004, 38, 58155823.
Trost, B. M.; Frederiksen, M. U.; Rudd, M. T. Angew. Chem., Int. Ed. 2005, 44, 66306666.
Trost, B. M. Acc. Chem. Res. 2002, 35, 695705.
Trost, B. M.; Toste, F. D.; Pinkerton, A. B. Chem. Rev. 2001, 101, 20672096.
Wender, P. A.; Bi, F. C.; Gamber, G. G.; Gosselin, F.; Hubbard, R. D.; Scanio, M. J. C.; Sun, R.; Williams, T. J.; Zhang, L. Pure Appl. Chem.
2002, 74, 2531.
Wender, P. A.; Baryza, J. L.; Brenner, S. E.; Clarke, M. O.; Gamber, C. G.; Horan, J. C.; Jessop, T. C.; Kan, C.; Pattabiraman, K.; Williams, T. J.
Pure Appl. Chem. 2003, 75, 143155.
Crochet, P.; Dez, J.; Fernandez-Zumel, M. A.; Gimeno, J. Adv. Synth. Catal. 2006, 348, 93100.
Murai, S.; Kakiuchi, F.; Sekine, S.; Tanaka, Y.; Kamatani, A.; Sonoda, M.; Chatani, N. Nature 1993, 366, 529531.
Kakiuchi, F.; Murai, S. Acc. Chem. Res. 2002, 35, 826834.
Kakiuchi, F.; Chatani, N. Adv. Synth. Catal. 2003, 345, 10771101.
Herrmann, W. A.; Kohlpaintner, C. W. Angew. Chem., Int. Ed. 1993, 32, 15241544.
Papadogianakis, G.; Sheldon, R. A. New J. Chem. 1996, 20, 175185.

859

860

Green Organometallic Chemistry

43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.

Lindstrom, U. M. Chem. Rev. 2002, 102, 27512772.


