Microwaves in Organic Synthesis
Microwaves in Organic Synthesis
Microwaves in Organic Synthesis
Edited by
A. Loupy
474 pages
2002
Hardcover
ISBN 3-527-30404-5
2nd Edition
ca. 850 pages
2002
Hardcover
ISBN 3-527-30603-X
K. Tanaka (ed.)
Edited by
Andr Loupy
Dr. Andr Loupy & This book was carefully produced. Never-
Laboratoire des Ractions Slectives theless, authors, editors and publisher do not
sur Supports warrant the information contained therein to
Universit Paris-Sud be free of errors. Readers are advised to keep
bat. 410 in mind that statements, data, illustrations,
91405 Orsay cedex procedural details or other items may
France inadvertently be inaccurate.
ISBN 3-527-30514-9
V
Contents
Preface XVIII
List of Authors XXI
N-alkylation of 2-halopyridines 85
Nucleophilic aromatic substitutions 86
Synthesis of phosphonium salts 86
3.7.2 Bimolecular Reactions with One Charged Reactant 87
3.7.2.1 Anionic SN2 Reactions Involving Charge-localized Anions 88
Selective dealkylation of aromatic alkoxylated compounds 88
Alkylation of dianhydrohexitols under phase-transfer
catalysis (PTC) conditions 89
The Krapcho reaction 90
Anionic b-elimination 91
3.7.2.2 Anionic SN2 Reactions Involving Charge-delocalized Anions 92
Alkylation of potassium benzoate 92
Pyrazole alkylation in basic media 93
Selective alkylation of b-naphthol in basic media 93
3.7.2.3 Nucleophilic Additions to Carbonyl Compounds 94
Saponification of hindered aromatic esters 94
PTC transesterification in basic medium 94
Ester aminolysis in basic medium 95
3.7.2.4 Reactions Involving Positively Charged Reactants 97
Friedel±Crafts acylation of aromatic ethers 97
Formylation using Vilsmeier reagent 98
SN2 reactions with tetralkylammonium salts 99
3.7.3 Unimolecular Reactions 99
3.7.3.1 Imidization Reaction of a Polyamic Acid 99
3.7.3.2 Cyclization of Monotrifluoroacetylated o-Arylenediamines 100
3.7.3.3 Intramolecular Nucleophilic Aromatic Substitution 101
3.7.3.4 Intramolecular Michael Additions 102
3.7.3.5 Deprotection of Allyl Esters 103
3.8 Illustrative Examples of the Effects of Selectivity 103
3.8.1 Benzylation of 2-Pyridone 104
3.8.2 Addition of Vinylpyrazoles to Imine Systems 104
3.8.3 Stereo Control of b-Lactam Formation 105
3.8.4 Cycloaddition to C70 Fullerene 106
3.8.5 Selective Alkylation of 1,2,4-Triazole 106
3.8.6 Rearrangement of Ammonium Ylides 108
3.9 Concerning the Absence of Microwave Effects 108
3.10 Conclusions 109
References 110
Index 487
XIX
Preface
The domestic microwave oven is a serendipitous invention. Percy Spencer was work-
ing for Raytheon, a company heavily involved with radar during World War II, when
he noticed the heat generated by a radar antenna. In 1947 an appliance called a Ra-
darange appeared on the market for food processing. The first kitchen microwave
oven was introduced by Tappan in 1955. Sales of inexpensive domestic ovens now re-
present a multibillion-dollar (euro) annual market.
Numerous other uses of microwaves have appeared in the recent years ± essen-
tially drying of different types of material (paper, rubber, tobacco, leather ¼), treat-
ment of elastomers and vulcanization, extraction, polymerization, and many applica-
tions in the food-processing industry.
The field of microwave-assisted organic chemistry is therefore quite young. The
first two pioneering publications from the groups of R. Gedye and R.J. Giguere ap-
peared in 1986. These authors described several reactions completed within a few
minutes when conducted in sealed vessels (glass or Teflon) in domestic ovens. If the
feasibility of the procedure was thus apparent, a few explosions caused by the rapid
development of high pressure in closed systems were also reported. To prevent such
drawbacks safer techniques were developed ± reactions in open beakers or flasks or
solvent-free reactions, as developed essentially since 1987 in France ± in Caen (D. Vil-
lemin), Orsay (G. Bram and A. Loupy), and Rennes (J. Hamelin and F. Texier-Boul-
let). Combination of solvent-free procedures with microwave irradiation constitutes
an interesting and well-admitted approach within the concepts of Green Chemistry.
This coupling takes advantage both of the absence of solvent and of microwave tech-
nology under economical, efficient, and safe conditions with minimization of waste
and pollution.
The goal of this book is to focus on the different fields of application of this tech-
nology in different aspects of organic synthesis. The chapters, which complement
each other, were written by the most eminent scientists well-recognized in their own
field.
After essential revision, and description of wave±material interactions, microwave
technology, and equipment (Chapt. 1) the concepts of microwave-assisted organic
chemistry in pressurized reactors are described (Chapt. 2). Special emphasis on the
possible intervention of a specific (non-purely thermal) microwave effect is discussed
in Chapt. 3 and this is followed by up-to-date reviews of microwave-assisted organic
synthesis in homogeneous media (Chapt. 4), under the action of phase-transfer cata-
lysis (Chapt. 5), using mineral solid supports under ªdry mediaº conditions
(Chapt. 6), and more specifically on graphite (Chapt. 7). Applications in which mi-
crowave-assisted technology has afforded spectacular results and applications are
discussed extensively in Chapt. 8 (heterocyclic chemistry) and Chapt. 9 (on cycload-
ditions). Finally, the techniques have led to fruitful advances in microwave catalysis
(Chapt. 10) and when applied to organometallic chemistry using transition metal
complexes (Chapt. 11) and new, very promising, techniques are now under develop-
ment as a result of applying microwave irradiation to combinatorial chemistry
(Chapt. 12), radiochemistry (Chapt. 13), and photochemistry (Chapt. 14).
I wish to thank sincerely all my colleagues and, nevertheless, (essentially) friends
involved in the realization of this book. I hope to express to all of them my friendly
and scientific gratitude as eminent specialists who agreed to devote their compe-
tence and time to submitting and reviewing papers to ensure the success of this
book.
List of Authors
1
Wave±Material Interactions, Microwave Technology
and Equipment
Didier Stuerga and Michel Delmotte
1.1
Fundamentals of Microwave±Matter Interactions
The objective of this first part of the book is to explain in a chemically intelligible
fashion the physical origin of microwave±matter interactions. After consideration of
the history of microwaves, and their position in the electromagnetic spectrum, we
will examine the notions of polarization and dielectric loss. The orienting effects of
the electric field, and the physical origin of dielectric loss will be analyzed, as will
transfers between rotational states and vibrational states within condensed phases. A
brief overview of thermodynamic and athermal effects will also be given.