Joo, F.; Katho, A. J. Mol. Catal. A: Chem. 1997, 116, 326.
Dwars, T.; Oehme, G. Adv. Synth. Catal. 2002, 344, 239260.
Li, C. J. Acc. Chem. Res. 2002, 35, 533538.
Amengual, R.; Genin, E.; Michelet, V.; Savignac, M.; Genet, J.-P. Adv. Synth. Catal. 2002, 344, 393398.
Venkatraman, S.; Huang, T.; Li, C. J. Adv. Synth. Catal. 2002, 344, 399405.
Verspui, G.; Feiken, J.; Papadogianakis, G.; Sheldon, R. A. J. Mol. Catal. A: Chem. 1999, 146, 299307.
Held, A.; Mecking, S. Chem. Eur. J. 2000, 6, 46234629.
Lynn, D. M.; Mohr, B.; Grubbs, R. H. J. Am. Chem. Soc. 1998, 120, 16271628.
Breslow, R. Acc. Chem. Res. 1991, 24, 159164.
Furlani, T. R.; Gao, J. J. Org. Chem. 1996, 61, 54925497.
Narayan, S.; Muldoon, J.; Finn, M. G.; Fokin, V. V.; Kolb, H. C.; Sharpless, K. B. Angew. Chem., Int. Ed. 2005, 44, 32753279.
Pinault, N.; Bruce, D. W. Coord. Chem. Rev. 2003, 241, 125.
Kuntz, E., Fr. Pat. 2.314.910, 1975.
Kuntz, E. G. Chem. Tech. 1987, 17, 570575.
Buhling, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Mol. Catal. A: Chem. 1995, 98, 6980.
Buhling, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Elgersma, J. W. J. Mol. Catal. A: Chem. 1997, 116, 297308.
Grassert, I.; Vill, V.; Oehme, G. J. Mol. Catal. A: Chem. 1997, 116, 231236.
Grassert, I.; Schmidt, U.; Ziegler, S.; Fischer, C.; Oehme, G. Tetrahedron: Asymmetry 1998, 9, 41934202.
Chen, H.; Li, Y.; Chen, J.; Cheng, P.; He, Y.-E.; Li, X. J. Mol. Catal. A: Chem. 1999, 149, 16.
Hanson, B. E. Coord. Chem. Rev. 1999, 185186, 795807.
Uozumi, Y.; Nakazono, M. Adv. Synth. Catal. 2002, 344, 274277.
Thomas, C. M.; Ward, T. R. Chem. Soc. Rev. 2005, 34, 337346.
Bertucci, C.; Botteghi, C.; Giunta, D.; Marchetti, M.; Paganelli, S. Adv. Synth. Catal. 2002, 344, 556562.
Ward, T. R. Chem. Eur. J. 2005, 11, 37983804.
Li, C. J.; Chen, L. Chem. Soc. Rev. 2006, 35, 6882.
Li, C. J. Green Chem. 2002, 4, 14.
Li, C.-J. Tetrahedron 1996, 52, 56435668.
Sakai, M.; Hayashi, H.; Miyaura, N. Organometallics 1997, 16, 42294231.
Sakai, M.; Ueda, M.; Miyaura, N. Angew. Chem., Int. Ed. 1998, 37, 32793281.
Huang, X.; Anderson, K. W.; Zim, D.; Jiang, L.; Klapars, A.; Buchwald, S. L. J. Am. Chem. Soc. 2003, 125, 66536655.
Leadbeater, N. E. Chem. Commun. 2005, 28812902.
Wei, C. M.; Li, C. J. Green Chem. 2002, 4, 3941.
Wei, C.; Li, C.-J. J. Am. Chem. Soc. 2003, 125, 95849585.
McHugh, M. A.; Krukonis, V. J. Supercritical Fluid Extraction; Butterworth Heinemann: Boston, 1994.
Reverchon, E. J. Supercrit. Fluid 1999, 15, 121.
DeSimone, J. M.; Maury, E. E.; Menceloglu, Y. Z.; Combes, J. R.; McClain, J. B.; Romack, T. Science 1994, 265, 356359.
Licence, P; Ke, J.; Sokolova, M.; Ross, S. K.; Poliakoff, M. Green Chem. 2003, 5, 99104.
Jessop, P. G.; Ikariya, T.; Noyori, R. Chem. Rev. 1999, 99, 475493.
Leitner, W. Acc. Chem. Res. 2002, 35, 746756.
Campestrini, S.; Tonellato, U. Curr. Org. Chem. 2005, 9, 3147.
Prajapati, D.; Gohain, M. Tetrahedron 2004, 60, 815833.
Francio, G.; Leitner, W. Transition Metals for Organic Synthesis, 2nd ed.; Wiley-VCH: Weinheim, 2004; pp 545558.
Jessop, P. G. In Supercritical Fluid Technology for Drug Development: Vol. 138; York, R., Kompella, U. B., Shekunov, B. Y., Eds.; Marcel
Dekker, Inc: New York, 2003, pp 461495.
Alonso, F.; Beletskaya, I. P.; Yus, M. Tetrahedron 2005, 61, 1177111835.
Jessop, P. G. Top. Catal. 1998, 5, 95103.
Sellin, M. F.; Bach, I.; Webster, J. M. J. Chem. Soc., Dalton Trans. 2002, 45694576.
Koch, D.; Leitner, W. J. Am. Chem. Soc. 1998, 120, 1339813404.
Abbott, A. P.; Eltringham, W.; Hope, E. G.; Nicola, M. Green Chem. 2005, 7, 721725.
Kainz, S.; Brinkmann, A.; Leitner, W.; Pfaltz, A. J. Am. Chem. Soc. 1999, 121, 64216429.
Goetheer, E. L. V.; Verkerk, A. W.; van den Broeke, L. J. P.; de Wolf, E.; Deelman, B-J.; van Koten, G.; Keurentjes, J. T. F. J. Catal. 2003,
219, 126133.
Sellin, M. F.; Webb, P. B.; Cole-Hamilton, D. J. Chem. Commun. 2001, 781782.
Blanchard, L. A.; Hancu, D.; Beckman, E. J.; Brennecke, J. F. Nature 1999, 399, 2829.
Blanchard, L. A.; Brennecke, J. F. Ind. Eng. Chem. Res. 2001, 40, 287292.
Leitner, W. Pure Appl. Chem. 2004, 76, 635644.
Sun, Y.; LeBlond, C.; Wang, J.; Blackmond, D. G. J. Am. Chem. Soc. 1995, 117, 1264712648.
Jessop, P. G.; Ikariya, T.; Noyori, R. Nature 1994, 368, 231233.
Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1996, 118, 344355.
Jessop, P. G.; Ikariya, T.; Noyori, R. Organometallics 1995, 14, 15101513.
Naesnik, T. E.; Freudenberger, J. H.; Orchin, M. J. Organomet. Chem. 1982, 236, 95100.
Matsui, Y.; Orchin, M. J. Organomet. Chem. 1983, 244, 369373.
Jessop, P. G.; Hsiao, Y.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1994, 116, 88518852.
Lange, S.; Brinkmann, A.; Trautner, P.; Woelk, K.; Bargon, J.; Leitner, W. Chirality 2000, 12, 450457.
Burk, M. J.; Feng, S.; Gross, M. F.; Tumas, W. J. Am. Chem Soc. 1995, 117, 82778278.
Oakes, R. S.; Clifford, A. A.; Rayner, C. M. J. Chem. Soc., Perkin Trans. 1 2001, 917941.
Adams, D. J.; Chen, W.; Hope, E. G.; Lange, S.; Stuart, A. M.; West, A.; Xiao, J. Green Chem. 2003, 5, 118122.
Xiao, J.; Nefkens, S. C. A.; Jessop, P. G.; Ikariya, T.; Noyori, R. Tetrahedron Lett. 1996, 37, 28132826.
Hu, Y.; Birdsall, D. J.; Stuart, A. M.; Hope, E. G.; Xiao, J. J. Mol. Cat. A: Chem. 2004, 219, 5760.

Green Organometallic Chemistry

111.
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.

Horvath, I. T.; Rabai, J. Science 1994, 266, 7275.