1.1.1
Introduction
According to the famous chemistry dictionary of N. Macquer edited in 1775, ªAll the
chemistry operations could be reduced to decomposition and combination; hence, the fire
appears as an universal agent in chemistry as in natureº [1]. Heating has remained the
primary means of stimulating chemical reactions which proceed slowly under ambi-
ent conditions, although several other techniques have been used, e. g. photochemi-
cal, ultrasonic, high pressure, and plasma. In this book, we describe results obtained
with the help of microwave heating. Microwave heating or dielectric heating, an al-
ternative to conventional conductive heating, uses the property of some products
(liquids and solids) to transform electromagnetic energy into heat. This ªin situº
mode of energy conversion is very attractive for applications in chemistry and mate-
rial processing.
If the effect of the temperature on reaction rate is well known, and is very easy
to express, the problem is very different for effects of electromagnetic waves. What
can be expected from the orienting action of electromagnetic fields at molecular
levels? Are electromagnetic fields able to enhance or modify collisions between re-
agents? All these questions are raised by the use of microwaves energy in chemis-
try.
1.1.1.1 History
The first announcement of a microwave oven was probably a magazine article about
a newly developed Radarange for airline use [2, 3]. This device, it was claimed, could
bake biscuits in 29 s, cook hamburgers in 35 s, and grill frankfurters in 10 s. The
first commercial microwave oven was developed by P. Spencer, of a company called
Raytheon, in 1952 [4]. There is a legend that P. Spencer, who studied high-power mi-
crowave sources for radar applications, observed the melting of a chocolate bar in his
pocket. Another story says that Spencer had some popping corn in his pocket that
began to pop as he was standing alongside a live microwave source [5]. This idea led
to the microwave oven in 1961 and the generation of the mass market. The wide-
spread domestic use of microwave ovens occurred during 1970s and 1980s as a re-
sult of Japanese technology transfer and global marketing. Curiously, industrial ap-
plications were initiated by the domestic oven.
Originally, microwaves played a leading role during the World War II, especially in
the battle of Britain which, thanks to radar, English planes won despite being out-
numbered three-to-one. The first generator of microwave power for radar, called the
magnetron, was designed by Randall and Booth at the University of Birmingham
during the 1940s. They were mass produced in the United States by companies such
as Raytheon.
hc
E ko hn
1
l
1.1.1.3 Energetics
It is well known that g or X photons have energies suitable for excitation of inner
electrons. We can use ultraviolet and visible radiation to initiate chemical reactions
(photochemistry). Infrared radiation excites bond vibrations only whereas hyperfre-
quencies excite molecular rotation. In Tab. 1.1 the energies associated with chemical
bonds and Brownian motion are compared with the microwave photon correspond-
ing to the frequency used in microwave heating systems such as domestic and in-
dustrial ovens (2.45 GHz, 12.22 cm).
4.51 (C±H)
Energy (eV) 0.017 (200 K) 0.04 to 0.44 7.6
3.82 (C±C)
435 (C±H)
Energy (kJ mol±1) 1.64 3.8 to 42 730
368 (C±C)
nian motion, and it obviously cannot induce chemical reactions. If no bond breaking
can occur by direct absorption of electromagnetic energy, then what can be expected
from the orienting effects of electromagnetic fields at molecular levels? Are electro-
magnetic fields able to enhance or to modify collisions between reagents? Do reac-
tions proceed with the same reaction rate with and without electromagnetic irradia-
tion at the same bulk temperature? In the following discussion the orienting effects
of the electric field, the physical origin of the dielectric loss, transfers between rota-
tional and vibrational states in condensed phases, and thermodynamic effects of
electric fields on chemical equilibrium will be analyzed.
1.1.2
The Complex Dielectric Permittivity
where e0 is the dielectric permittivity of vacuum, e' and e@ are the real and imaginary
parts of the complex dielectric permittivity, and e'r and e@r are the real and imaginary
parts of the relative complex dielectric permittivity. The storage of electromagnetic
energy is expressed by the real part whereas the thermal conversion is proportional
to the imaginary part.
zation are linked by Maxwell's equations. The constitutive equation for vacuum is
given by Eq. (3):
D~ e0 E~ 3
where D~ is the electric displacement and E~ the electric field. For a dielectric medium
characterized by e~ , the constitutive equation is given by Eq. (4):
D~ e~ E~ e0 E~ P~ 4
In the global formulation of Eq. (4), we can express the term corresponding to
vacuum and given by Eq. (3). The second and complementary term then defines the
~ The higher the di-
contribution of matter to polarization processes, or polarization P.
electric permittivity of a material, the greater the polarization processes. The polari-
zation process described by P~ has its physical origin in the response of dipoles and
charges to the applied field. Depending on the frequency, electromagnetic fields put
one or more types of charge association under oscillation. In any material there is a
variety of types of charge association:
Depending on the frequency the electromagnetic field can induce one or more
types of charge association under oscillation. Each configuration has its own critical
frequency above which interaction with the field becomes vanishingly small, and the
lower the frequency and the more configurations are excited. For electrons of the in-
ner atomic shells the critical frequency is of the order of that of X-rays. Consequently
an electromagnetic field of wavelength more than 10±10 m cannot excite any vibra-
tions, but rather induces ionization of these atoms. There is no polarizing effect on
the material, which for this frequency has the same dielectric permittivity as
vacuum. For ultraviolet radiation the energy of photons is sufficient to induce transi-
tions of valence electrons. In the optical range an electromagnetic field can induce
distortion of inner and valence electronic shells. Polarization processes result from a
dipole moment induced by distortion of electron shells and are called electronic po-
larizability. In the infrared range electromagnetic fields induce atomic vibrations in
molecules and crystals and polarization processes result from the dipole moment in-
duced by distortion of nuclei positions. These polarization processes are called
atomic polarization. In all the processes mentioned so far, the charges affected by
the field can be considered to be attracted towards their central position by forces
6 1 Wave±Material Interactions, Microwave Technology and Equipment
which are proportional to their displacement by linear elastic forces. This mechani-
cal approach of electronic resonance is only an approximation, because electrons
cannot be properly treated by classical mechanics. Quantitative treatment of these
processes requires the formalism of quantum mechanics. The two types of polariza-
tion process described above can be connected in distortion polarization.
The characteristic material response times for molecular reorientation are 10±12 s.
Then, in the microwave band, electromagnetic fields lead to rotation of polar mole-
cules or charge redistribution. The corresponding polarization processes are denoted
orientation polarization.
thermal agitation. The relationship between the ratio of effective to maximum polari-
zation and the ratio of the potential interaction energy to the thermal agitation is de-
scribed by Fig. 1.3.