Schneider, S.; Bannwarth, W. Helv. Chim. Acta 2001, 84, 735742.
Horvath, I. T. Acc. Chem. Res. 1998, 31, 641650.
Hope, E. G.; Stuart, A. M. J. Fluorine Chem. 1999, 100, 7583.
De Wolf, E.; Van Koten, G.; Deelman, B.-J. Chem. Soc. Rev. 1999, 28, 3741.
Cavazzini, M.; Montanari, F.; Pozzi, G.; Quici, S. J. Fluorine Chem. 1999, 94, 183193.
Fish, R. H. Chem. Eur. J. 1999, 5, 16771680.
Barthel-Rosa, L. P.; Gladysz, J. A. Coord. Chem. Rev. 1999, 190192, 587605.
Dobbs, A. P.; Kimberley, M. R. J. Fluorine Chem. 2002, 118, 317.
Curran, D. P. Synlett 2001, 9, 14881496.
Curran, D. P.; Fischer, K.; Moura-Letts, G. Synlett 2004, 13791382.
Matsugi, M.; Curran, D. P. J. Org. Chem. 2005, 70, 16361642.
Tzschucke, C. C.; Markert, C.; Glatz, H.; Bannwarth, W. Angew. Chem., Int. Ed. 2002, 41, 45004503.
Tzschucke, C. C.; Andrushko, V.; Bannwarth, W. Eur. J. Org. Chem. 2005, 52485261.
Tzschucke, C. C.; Bannwarth, W. Helv. Chim. Acta 2004, 87, 28822889.
Audic, N. A.; Bennett, J. A.; Dyer, P. W.; Hope, E. G.; Stuart, A. M.; Suhard, S. 1st International Symposium on Fluorous Technologies;
Bordeaux: France, 2005.
Pozzi, G.; Cavazzini, M.; Quici, S.; Maillard, D.; Sinou, D. J. Mol. Catal. A: Chem. 2002, 182183, 455461.
Pozzi, G.; Shepperson, I. Coord. Chem. Rev. 2003, 242, 115124.
Fache, F. New J. Chem. 2004, 28, 12771283.
Wasserscheid, P.; Keim, W. Angew. Chem., Int. Ed. 2000, 39, 37733789.
Sheldon, R. Chem. Commun. 2001, 23992407.
Olivier-Bourbigou, H.; Magna, L. J. Mol. Catal. A: Chem. 2002, 182, 419437.
Gordon, C. M. Appl. Catal. A: Gen. 2001, 222, 101117.
Dupont, J.; de Souza, R. F.; Suarez, P. A. Z. Chem. Rev. 2002, 102, 36673691.
Welton, T. Coord. Chem. Rev. 2004, 248, 24592477.
Chauvin, Y.; Gilbert, B.; Guibard, I. J. Chem. Soc., Chem. Commun. 1990, 17151716.
Suarez, P. A. Z.; Dullius, J. E. L.; Einloft, S.; de Souza, R. F.; Dupont, J. Polyhedron 1996, 15, 12171219.
Chauvin, Y.; Mussmann, L.; Olivier, H. Angew. Chem., Int. Ed. Engl. 1995, 34, 26982700.
Dullius, J. E. L.; Suarez, P. A. Z.; Einloft, S.; de Souza, R. F.; Dupont, J.; Fischer, J.; de Cian, A. Organometallics 1998, 17, 815819.
Zim, D.; de Souza, R. F.; Dupont, J.; Monteiro, A. L. Tetrahedron Lett. 1998, 39, 70717074.
Knifton, J. F. J. Mol. Catal. 1987, 43, 6577.
Song, C. E.; Roh, E. J. Chem. Commun. 2000, 837838.
Owens, G. S.; Abu-Omar, M. M. Chem. Commun. 2000, 11651166.
Buijsman, R. C.; van Vuuren, E.; Sterrenburg, J. G. Org. Lett. 2001, 3, 37853787.
Semeril, D.; Olivier-Bourbigou, H.; Bruneau, C.; Dixneuf, P. H. Chem. Commun. 2002, 146147.
Suarez, P. A. Z.; Dullius, J. E. L.; Einloft, S.; de Souza, R. F.; Dupont, J. Inorg. Chim. Acta 1997, 255, 207209.
Dyson, P. J.; Ellis, D. J.; Parker, D. G.; Welton, T. Chem. Commun. 1999, 2526.
Bonilla, R. J.; James, B. R.; Jessop, P. G. Chem. Commun. 2000, 941942.
Monteiro, A. L.; Zinn, F. K.; de Souza, R. F.; Dupont, J. Tetrahedron: Asymmetry 1997, 8, 177179.
Guernik, S.; Wolfson, A.; Herskowitz, M.; Greenspoon, N.; Geresh, S. Chem. Commun. 2001, 23142315.
Berger, A.; de Souza, R. F.; Delgado, M. R.; Dupont, J. Tetrahedron: Asymmetry 2001, 12, 18251828.
Dyson, P. J.; Laurenczy, G.; Ohlin, C. A.; Vallance, J.; Welton, T. Chem. Commun. 2003, 24182419.
Brown, R. A.; Pollet, P.; McKoon, E.; Eckert, C. A.; Liotta, C. L.; Jessop, P. G. J. Am. Chem. Soc. 2001, 123, 12541255.
Kaufmann, D. E.; Nouroozian, M.; Henze, H. Synlett 1996, 10911092.
Carmichael, A. J.; Earle, M. J.; Holbrey, J. D.; McCormac, P. B.; Seddon, K. R. Org. Lett. 1999, 1, 9971000.
Bohm, V. P. W.; Herrmann, W. A. Chem. Eur. J. 2000, 6, 10171025.
Herrmann, W. A.; Bohm, V. P. W. J. Organomet. Chem. 1999, 572, 141145.
Xu, L.; Chen, W.; Xiao, J. Organometallics 2000, 19, 11231127.