We can see that the Langevin function increases from 0 to 1 as the strength of the
electric field is increased and/or the temperature is reduced. The molecules tend to
align with the field direction. For high values of the field orientation dominates over
the disorientation induced by temperature, and all the dipoles tend to become paral-
lel to the applied field. The complete alignment corresponds to saturation of the in-
duced polarization. In many practical situations field strengths are well below their
saturation values. The arrow in Fig. 1.3 corresponds to the usual conditions of micro-
wave heating (temperature close to room temperature, 25 8C, and electric field
strength close to 105 V m±1). According to these results, the electric field strength
commonly used in microwave heating is not sufficient to induce a consequent freez-
ing of the media.
Relaxation processes are probably the most important of the interactions between
electric fields and matter. Debye [6] extended the Langevin theory of dipole orienta-
tion in a constant field to the case of a varying field. He showed that the Boltz-
mann factor of the Langevin theory becomes a time-dependent weighting factor.
When a steady electric field is applied to a dielectric the distortion polarization,
P~Distor, will be established very quickly ± we can say ªinstantaneouslyº compared
with time intervals of interest. But the remaining dipolar part of the polarization
(orientation polarization, P~Orient ) takes time to reach its equilibrium value. When
the polarization becomes complex, the permittivity must also become complex, as
shown by Eq. (5):
e s n2
e~ e0 j e00 n2
5
1 jot
where n is the refractive index and t the relaxation time. All polar substances have a
characteristic time t called the relaxation time (the characteristic time of reorienta-
tion of the dipole moments in the direction of the electric field). The refractive index
corresponding to optical frequencies or very high frequencies is given by Eq. (6):
e n2 6
e s n2
e 0 n2
7
1 o2 t 2
e s n2
e00 ot
8
1 o2 t2
Changes of e' and e@ with frequency are shown in Fig. 1.4. The frequency is dis-
played on a logarithmic scale. The dielectric dispersion covers a wide range of fre-
quencies.
The dielectric loss reaches a maximum given by Eq. (9):
es n2
e00max
9
2
lows the electric field, whereas the other component of the polarization undergoes
chaotic motion leading to thermal dissipation of the electromagnetic energy. This de-
scription is well adapted to gases (low particle density). In fact, for a liquid we must
take into account the effect of collisions with the surroundings and the equilibrium
distribution function is no longer applicable. The Debye theory can define a distribu-
tion function which obeys a rotational diffusion equation. Debye [6, 7] based his the-
ory of dispersion on Einstein's theory of the Brownian motion. He supposed that the
rotation of a molecule caused by an applied field is constantly interrupted by colli-
sions with neighbors, and the effect of these collisions can be described by a resistive
couple proportional to the angular velocity of the molecule. This description is well
adapted to liquids, but not to gases.
The general equation for complex dielectric permittivity is then given by Eq. (11):
e~r 1 rN m2
a
11
e~r 2 3e0 M 3kT
1 j o t
where N is Avogadro number, M is the molar mass, r the specific mass, and a the
atomic polarizability. The relaxation time, t, is a microscopic relaxation time that de-
pends on the average resistive force experienced by the individual molecules. In the
limit of low frequency the Debye expression is obtained for the static permittivity
whereas in the high frequency limit the permittivity will fall to a value which can be
written as the square of the optical index. The first term of the left side corresponds
to the distortion polarization whereas the other term corresponds to the orientation
polarization. For apolar molecules, we obtain the famous Clausius±Mosotti±Lorentz
equation.
Relaxation times
Debye [6] suggested that a spherical or nearly spherical molecule can be treated as a
sphere (radius r) rotating in a continuous viscous medium with the viscosity, Z, of
the bulk liquid. The relaxation time is given by Eq. (12):
10 1 Wave±Material Interactions, Microwave Technology and Equipment
8p Zr 3
t
12
2 kT
For a given molecular system it is, in fact, better to use a formula containing
tinter (T ), a part which depends on the temperature, and a part totally independent of
the temperature, tSteric , as described by Eq. (13):
Relaxation times for dipole orientation at room temperature are between 10±10 s
for small dipoles diluted in a solvent of low viscosity and more than 10±4 s for large
dipoles in a viscous medium such as polymers (polyethylene), or dipole relaxation in
crystals (the relaxation associated with pairs of lattice vacancies). The relaxation time
of ordinary organic molecules are close to a few picoseconds. Thus for a frequency of
2.45 GHz these molecules can follow electric field oscillations, unlike substances
which are strongly associated, e. g. water and alcohols, and therefore are subject to
dielectric loss at 2.45 GHz. Consequently, the solvents for which dielectric loss is ob-
served are water, MeOH, EtOH, DMF, DMSO, and CH2Cl2. For nonpolar solvents
such as C6H6, CCl4, and ethers dielectric loss is negligible, although addition of
small amounts of alcohols can strongly increase dielectric loss and microwave cou-
pling of these solvents.
It is clear that for a substance with dielectric loss, e. g. water and the alcohols, the
molecules do not perfectly follow the oscillations of the electric field. For media with-
out dielectric loss, and for the same reasons as under static conditions, the strength
of the electric field cannot induce rotation of all polar molecules but, statistically, for
a small part only (less than 1%). This means that all the molecules oscillate around
an average direction (precession motion), as shown by Fig. 1.5.
The principal axis of the cone represents the component of the dipole under the
influence of the thermal agitation. The component of the dipole in the cone results
from the field that oscillates in its polarization plane. In this way, in the absence of
Brownian motion the dipole follows a conical orbit. In fact the direction of the cone
changes continuously (because of the Brownian movement) faster than the oscilla-
tion of the electric field; this leads to chaotic motion. Hence the structuring effect of
electric field is always negligible, because of the value of the electric field strength,
and even more so for lossy media.
It is well known that in condensed phases energy transfer can occur between rota-
tional and vibrational states. Indeed, molecular rotation does not actually occur in li-
quids; rotational states turn into vibrational states because of an increase in collisions.
For liquids, the collision rate is close to 1030 collisions s±1. Microwave spectro-
scopy, which studies molecular rotation, only uses dilute gases to obtain pure rota-
tional states of sufficient lifetime. Rotational transitions are broadened by molecular
collisions, because the pressure is close to a few tenths of a bar, as shown in Fig. 1.6.
In conclusion, for condensed phases molecular rotations have quite a short life-
time, because of collisions. The eventual oscillations induced by the electric field are
then dissipated in the liquid state leading to vibration. At collision densities corre-
sponding to liquids the frequency of the collisions become comparable with the fre-
quency of a single rotation, and because the probability of a change in rotational
state on collision is high, the time a molecule exists in a given state is small. It is,
therefore, obvious that the electric field cannot induce organization in condensed
phases such as in the liquid state.