Moreno-Manas,
M.; Pleixats, R. Acc. Chem. Res. 2003, 36, 638643.
Suzuki, A. J. Organomet. Chem. 1999, 576, 147168.
Mathews, C. J.; Smith, P. J.; Welton, T. Chem. Commun. 2000, 12491250.
McLachlan, F.; Mathews, C. J.; Smith, P. J.; Welton, T. Organometallics 2003, 22, 53505357.
Dupont, J.; Spencer, J. Angew. Chem., Int. Ed. 2004, 43, 52965297.
Seddon, K. R.; Stark, A.; Torres, M.-J. Pure Appl. Chem. 2000, 72, 22752287.
Daguenet, C.; Dyson, P. J. Organometallics 2004, 23, 60806083.
Swatloski, R. P.; Holbrey, J. D.; Rogers, R. D. Green Chem. 2003, 5, 361363.
Arhancet, J. P.; Davis, M. E.; Merola, J. S.; Hanson, B. E. Nature 1989, 339, 454455.
Mehnert, C. P.; Cook, R. A.; Dispenziere, N. C.; Afeworki, M. J. Am. Chem. Soc. 2002, 124, 1293212933.
Stock, F.; Hoffmann, J.; Ranke, J.; Stormann, R.; Ondruschka, B.; Jastorff, B. Green Chem. 2004, 6, 28690.
Pernak, J.; Goc, I.; Mirska, I. Green Chem. 2004, 6, 323329.
Loupy, A., Ed. Microwaves in Organic Synthesis; Wiley-VCH: Weinheim, 2002.
Hayes, B. L. Aldrichim. Acta 2004, 37, 6676.
Perreux, L.; Loupy, A. Tetrahedron 2001, 57, 91999223.
Lidstrom, P.; Tierney, J.; Wathey, B.; Westman, J. Tetrahedron 2001, 57, 92259283.
Nuchter, M.; Ondruschka, B.; Bonrath, W.; Gum, A. Green Chem. 2003, 6, 128141.
Blackwell, H. E. Org. Biomol. Chem. 2003, 1, 12511255.
Kappe, C. O. Curr. Opin. Chem. Biol. 2002, 6, 314320.
VanAtta, S. L.; Duclos, B. A.; Green, D. B. Organometallics 2000, 19, 23972399.
Ardon, M.; Hogarth, G.; Oscroft, D. T. W. J. Organomet. Chem. 2004, 689, 24292435.

861

862

Green Organometallic Chemistry

180.
181.
182.
183.
184.
185.
186.
187.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
203.
204.
205.
206.
207.
208.
209.
210.
211.
212.

213.
214.
215.
216.
217.
218.
219.
220.
221.
222.
223.
224.
225.
226.
227.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
240.
241.
242.

Konno, H.; Sasaki, Y. Chem. Lett. 2003, 32, 252253.