Fig. 1.6 Absorption spectrum for water (gaseous, solution, and liquid).
Above the vapor band is Mecke's rotational analysis [11, 12].
12 1 Wave±Material Interactions, Microwave Technology and Equipment
According to the value of the frequency of the field, and the relaxation time band
in relation to the temperature considered, one can find the three general changes
with temperature of the dielectric properties. Fig. 1.7 gives the three-dimensional
curves describing the dielectric properties in relation to frequency and temperature.
According to the value of the working frequency compared with the relaxation fre-
quency, three general cases could be found:
1. the real and imaginary parts of the dielectric permittivity decrease with tempera-
ture (working frequency lower than relaxation frequency);
2. the real and/or imaginary parts of the dielectric pass through a maximum (work-
ing frequency very close to relaxation frequency).
1.1 Fundamentals of Microwave±Matter Interactions 13
3. the real and imaginary parts of the dielectric permittivity increase with tempera-
ture (working frequency higher than relaxation frequency).
The two solvents most commonly used in microwave heating are ethanol and water.
Values for water are given by Kaatze [13] and those for ethanol by Chahine et al. [14].
Water is close to case 1 because both values decrease with temperature. In contrast,
for ethanol the real part increases and the dielectric loss reaches a maximum at 45 8C
(case 2). For ethanol, in fact, the working frequency is higher than relaxation fre-
quency at room temperature. Ethanol has a single relaxation frequency, close to
1 GHz at 25 8C, and, furthermore, its relaxation frequency increases fairly rapidly
with temperature (3 GHz at 65 8C). For water the working frequency is smaller than
the relaxation frequency at all temperature (17 GHz at 20 8C and 53 GHz at 80 8C).
The pioneering work of Von Hippel [15] and his coworkers, who obtained dielec-
tric data for organic and inorganic materials, still remains a solid basis. Study of di-
electric permittivity as a function of temperature is, however, less well developed,
particularly for solids.
Conduction losses
For highly conductive liquids and solids the loss term not only results from a single
relaxation term, as given by Eq. (8), but also from term resulting from ionic conduc-
tivity, s, as described by Eq. (14):
e s n2 s
e00 ot
14
1 o2 t2 o
The hydroxide ion is the typical example of ionic species with both ionic and dipo-
lar characteristics. For solutions containing large amount of ionic salts this conduc-
tive loss effect can become larger than dipolar relaxation. For solids conduction
losses are usually very slight at room temperature but can change substantially with
temperature. A typical example is alumina, for which dielectric losses are very small
(~10±3) at room temperature but which can reach fusion in several minutes in a mi-
crowave cavity [16]. This effect is because of a strong increase of conduction losses
associated with the thermal activation of the electrons which pass from the oxygen
2p valence band to the 3s3p conduction band. In solids, moreover, conduction losses
are usually enhanced by material defects which sharply reduce the energy gap be-
tween the valence and conduction bands.
Because conduction losses are high for carbon black powder it can be used as lossy
impurities or additives to induce losses within solids for which dielectric losses are
too small.
Magnetic losses
Chemical reagents are primarily concerned with dielectric liquids or solids. For me-
tal oxides such as ferrites, however, magnetic losses occur in the microwave region.
As for a dielectric material, a complex magnetic permeability is defined as given by
Eq. (16):
14 1 Wave±Material Interactions, Microwave Technology and Equipment
m~ m0 j m00 15
The real part is the magnetic permeability whereas the imaginary part is the mag-
netic loss. These losses are quite different from hysteresis or eddy current losses, be-
cause they are induced by domain wall and electron-spin resonance. These materials
should be placed at position of magnetic field maxima for optimum absorption of
microwave energy. For transition metal oxides such as iron, nickel, and cobalt mag-
netic losses are high. These powders can, therefore, be used as lossy impurities or
additives to induce losses within solids for which dielectric loss is too small.
o e0 e00r ~ 2
PDiss
r E
r
16
2
where o is the angular frequency, e0 the dielectric permittivity of vacuum, e (r) the di-
electric loss, and E~ (r) the electric field amplitude. Depending on the dielectric loss
and electric field strength, the dissipation of electromagnetic energy leads to heating
of the irradiated medium. Hence estimation of dissipated power density within the
heated object depends directly on the electric field distribution within the heated ob-
ject and on the dielectric loss. Maxwell's equations can be used to describe the elec-
tromagnetic fields in a lossy medium and an energy balance can be solved to provide
the temperature profiles within the heated reactor. The specificity of microwave heat-
ing results from the thermal dependence of dielectric properties. The complex di-
electric permittivity is highly dependent on temperature and the dynamic behavior
of microwave heating is governed by this thermal change. The electric field ampli-
tude depends, moreover, on the real and imaginary parts of the dielectric permittiv-
ity, which themselves depend on temperature, as described by Eq. (17):
o e0 e00r
T ~ 2
PDiss
r; T E r; e00r
T; e00r
T
17
2
Fig. 1.8 Dependence on temperature of the crowave heating (with and without the elec-
heating expected with conventional and mi- tric field effect) for water and ethanol [19].
1.1.3
Thermodynamic and other Effects of Electric Fields
The thermodynamic effects of electric fields and are well known. Application of an
electric field to a solution can affect the chemical equilibrium. For example, in
Eq. (18) where C has a large dipole moment and B has a small dipole moment the
equilibrium is shifted toward C under the action of an electric field.
ABC 18
Typical examples are the conversion of the neutral form of an amino acid into its
zwitterionic form, the helix±coil transitions in polypeptides and polynucleotides, and
other conformational changes in biopolymers. Reactions of higher molecularity in
which reactants and products have different dipole moments are subject to the same
effect (association of the carboxylic acids to form hydrogen-bonded dimers). Equili-
brium involving ions are often more sensitive to the application of an electric field;
16 1 Wave±Material Interactions, Microwave Technology and Equipment
the field induces a shift toward producing more ions. This is known as the dissocia-
tion field effect (DFE) or the ªsecond Wien's effectª [20].
In principle the effect of an electric field on chemical equilibria can be described
by the thermodynamic relationship described by Eq. (19):
!
u ln K Eq: Dm
19
~
u jE j RT
P;T
where Keq. is the equilibrium constant, |E~ | the field strength (V m±1), and Dm the
molar change of the macroscopic electric moment or the molar polarization for non-
ionic systems. For ionic equilibria it must be pointed out that a true thermodynamic
equilibrium can never be reached, because of the field-induced flow of the ions. The
DFE theory was been developed by Onsager [21]. The most notable result is that Dm
is proportional to the field strength, E. Hence, according to Eq. (20):
ZE ZE
1
d ln K Eq: Dm dE / E 2
20
RT
0 0
which is the integration of Eq. (19), the change in equilibrium constant is propor-
tional to the square of the electric field strength and the effect on equilibrium con-
stant is noticeable only at high field strengths. In practice electric field strengths up
to 107 V m±1 are required to produce a measurable effect upon normal chemical re-
actions. For water at 25 8C, K changes by about 14 % if a field of 100 kV m±1 is ap-
plied. Smaller fields are required to achieve a comparable shift in less polar solvents.