Larhed, M.; Moberg, C.; Hallberg, A. Acc. Chem. Res. 2002, 35, 717727.
Leadbeater, N. E. Chem. Commun. 2005, 28812902.
Leadbeater, N. E.; Marco, M. Angew. Chem., Int. Ed. 2003, 42, 14071409.
Leadbeater, N. E.; Marco, M. J. Org. Chem. 2003, 68, 56605667.
Appukkuttan, P.; Dehaen, W.; Van der Eycken, E. Eur. J. Org. Chem. 2003, 47134716.
Arvela, R. K.; Leadbeater, N. E.; Songi, M. S.; Williams, V. A.; Grenados, P.; Singer, J. D. J. Org. Chem. 2005, 70, 161168.
Arvela, R. K.; Leadbeater, N. E. J. Org. Chem. 2005, 70, 17861790.
He, P.; Haswell, S. J.; Fletcher, P. D. I. Appl. Cat. A: Gen. 2004, 274, 111114.
Arvela, R. K.; Leadbeater, N. E.; Collins, J.; Michael, J. Tetrahedron 2005, 61, 93499355.
Kaiser, N. F. K.; Bremberg, U.; Larhed, M.; Moberg, C.; Hallberg, A. J. Organomet. Chem. 2000, 603, 25.
Kaiser, N. F. K.; Bremberg, U.; Larhed, M.; Moberg, C.; Hallberg, A. Angew. Chem., Int. Ed. 2000, 39, 35963598.
Fischer, S.; Groth, U.; Jung, M.; Schneider, A. Synlett 2002, 20232026.
Bytschkov, I.; Doye, S. Eur. J. Org. Chem. 2001, 44114418.
Villemin, D.; Heroux, M.; Blot, V. Tetrahedron Lett. 2001, 42, 37013703.
Tan, K. L.; Vasudevan, A.; Bergman, R. G.; Ellman, J. A.; Souers, A. J. Org. Lett. 2003, 5, 21312134.
Loupy, A.; Chatti, S.; Delamare, S.; Lee, D.-Y.; Chung, J.-H.; Jun, C.-H. J. Chem. Soc., Perkin Trans. 1 2002, 12801285.
Cintas, P.; Luche, J.-L. Green Chem. 1999, 1, 115125.
Disselkamp, R. S.; Hart, T. R.; Williams, A. M.; White, J. F.; Peden, C. H. F. Ultrason. Sonochem. 2005, 12, 319324.
Mikkola, J.-P.; Kubicka, D.; Kuusisto, J.; Granholm, N.; Salmi, T.; Holmborn, B. J. Chem. Technol. Biotechnol. 2003, 78, 203207.
Toukoniitty, B.; Toukoniiitty, E.; Maki-Arvela, P.; Mikkola, J.-P.; Salmi, T.; Murzin, D. Yu.; Kooyman, P. J. Ultrason. Sonochem. 2006, 13,
6875.
Polac kova, V.; Hutka, M.; Toma, S. Ultrason. Sonochem. 2005, 12, 99102.
Cravotto, G.; Palmisano, G.; Tollari, S.; Nano, G. M.; Penoni, A. Ultrason. Sonochem. 2005, 12, 9194.
Gogate, P. R.; Pandit, A. B. Ultrason. Sonochem. 2004, 11, 105117.
Ehrfeld, W.; Hessel, V.; Lowe, H. Microreactors: New Technology for Modern Chemistry; Wiely-VCH: Weinheim, 2000.
Fletcher, P. D. I.; Haswell, S. J.; Pombo-Villar, E.; Warrington, B. H.; Watts, P.; Wong, S. Y. F.; Zhang, X. L. Tetrahedron 2002, 58,
47354757.
Haswell, S. J.; Watts, P. Green Chem. 2003, 5, 240249.
Watts, P.; Haswell, S. J. Chem. Soc. Rev. 2005, 34, 235246.
Madou, M. Fundamentals of Microfabrication; CRC Press: Boca Raton, FL, 1997.
Manz, A.; Harrison, D. J.; Verpoorte, E.; Fettinger, J. C.; Ludi, H.; Widmer, H. M. Chimia 1991, 45, 103105.
Jensen, K. F. In Microreactor Technology and Process Intensification; Wang, Y., Halladay, J. D., Eds.; ACS Symposium Series 914; American
Chemical Society: Washington, DC, 2005; pp 222.
Kiwi-Minsker, L.; Renken, A. Catal. Today 2005, 110, 214.
Dietzsch, E.; Honicke, D.; Fichtner, M.; Schubert, K.; Weimeier, G. IMRET 4: 4th International Conference of Micro Reaction Technology
Topical Conference Proceedings, AIChE, Spring National Meeting, Atlanta, GA, USA, March 59, 2000, American Institute of Chemical
Engineers, Indianapolis, IN, USA, p 89.
Kobayashi, J.; Mori, Y.; Okamoto, K.; Akiyama, R.; Ueno, J.; Kitamori, T.; Kobayashi, S. Science 2004, 304, 13051308.
Greenway, G. M.; Haswell, S. J.; Morgan, D. O.; Skelton, V.; Styring, P. Sens. Actuators B 2000, 63, 153158.
Haswell, S. J.; OSullivan, B.; Styring, P. Lab Chip 2001, 1, 164166.
Phan, N. T. S.; Brown, D. H.; Styring, P. Green Chem. 2004, 6, 526532.
Phan, N. T. S.; Kan, J.; Styring, P. Tetrahedron 2005, 61, 1206512073.
de Bellefon, C.; Lamouille, T.; Pestre, N.; Bornette, F.; Pennemann, H.; Neumann, F.; Hessel, V. Catal. Today 2005, 110, 179187.
nal, Y; Lucas, M.; Claus, P. Chem.-Ing.-Tech. 2005, 77, 101105.
O
Sanchez Marcano, J.; Tsotsis, T. T. Catalytic Membranes and Membrane Reactors; Wiley-VCH: Weinheim, 2002.
Sanchez Marcano, J. In Aqueous-Phase Organometallic Catalysis, 2nd ed.; Cornils, B., Herrmann, W. E., Eds.; Wiley-VCH: Weinheim, 2004;
pp 122131.
Laue, S.; Greiner, L.; Woltinger, J.; Liese, A. Adv. Synth. Catal. 2001, 343, 711720.
Nair, D.; Luthra, S. S.; Scarpello, J. T.; White, L. S.; Freitas dos Santos, L. M.; Livingston, A. G. Desalination 2002, 147, 301306.
Bergbreiter, D. E. Catal. Today 1998, 42, 389397.
Bergbreiter, D. E.; Zhan, L.; Mariagnanam, V. M. J. Am. Chem. Soc. 1993, 115, 92959296.
Bergbreiter, D. E.; Liu, Y. S. Tetrahedron Lett. 1997, 38, 78437846.
Wang, Y.; Jin, Z. Trends in Organometallic Chemistry 2002, 4, 7179. CA141: 190824
Jin, Z.; Wang, Y.; Jiang, J.; Wen, F. Current Topics in Catalysis 2002, 3, 1532. CA141: 206536
Wang, Y.; Jiang, J.; Miao, Q.; Wu, X.; Jin, Z. Catal. Today 2002, 74, 8590.
Liu, C.; Jiang, J.; Wang, Y.; Cheng, F.; Jin, Z. J. Mol. Catal. A: Chem. 2003, 198, 2327.
Wang, Y.; Wu, X.; Cheng, F.; Jin, Z. J. Mol. Catal. A: Chem. 2003, 195, 133137.
Diressen-Hollscher, B. Adv. Catal. 1998, 42, 473505.
Burgess, K. Solid-Phase Organic Synthesis; Wiley: New York, 2000.
De Vos, D. E., Vankelecom, I. F. J., Jacobs, P. A., Eds. Chiral Catalyst Immobilization and Recycling; Wiley-VCH: Weinheim, 2000.
Seneci, P. Solid-phase Synthesis and Combinatorial Techniques; Wiley: New York, 2001.
Shuttleworth, S. J.; Allin, S. M.; Sharma, P. K. Synthesis 1997, 12171239.
Ley, S. V.; Baxendale, I. R.; Bream, R. N.; Jackson, P. S.; Leach, A. G.; Longbottom, D. A.; Nesi, M.; Scott, J. S.; Storer, I.; Taylor, R. S.
J. Chem. Soc., Perkin Trans. 1 2000, 38154195.
Montheard, P.; Jegat, C.; Camps, C. J. M. J. Macromol. Sci., Rev. Macromol. Chem. Phys. 1999, C39, 135174.
Astruc, D.; Chardac, F. Chem. Rev. 2001, 101, 29913023.
Leadbeater, N. E.; Marco, M. Chem. Rev. 2002, 102, 32173274.
McNamara, C. A.; Dixon, M. J.; Bradley, M. Chem. Rev. 2002, 102, 32753300.
Barrett, A. G. M.; Hopkins, B. T.; Kobberling, J. Chem. Rev. 2002, 102, 33013324.

Green Organometallic Chemistry

243.
244.
245.
246.
247.
248.
249.
250.
251.
252.
253.
254.
255.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
269.
270.
271.
272.
273.
274.
275.
276.
277.
278.
279.
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
291.
292.
293.
294.
295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.

Dickerson, T. J.; Reed, N. N.; Janda, K. D. Chem. Rev. 2002, 102, 33253344.
Bergbreiter, D. E. Chem. Rev. 2002, 102, 33453384.
Fan, Q.-H.; Li, Y.-M.; Chan, A. S. C. Chem. Rev. 2002, 102, 33853466.
Clark, J. H.; MacQuarrie, D. J. Org. Process Rev. Dev. 1997, 1, 149162.
Song, C. E.; Lee, S. Chem. Rev. 2002, 102, 34953524.
Duchateau, R. Chem. Rev. 2002, 102, 35253542.
De Vos, D. D.; Dams, M.; Sels, B. F.; Jacobs, P. A. Chem. Rev. 2002, 102, 36153640.
van Heerbeek, R.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Reek, J. N. H. Chem. Rev. 2002, 102, 37173756.
Wight, A. P.; Davis, M. E. Chem. Rev. 2002, 102, 35893624.
Lu, Z.-L.; Lindner, E.; Mayer, H. A. Chem. Rev. 2002, 102, 35793588.
Cornils, B. In New Syntheses with Carbon Monoxide; Falbe, J., Ed.; Springer: Berlin, 1980; Chapter 1.
Johnson, T. H. U.S. Patent 4,584,411, 1985.
Frohning, C. D.; Kohlpainter, C. W. In Applied Homogeneous Catalysis with Organometallic Compounds; Cornils, B., Herrmann, W. E., Eds.;
Wiley-VCH: Weinheim, 1996; pp 6165.
Sandee, A. J.; Reek, J. N. H.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Am. Chem. Soc., 2001, 128, 84688476.
Tulchinsky, M. L.; Miller, D. J. U.S. Patent 6,350,819, 2002.
Banet Osuna, A. M.; Chen, W.; Hope, E. G.; Kemmitt, R. D. W.; Paige, D. R; Stuart, A. M.; Xiao, J.; Xu, L J. Chem. Soc., Dalton Trans. 2000,
40524055.
Hu, Y.; Chen, W.; Banet Osuna, A. M.; Stuart, A. M.; Hope, E. G.; Xiao, J. Chem. Commun. 2001, 725726.
Hu, Y.; Chen, W.; Banet Osuna, A. M.; Iggo, J. A.; Xiao, J. Chem. Commun. 2002, 788789.
Meehan, N. J.; Sandee, A. J.; Reek, J. N. H.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Poliakoff, M. Chem. Commun. 2000, 14971498.
Solinas, M.; Jiang, J.; Stelzer, O.; Leitner, W. Angew Chem., Int. Ed. 2005, 44, 22912295.
Frohning, C. D.; Kohlpainter, C. W. In Applied Homogeneous Catalysis with Organometallic Compounds; Cornils, B., Herrmann, W. E., Eds.;
Wiley-VCH: Weinheim, 1996; pp 8082.
Chen, J.; Alper, H. J. Am. Chem. Soc. 1997, 119, 893895.
Ajjou, A. N.; Alper, H. J. Am. Chem. Soc. 1998, 120, 14661468.
Gong, A.; Fan, Q.; Chen, Y.; Liu, H.; Chen, C.; Xi, F. J. Mol. Catal. A: Chem. 2000, 159, 225232.
Reek, J. N. H.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. In Rhodium Catalyzed Hydroformylation; van Leeuwen, P. W. N. M., Claver, C.,
Eds.; Kluwer: Dordrecht, 2000; pp 253279.
Monflier, E.; Fremy, G.; Castanet, Y.; Mortreux, A. Angew Chem., Int. Ed. Engl. 1995, 34, 22692271.
Reetz, M. T.; Waldvogel, S. R. Angew. Chem., Int. Ed. Engl. 1997, 36, 865867.
Shimizu, S.; Shirakawa, S.; Sasaki, Y.; Hirai, C. Angew. Chem., Int. Ed. 2000, 39, 12561258.
Sueur, B.; Leclercq, L.; Sauthier, M.; Castanet, Y.; Mortreux, A.; Bricout, H.; Tilloy, S.; Monflier, E. Chem. Eur. J. 2005, 11, 62286236.