Nonionic equilibria can also be perturbed by the DFE if they are coupled to a rapid
ionic equilibrium. A possible scheme is described by Eq. (21):
A H AH BH 21
in which the slow equilibrium is coupled with an acid±base equilibrium. This is the
same principle as coupling a temperature-independent equilibrium to a strongly de-
pendent equilibrium. Such a scheme has been studied in the helix±coil transition of
poly-a-l-glutamic acid by Yasunaga et al. [22] ± dissociation of protons from the side
chains increases the electric charge of the polypeptide, which in turn induces a tran-
sition from the helix to the coil form, in the dissociation of acetic acids by Eigen and
DeMayer [23], and in the dissociation of water by Eigen and DeMayer [24].
Hence, if thermodynamic effects of electric effect occur, the electric field strengths
necessary are too high compared with ordinary operating conditions of microwave
heating.
1.1 Fundamentals of Microwave±Matter Interactions 17
1.1.4
The Athermal and Specific Effects of Electric Fields
A chemical reaction is characterized by the difference between the free energy of the
reagents and products. According to thermodynamics the reaction is feasible only if
the free energy change is negative. The more negative the free energy change, the
more feasible the reaction. This free energy change for the reaction is the balance be-
tween broken and created chemical bonds. This thermodynamic condition is not,
however, sufficient to ensure the chemical reaction occurs rapidly (i. e. with a signifi-
cant rate of reaction). Kinetic conditions must also be verified to achieve the reaction.
The free energy of activation depends on the enthalpy of activation which expresses
the height of the energy barrier which must be surmounted. This energy condition
is only a necessary condition but not sufficient to ensure the transformation of the
reagents. The relative orientation of the molecules which react is crucial, and this
condition is expressed by the entropy of activation. This entropic term expresses the
need for a geometrical approach to ensure the effectiveness of collisions between re-
agents.
The essential questions raised by the assumption of ªathermalº or ªspecificº ef-
fects of microwaves are, then, the change of these characteristic terms (free energy
of reaction and of activation) of the reaction studied. Hence, in relation to previous
conclusions, five criteria or arguments (in a mathematical sense) relating to the oc-
currence of microwave athermal effects have been formulated by the author [25].
More details can be found in comprehensive papers which analyze and quantify the
likelihood of nonthermal effects of microwaves. This paper provides guidelines
which clearly define the character of nonthermal effects.
Hence, according to these five criteria there can be no doubt that an electric field
cannot have any molecular effect for solutions. First, the orienting effect of electric
field is small compared with thermal agitation, which results from the weakness of
the electric field amplitude. Even if the electric field amplitude were sufficient, the
presence of dielectric loss results in a delay between dipole moment oscillations and
electric field oscillations. Heating of the medium reveals the stochastic character of
molecular motion induced by dissipation of the electromagnetic wave. The third lim-
itation is the annihilation of molecular rotation in liquid state condensed phases. Ac-
cording to our demonstration, under normal operating conditions, it will be proved
that the frequently propounded idea that microwaves rotates dipolar groups is,
mildly speaking, misleading.
If the molecular effects of the electric field are irrelevant to microwave heating of
solutions, this assumption could be envisaged in the use of operating conditions
very far from current conditions. On one hand, it will be necessary to use an electric
field of higher amplitude, or to reduce the temperature according to the Langevin
function. This last solution is obviously antinomic with conventional chemical ki-
netics, and the first solution is, currently, technologically impossible. It will, on the
other hand, be necessary to avoid reaction media with dielectric loss. The molecular
effects of the microwave electric field could, paradoxically, be observed for a medium
which is not heated by the action of microwave irradiation.
18 1 Wave±Material Interactions, Microwave Technology and Equipment
1.1.5
Conclusions
The interaction between a dipole and an electric field is clearly interpreted by quan-
tum theory. Coupling is weaker than with magnetic fields, and when a dipole popu-
lation is subjected to an electric field there is such a demultiplication of quantum le-
vels that they are very close to each other. The interaction energy is continuous, and
we have to use Boltzmann or Langevin theories. Because of the weak coupling be-
tween the dipole and electric field, and the lack of quantified orientations, the study
of electric dipole behavior gives less information about the dipole itself than about
its surroundings. Indeed electric dipoles are associated with molecular bonds (the
electric dipole moment results from the distribution of positive and negative charges
on the studied molecule; if they are centered at different points then the molecule
has a permanent dipole and the molecule is polar). Any motion of electric dipoles in-
duced, for example, by interaction with an electric field, lead to correlative motion of
molecular bonds, whereas motion of magnetic moments is totally independent of
any molecular motion. Consequently, studies of dielectric properties must be those
of ªgroup propertiesª. Those properties cannot be modeled by a single dipole; a
group of dipoles interacting among themselves would be a key ingredient for these
models. The origin of confusion between the behavior of a single dipole and a collec-
tion of dipoles (in other words differences between dilute and condensed phases) is
the most important problem, and the source of the confusion of those who claim
that microwave effect results from orienting effect of the electric field.
In conclusion, is it necessary to obtain a microwave athermal effect to justify mi-
crowave chemistry? Obviously not ± it is not necessary to present microwaves effects
in a scientific disguise. There are many examples in which microwave heating re-
sults in specific time±temperature histories and gradients which cannot be achieved
by other means especially for solid materials. Hence, rather than claiming nonther-
mal effects it is better to claim a means or a tool which induces a specific thermal
history.
1.2
Overview of Microwave Reactor Design and Laboratory and Industrial Equipment
The objective of this section is to provide the reader with a basic knowledge of micro-
wave techniques used in the design and the construction of microwave ovens.
After some considerations relating to microwave technology, we will examine mi-
crowave ovens and reactor background. The limits of domestic ovens and tempera-
ture measurements will be analyzed, as well as design principles of microwave appli-
cators. Next, a brief overview of laboratory, experimental and industrial equipment
will be given.
1.2 Overview of Microwave Reactor Design and Laboratory and Industrial Equipment 19
1.2.1
Microwave Ovens and Reactors ± Background
This part has been designed by the authors as a practical tool for people interesting
by using laboratory or industrial equipment. An overall coverage of microwave ovens
design would require more space than is available in this chapter. An excellent cover-
age of applicator theory can be found in [5, 26] and basis of electromagnetic wave-
guides and cavities can be found elsewhere [27].
the reflected waves. The goal of microwave oven designer is to ensure that all the in-
cident power is absorbed by the load. In other words, the resonant frequency of the
loaded oven (and not the empty oven) should be close to frequency of the magnetron
(i. e. 2.45 GHz). If too much energy is reflected back to the source, the magnetron
may be damaged. That is the reason why it is not advised to run empty domestic
ovens. However, most commercial ovens are protected by a thermal automatic cutoff
in case of poor matching between magnetron and oven.
1.2.2
Commercial Laboratory Microwave Reactors
Only a few microwave reactors equipped with efficient temperature control systems
for safe microwave synthesis at laboratory scale are currently available on the market.
These systems lead to reproducible operating conditions.
eter [49]. This system allow measurements on the bottom of the heated tube. This
system was a good solution for laboratory experiments. Users should be aware of the
problem of temperature measurements. It is better to calibrate the system for each
media to be tested [50]. This setup could be associated with a computer in order to
regulate reagents temperature. The Synthewave 402 and Synthewave 1000 are de-
scribed by Fig. 1.9.
The Soxwave 100 is a variant which has been designed for extractions such as
Soxhlet which are laborious. The extraction tube is capped with a cooling column.
Optional temperature control during extraction was available. Maxidigest (one mi-
crowave unit 15 W to 250 W) and Microdigest (3 up to 6 independent digestions at
one time with integrated magnetic stirrers) are other variants for digestion.
perature control and variable speed stirring are available. The MARS-S is described
by Fig. 1.10 whereas the Discover system is described by Fig. 1.11.
The Star Systems 100 is device designed for digestions and extractions such as
Soxhlet which are laborious
1.2.3
Experimental Microwave Reactors
The microwave power could be adjusted in order to allow constant pressure within
the vessel. A incorporated pressure release valve permits to use this experimental de-
vice routinely and safely. Furthermore, an inert gas as argon could be introduced
within the reactor to avoid sparking risk with flammable solvents. This experimental
device is able to raise temperature from ambient to 200 8C in less than 20 s (pressure
is close to 1.2 Mpa and heating rate is close to 7 8 s±1). The RAMO system has been
designed for nanoparticles growing and elaboration [59±62]. The RAMO system is a
batch system. It could be easily transpose to continuous process with industrial scale
(several hundred kilograms by seconds).
1.2.4
Industrial Equipment ± Batch or Continuous Flow?
Industrial equipment is classically divided into batch reactor and continuous flow
systems. The typical sizes of batch reactor for industrial heterogeneous or homoge-
neous reactions are close to several cubic meters of liquids reagents. The design of
microwave applicators able to heat these volumes meets scientific and technological
limits of microwave penetration within high lossy media. The penetration depth of
microwaves are beneath decimeter for classical solvents as water, alcohols, etc¼ So,
the microwave treatment of a cubic meter of solvents is generally technologically im-
possible. Moreover, several cubic meters of liquids reagents inside batch reactors
should need microwave power close to one megawatt power. This microwave power
is out of range of the classical devices of microwave technology, generally close to
one hundred kilowatts at 915 MHz or 2.45 GHz frequencies.
Consequently, we consider that the industrial scale technological management of
microwave assisted chemical reaction is no compatible with batch reactors coupled
with multimode applicators. Some typical processes with a systematic decrease of
the dielectric losses of the concerned reactant, such as filtration and drying of
mineral or pharmaceutical powders are compatible with multimode applicators. To
our knowledge, the only industrial batch microwave device is the microwave variant
of the Turbosphere (ªall in one solutionº mixer/granulator/dryer designed by Moritz
1.2 Overview of Microwave Reactor Design and Laboratory and Industrial Equipment 29
company) described by Fig. 1.17. This industrial equipment is sold by the Pierre
Guerin company, well known in the pharmaceutical industry [64].
In contrast, several industrial continuous flows systems are commercially avail-
able. The continuous flow systems permit the spatial localization of a small quantity
of matter (between grams and one kilogram) in front of a microwave source with
power magnitude close to several kilowatts (typically close to 2 or 6 kW). According
to the flow and heating rates expected, several modular power units should be asso-
ciated and microwave power magnitude could be typically about ten kW. The indus-
trial continuous flows systems described in the following have been designed for
food industries and drying and not specially for chemistry industry. However, these
devices could be used for chemical processes.
This equipment could be used for chemical reactions based on a strong solid±gas
interaction with gas adsorbed on powder such as limited air oxidation or with gas re-
lease (water, ammonia) such as esterification. The oversized applicator structure per-
mits the design of dielectric pipe to manage such matter transfers. This equipment
can be also used for many reactions on solid supports. A typical unit is powered with
microwave generators units of 2 or 6 kW for a total microwave power close to 20 or
60 kW.
Fig. 1.19 General view of the Thermostar sys- crossing by a vertical pipe (left) or horizontal pipe
tem. Six successive parallelepiped applicators with screw (right) (courtesy of MES Company).
References
2
Microwave-assisted Organic Chemistry in Pressurized Reactors
Christopher R. Strauss
2.1
Introduction
The highest priorities for the chemical industry now are process and product safety
and the environment [1±5]. New technologies and methods for ªgreenº and sustain-
able chemistry are subjects of intense activity [6]. In that regard, the potential of
microwave heating for organic synthesis attracted interest soon after the first re-
ports [7, 8] appeared in 1986. Early reactions were performed with domestic micro-
wave ovens and relatively primitive containers. Rate enhancement of up to three or-
ders of magnitude was observed, but the time savings were offset by hazards such as
deformation of the vessels and explosions.
Thus, the major challenge confronting the early microwave chemists was to retain
or enhance the benefits of the technology while concurrently avoiding the risks. Sev-
eral groups concluded that microwave heating was incompatible with organic sol-
vents and investigated solvent-free conditions including ªdryº media, usually with
open vessels in domestic microwave ovens [9±12]. That area expanded rapidly, aided
by the ready availability of inexpensive microwave equipment and encouraged by a
diverse range of reactions awaiting exploration.
Synthetic chemists desire well defined reaction conditions. Process chemists de-
mand them. Nonuniform heating and difficulties with mixing and temperature
measurement are technical constraints that initially limited the scale of microwave
chemistry with ªdryº media and have not yet been overcome. Poor reproducibility
also has been reported, probably resulting from differences in performance and op-
eration of individual domestic microwave ovens [13±15]. Consequently, most, if not
all, of the disclosed applications of ªdryº media are laboratory-scale preparations.
However, as discussed in other chapters, this does not prevent their being interesting
and useful.