Zheng, X.; Jiang, J.; Liu, X.; Jin, Z. Catal. Today 1998, 44, 175182.
Jin, Z.; Zheng, X.; Fell, B. J. Mol. Catal. A: Chem. 1997, 116, 5558.
Jiang, J.; Wang, Y.; Liu, C.; Xiao, Q.; Jin, Z. J. Mol. Catal. A: Chem. 2001, 171, 8589.
Horvath, I. T.; Kiss, G.; Cook, R. A.; Bond, J. E.; Stevens, P. A.; Rabai, J.; Mozeleski, E. J. J. Am. Chem. Soc. 1998, 120, 31333143.
Foster, D. F.; Adams, D. J.; Gudmunsen, D.; Stuart, A. M.; Hope, E. G.; Cole-Hamilton, D. J. Chem. Commun. 2002, 722723.
Foster, D. F.; Gudmunsen, D.; Adams, D. J.; Stuart, A. M.; Hope, E. G.; Cole-Hamilton, D. J.; Schwarz, G. P.; Pogorzelec, P. Tetrahedron
2002, 58, 39013910.
Mathivet, T.; Monflier, E.; Castanet, Y.; Mortreux, A.; Couturier, J. L. Tetrahedron 2002, 58, 38773888.
Perperi, E.; Huang, Y.; Angeli, P.; Manos, G.; Mathison, C. R.; Cole-Hamilton, D. J.; Adams, D. J.; Hope, E. G. Dalton Trans. 2004,
20622064.
Brasse, C. C.; Englert, U.; Salzer, A.; Waffenschmidt, H.; Wasserscheid, P. Organometallics 2000, 19, 38183823.
Dupont, J.; Silva, S. M.; de Souza, R. F. Catal. Lett. 2001, 77, 131133.
Wasserscheid, P.; Waffenschmidt, H.; Machnitzki, P.; Kottsieper, K. W.; Stelzer, O. Chem. Commun. 2001, 451452.
Bronger, R. P. J.; Silva, S. M.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Chem. Commun. 2002, 30443045.
Kong, F.; Jiang, J.; Jin, Z. Catal. Lett. 2004, 96, 6365.
Horvath, I. T. Catal. Lett. 1990, 6, 4348.
van Leeuwen, P. W. N. M.; Sandee, A. J.; Reek, J. N. H.; Kamer, P. C. J. J. Mol. Catal. A: Chem. 2002, 182183, 107123.
Naughton, M. J.; Drago, R. S. J. Catal. 1995, 155, 383389.
Mehnert, C. P.; Cook, R. A.; Dispenziere, N. C.; Afeworki, M. J. Am. Chem. Soc. 2002, 124, 1293212933.
Mehnert, C. P. Chem. Eur. J. 2005, 11, 5056.
Riisager, A.; Wassercheid, P.; van Hal, R.; Fehrmann, R. J. Catal. 2003, 219, 452455.
Riisager, A.; Eriksen, K. M.; Wasserscheid, P.; Fehrmann, R. Catal. Lett. 2003, 90, 149153.
Brintzinger, H. H.; Fischer, D.; Mulhaupt, R.; Rieger, B.; Waymouth, R. Angew Chem., Int. Ed. Engl. 1995, 34, 1143--1170.
Bochmann, M. J. Chem. Soc., Dalton Trans. 1996, 255270.
Kaminsky, W.; Arndt, M. Adv. Polym. Sci. 1997, 127, 143187.
Britovsek, G. J. P.; Gibson, V. C.; Wass, D. F. Angew. Chem., Int. Ed. 1999, 38, 419447.
Ittel, S. D.; Johnson, L. K.; Brookhart, M. Chem. Rev. 2000, 100, 11691203.
Mecking, S. Angew. Chem., Int. Ed. 2001, 40, 534540.
Gibson, V. C.; Spitzmesser, S. Chem. Rev. 2003, 103, 283315.
Lovell, P. A., El-Aasser, M. S., Eds. Emulsion Polymerisation and Emulsion Polymers; Wiley: Chichester, 1997.
Mecking, S.; Held, A.; Bauers, F. M. Angew. Chem., Int. Ed. Engl. 2002, 144, 544561.
Claveriea, J. P.; Soula, R. Prog. Polym. Sci. 2003, 28, 619662.
Wang, L.; Lu, R. S.; Bau, R.; Flood, T. C. J. Am. Chem. Soc. 1993, 115, 69997000.
Held, A.; Bauers, F. M.; Mecking, S. Chem. Commun. 2000, 301302.
Bauers, F. M.; Mecking, S. Macromolecules 2001, 34, 11651171.
Mecking, S.; Bauers, F. M.; Thomann, R. Polym. Mater. Sci. Eng. 2001, 84, 10491050.
Tomov, A.; Broyer, J.-P.; Spitz, R. Macromol. Symp. 2000, 150, 5358.