An alternative approach, developed by Bose [16±19] and termed ªmicrowave-in-
duced organic reaction enhancementº (MORE) chemistry, employed polar, high-boil-
ing solvents with open vessels in unmodified domestic microwave ovens. The sol-
vents had dielectric properties suitable for efficient coupling of microwave energy
and rapid heating to temperatures that although high, were typically some 20±30 8C
below boiling. Apparent disadvantages were the need for high boiling polar solvents
such as dimethylsulfoxide, ethylene glycol, diglyme, triglyme, N-methylmorpholine,
N,N-dimethylformamide and 1,2-dichlorobenzene that have relatively similar boiling
points and present difficulties for recycling and for isolation of liquid products.
The above strategies were pursued concurrently with our approach, its objective
being to interface microwave technology with conventional organic chemistry in-
cluding the use of low boiling organic solvents. Consequently, dedicated reactors
were required that could operate reliably, safely and routinely in the presence of
flammable compounds. When we began, such equipment was not available and ap-
parently not contemplated for manufacture by others.
Our f irst tasks were first to design and fabricate appropriate systems and then to
apply them to a diverse range of organic reactions. The following features were con-
sidered essential:
2.2
Rationale for Pressurized Microwave Reactors
and cooling steps, the rates of temperature increase and decrease are usually low
and thermal gradients are difficult to avoid.
It was surmised that with appropriate vessels, microwaves would be absorbed pre-
dominantly throughout the sample. In addition, the energy could be applied or with-
drawn instantaneously and the input could be adjusted readily to match that re-
quired. Thus, it appeared likely that direct, bulk heating, combined with efficient
stirring of the sample would diminish temperature gradients, particularly near the
walls of the vessel.
2.2.1
The Continuous Microwave Reactor (CMR)
Two years after the 1986 reports of Gedye [7] and Giguere [8], a prototype continu-
ous microwave reactor (CMR) was produced [21, 22]. It was the first microwave sys-
tem designed to cope with organic solvents. Patents were obtained. Commercial
systems developed under license, typically comprised a microwave cavity fitted with
a vessel of microwave-transparent, inert material (Fig. 2.1). Plumbing in the micro-
wave zone was connected to a metering pump and pressure gauge at the inlet end
and to a heat exchanger and a pressure-regulating valve at the effluent end. The
heat exchanger enabled rapid cooling of the effluent, under pressure, immediately
after it exited the irradiation zone. Temperature was monitored immediately before
and after cooling. Variables such as internal volume of the plumbing within the mi-
crowave zone, flow rate and control of the applied microwave power allowed flex-
ible operation. To withstand corrosive acids and bases within the plumbing, contact
was avoided between metal surfaces and reaction mixtures. Feedback microproces-
sor control allowed setting of pump rates and temperatures for heating and cooling
of reactions. Failsafe measures were instituted to shut down the system if the tem-
2.2.2
The Microwave Batch Reactor (MBR)
2.2.3
Transfer of Microwave Energy
2.4
Contrasts between Synthesis and Digestion
Before the advent of the MBR and CMR for synthesis, microwave heating with pres-
sure vessels had been used in analytical laboratories, to speed the rate of digestion
and dissolution of solid samples, such as ores, hair and foodstuffs [32]. Digestion is
a degradative process, typically involving treatment of a small sample with an excess
of a strong oxidizing acid such as nitric, perchloric or sulfuric. The objective is to ob-
tain a clear solution usually for quantitative analysis of mineral components, without
charring and physical loss of material. Although it is essential that digestions be
standardized for high precision, parameters such as temperature, reaction time,
sample stirring and cooling rate are not essential to the outcome and so are not ac-
tively controlled.
As technologies for microwave-assisted digestion gained favor over slower, tradi-
tional heating methods, microwave equipment manufacturers emerged during the
1970s and 1980s to satisfy these analytical applications. The field of microwave-as-
sisted organic chemistry developed after microwave digestion had become an ac-
cepted analytical tool. Not surprisingly, some manufacturers attempted to employ di-
gestion technologies for synthesis. However, they did not appear to fully appreciate
the differences between the applications. This probably slowed the introduction of
appropriate equipment for synthesis, rather than assisting it.
In contrast with digestion, synthesis is a constructive process. Reactants that are
not particularly stable to strong acids and bases may be required and in much larger
amounts than those used for digestion. They may be expensive and highly reactive,
even on the laboratory-scale. The ability to define and control conditions including
temperature, time, sample stirring, addition or withdrawal of materials and postreac-
tion cooling, is almost always vital for satisfactory outcomes and for reproducibility.
Digestion techniques are unsatisfactory for these tasks.
2.5
Advantages of the MBR and CMR
The MBR and CMR have the following advantages for synthesis:
. reactions that are known to require high temperatures and higher boiling solvents
can be performed under pressure at these temperatures, but in lower boiling sol-
vents, facilitating workup;
. the available temperature range for many solvents is increased dramatically;
. low boiling reactants can be heated to high temperatures and rapidly cooled under
pressure; losses of volatile compounds are minimized;
. reactions can be sampled for analysis while material is being processed; with the
CMR, reaction mixtures can be subjected to multiple passes if required, or the
conditions can be altered during a run;
. reactions performed on a laboratory scale are amenable to scale up, because the
conditions are defined;
. moderate to high temperature reactions can be performed in vessels fabricated
from inert materials such as PFA Teflon, PTFE, or quartz; this is beneficial where
strongly acidic or basic reactants or products are incompatible with metals or boro-
silicate glass; and
. in multiphase systems, selective heating is possible.
2.6
Applications of the MBR and CMR
The CMR and MBR are enabling technologies that can stand alone. They also can be
integrated with other technologies and methodologies for processes that utilize re-
newable resources, limit waste and afford useful, recyclable products [6, 33]. They lie
at the heart of an approach toward development of new tools for preparative organic
chemistry, the aim being to be able to select appropriate combinations to solve speci-
fic problems in environmentally acceptable ways [6, 34]. This strategy has some simi-
larity with combinatorial chemistry, which revolutionized approaches to drug discov-
ery in the mid-1990s. However, it was designed more to establish multiple chemical
processes than to accomplish many reactions by parallel applications of the same
process. Downstream processing is a key element, as environmental damage can de-
stroy the viability of chemical manufacturing practices.
To complement the reactors, a catalytic membrane of Pd on porous glass tubing
was produced [35]. As mentioned in Sect. 2.8, this has been used to perform various
thermal reactions including Heck arylative coupling and to develop new reactions in-
cluding a novel coupling±dehydrogenation and a new concerted process incorporat-
ing an oxidation±dehydrogenation±double-Heck coupling [36]. Both reactions pro-
ceeded well in the MBR.
Other methodologies feature solvent-free conditions with ªneatº starting material,
tandem or cascade, catalyzed or uncatalyzed reactions, the use of aqueous media at
high temperature and nonextractive techniques for product isolation. Examples ap-
pear among the following microwave-assisted applications.