863

864

Green Organometallic Chemistry

307. Soula, R.; Novat, C.; Tomov, A.; Spitz, R.; Claverie, J.; Drujon, X.; Maligne, J.; Saudemont, T. Macromolecules 2001, 34, 20222026.
308. Held, A.; Weiss, F.; Mecking, S. Polym. Prepr. 2001, 42, 466467.
309. Brookhart, M. S.; Johnson, L. K.; Killian, C. M.; Arthur, S. D.; Feldman, J.; McCord, E. F.; Mclain, S. J.; Kreutzer, K. A.; Bennett, A. M. A.;
Coughlin, E. B., et al. (E.I. Du Pont De Nemours and Co.). WO 96/23010, 1996.
310. Brown, K. A.; Lamana, W. M.; Siedle, A. R.; Stewart, E. G.; Swanson, P. J. (3M). WO 97/17380, 1997.
311. Rau, A.; Schmitz, S.; Luft, G. Chem. Eng. Technol. 2002, 25, 494498 and references therein.
312. Brant, P.; Rix, F. C.; Kiss, G.; Reynolds, R. (ExxonMobil Chemical Co.). WO-A 06/025545, 2006.
313. Brant, P.; Luft, G. F.; Shutt, J. R.; Smith, L. C.; McLain, D. J.; Burkhardt, T. J. (ExxonMobil Chemical Co.). WO 04/026921, 2004.
314. Super, M. S.; Berluche, E.; Costello, C. A.; Beckman, E. Macromolecules 1997, 30, 368372.
315. Hori, H.; Six, C.; Leitner, W. Macromolecules 1999, 32, 31783182.
316. Kendall, J. L.; Canelas, D. A.; Young, J. L.; DeSimone, J. M. Chem. Rev. 1999, 99, 543563.
317. de Vries, T. J.; Duchateau, R.; Vorstman, M. A. G.; Keurentjes, J. T. F. Chem. Commun. 2000, 263264.
318. de Vries, T. J.; Kemmere, M. F.; Keurentjes, J. T. F. Macromolecules 2004, 11, 42414246.
319. Kemmere, M. F.; de Vries, T. J.; Vorstman, M.; Keurentjes, J. T. F. Chem. Eng. Sci. 2001, 56, 41974204.
320. Smart, B. E. In Organofluorine Chemistry, Principles and Commercial Applications; Banks, R. E., Smart, B. E., Tatlow, J. C., Eds.; Chapter 3
Plenum: New York, 1994; pp 5767.
321. Malanga, M. T.; Newman, T. H. Eur. Pat. Appl. EP 361309A2, 1990.
322. Jones, E.; Walker, J. Ger. Offen. DE 2501239, 1975.
323. Mitani, M.; Furuyama, R.; Mohri, J.-I.; Saito, J.; Ishii, S.; Terao, H.; Nakano, T.; Tanaka, H.; Fujita, T. J. Am. Chem. Soc. 2003, 125,
42934305.
324. Merle, P. G.; Cheron, V.; Hagen, H.; Lutz, M. L.; Spek, A. L.; Deelman, B.-J.; van Koten, G. Organometallics 2005, 24, 16201630.
325. Merle, P. G.; Deelman, B.-J.; van Koten, G. Second European Catalysis Symposium: Organic Catalysis for a Sustainable Development; Pisa, Italy,
September 2326, 2001; Book of Abstracts, Poster 48.
326. Hagerty, R. O.; Laird, R. B.; Risch, M. A.; Shirodkar, P. P.; Jiang, P. (ExxonMobil Chemical Co.). WO 06/009979, 2006.
327. Jiang, P.; Shutt, J. R.; Speed, C. S.; Hagerty, R. O.; Shirodkar, P. P. (ExxonMobil Chemical Co.). WO 06/002132, 2006.
328. Cheron, V.; Couturier, J.-L.; Hagen, H.; van Koten, G.; Deelman, B.-J. (to ATOFINA). Int. Pat. Appl. WO/0214337, 2002.
329. Welton, T. Chem. Rev. 1999, 99, 20712083.
330. Carlin, R. T.; Wilkes, J. S. J. Mol. Catal. 1990, 63, 125129.
331. Pinheiro, M. F.; Mauler, R. S.; de Souza, R. F. Macromol. Rapid Commun. 2001, 22, 425428.
332. Lavastre, O.; Bonnette, F.; Razavi, A. (Atofina res.). WO 05/047350, 2005.
333. Lavastre, O.; Bonnette, F.; Razavi, A. (Atofina res.). WO 05/030392, 2005.
334. Hlatky, G.G. (Equistar Chem.). WO 01/81436, 2001.
335. Carlin, R. T.; Osteryoung, R. A.; Wilkes, J. S.; Rovang, J. Inorg. Chem. 1990, 29, 30033009.
336. Shaughnessy, K. H.; Klingshirn, M. A.; PPool, S. J.; Holbrey, J. D.; Rogers, R. D. In Ionic Liquids As Green Solvents: Progress and Prospects; ACS
symposium series 856; Rogers, R. D., Seddon, K. R., Eds.; American Chemical Society: Washington, DC, 2003; pp 300313.
337. Wasserscheid, P.; Gordon, C. M.; Hilgers, C.; Muldoon, M. J.; Dunkin, I. R. Chem. Commun. 2001, 11861187.
338. Wasserscheid, P.; Gordon, C. M.; Hilgers, C.; Muldoon, M. J.; Dunkin, I. R. Chem. Commun. 2001, 1700.
339. Wasserscheid, P.; Hilgers, C.; Keim, W. J. Mol. Catal. A: Chem. 2004, 214, 8390.
340. Gu, Y. L.; Shi, F.; Deng, Y. Q. Catal. Commun. 2003, 4, 597601.
341. Shaughnessy, K. H.; PPool, S. J.; Klingshirn, M. A.; Rogers, R. D. Polym. Prepr. 2004, 44, 317318.
342. Bernardo-Gusmao, K.; Queiroz, L. F. T.; de Souza, R. F.; Leca, F.; Loup, C.; Reau, R. J. Catal. 2003, 219, 5962.

You might also like