42 2 Microwave-assisted Organic Chemistry in Pressurized Reactors
2.6.1
Reactions with Sterically Constrained Molecules
Reactions that are unfavored through steric constraints are difficult to achieve con-
ventionally, but often proceed at higher than normal temperatures in the CMR or
MBR.
Heating of isopropanol under conventional reflux conditions, with 2,4,6-trimethyl-
benzoic acid and a catalytic amount of sulfuric acid, afforded the corresponding iso-
propyl ester in only 2 % conversion after 28 h. With the MBR, the product was iso-
lated in 56 % yield after 1 h at 148 8C [26] (Scheme 2.1 ± please note that in all
schemes herein, the use of a double headed arrow does not imply a balanced equa-
tion).
In the CMR, a yield of 81% was obtained after four passes (total residence time,
6.4 min) at 155±164 8C [22]. As indicated by the pressure, these reactions were per-
formed at temperatures below those enabling esterification by elimination of the al-
cohol to propene, followed by addition of the acid to the olefin. Kinetics studies de-
monstrated that the significantly enhanced yields obtained under the microwave
conditions, resulted from the higher temperatures employed and not from any in-
trinsic microwave ªeffectsº [37, 38].
Similarly, hydrolysis of tertiary amides of carboxylic acids is usually slow under
conventional conditions. Hydrolysis of a morpholide occurred in only 48 % yield
with 2M HCl at reflux after 4 h, yet proceeded in 70 % yield after only 10 min at
200 8C in the MBR (Scheme 2.2) [26]. The convenience of operation and the rapid
throughput enabled preparation of multiple batches of the corresponding acid in a
few hours.
2.6.2
Preparation of Thermally Labile Products
Vessels for microwave-assisted chemistry are usually made from thermal insulators
such as PEEK, quartz, borosilicate glass or PTFE. Thus, the benefits of rapid heating
can be diminished if the opportunity for workup is delayed by slow cooling. Decom-
position of thermally unstable products also can occur.
In the CMR, rapid cooling takes place through an in-line heat exchanger adjacent
to the microwave zone. Mixtures can be cooled immediately, while still under pres-
sure, to prevent losses of volatile compounds and to minimize decomposition of
thermally labile products [22]. The MBR has a cold-finger that contacts the reaction
mixture directly. Cooling can be initiated at any time during operation and is effi-
cient because it is not via the container [26]. Temperature and pressure monitoring,
as well as stirring, can be maintained during the process, allowing access to the ves-
sel at the earliest opportunity.
2.6.3
New Reactions that Require High Temperature
2.6.3.1 Etherification
Conditions employing elevated temperatures with less catalyst, a milder catalyst or
without addition of catalyst, can supplant those utilizing aggressive reagents at lower
temperatures. A recent example concerns a catalytic, thermal etherification that can
be performed near neutrality and that produces minimal waste [41]. This represents
a cleaner alternative to the traditional Williamson synthesis, in which the ether is
2.6 Applications of the MBR and CMR 45
This work elaborated upon that of Malwitz and Metzger [43] (who conducted the
majority of their investigations at 400 8C) and was performed by conventional heat-
ing. It further emphasizes opportunities for the discovery of new, clean reactions
that require high temperature.
2.6.4
Reactions Known to Require High Temperatures
sulted from the convenience in the operation of the reactor in comparison with a
standard autoclave, indicating broader opportunities for microwave heating to im-
prove the conditions for established reactions that require high temperatures.
2.6.5
Viscous Reaction Mixtures
Typically, viscous materials transfer energy poorly. With conductively heated vessels,
pyrolytic degradation on the walls can co-occur with incomplete reaction towards the
center of the container. Large thermal gradients can result in suboptimum conver-
sions, loss of product and laborious cleanup procedures. Also, when high tempera-
tures are required, heat losses increase and conductive heating becomes inefficient.
Under microwave conditions these problems are diminished. This accounts for the
numerous applications for curing of polymers by microwave heating.
Reactions such as the acid catalyzed preparation of isopropylideneglycerol pro-
ceeded well in the CMR. A yield of 84 % was obtained after a residence time of
48 2 Microwave-assisted Organic Chemistry in Pressurized Reactors
1.2 min, the exit temperature being 133 8C. A conventional literature method in
which the upper temperature was limited by the boiling point of acetone, required
one day for a comparable result [22].
2.6.6
Reaction Vessels
Being microwave-transparent, the reaction vessels will be no hotter than their con-
tents. As mentioned above, they usually are made from insulating polymeric materi-
als like polytetrafluoroethylene (PTFE), which have inherent advantages for cleaner
processing. In contrast with other materials, PTFE is resistant to attack by strong
bases or HF and is not corroded by halide ions.
This was significant in the preparation of 1,2-dimethyl-3-hydroxy-4-pyridone, em-
ployed clinically as an iron chelating agent. The aminoreductone is obtained by reac-
tion of methylamine with maltol. Traces of metal within the system readily form
highly colored complexes with reactant or product and these are difficult to remove.
With the CMR, the preparation was achieved in 65 % yield without the need for deco-
lorizing charcoal and the product was crystallized by collecting the effluent in acet-
one (Scheme 2.10) [22].
Another advantage of PTFE vessels is low adhesivity, which can help to reduce de-
tergent and organic solvent usage during cleaning operations that would otherwise
generate considerable effluent [6].
2.6.7
Reactions with a Distillation Step
With conductive heating, several factors militate against efficient distillation. Transf-
eral of heat to the liquid usually occurs from the inner surfaces of the vessel. Vapori-
zation from the surface and convection preclude a uniform temperature within the
liquid. To achieve distillation under those circumstances, the pot temperature must
be considerably higher than that of the distillate. With microwaves, energy is ab-
sorbed more uniformly, a larger volume of the sample in the pot is heated simulta-
neously, convection is reduced and distillation is more rapid.
The MBR can accommodate distillation through an outlet tube connected to a
port on the top flange. With that arrangement, a monodehydrobromination of 1,6-di-
bromohexane was performed in which the product 6-bromohex-1-ene was removed
from the mixture before a second molecule of HBr could be eliminated [26, 33].
2.6 Applications of the MBR and CMR 49
2.6.8
Miscellaneous Reactions
The MBR and CMR have been instrumental for transformations including nucleophi-
lic substitution, oxidation, addition, elimination, isomerization, esterification, trans-
esterification, hydrolysis, oximation, acetalization, amidation, decarboxylation and
coupling. Most of the common organic solvents have been employed. ªNameº reac-
tions include Michael, Diels±Alder and aza Diels±Alder additions,Williamson etherifi-
cation, Claisen and Rupe rearrangements, Finkelstein, Knoevenagel, Claisen±
Schmidt, Mannich, Baylis±Hillman,Willgerodt, Meyer±Schuster, reactions, Heck ary-
lative coupling and Hofmann elimination. Improved conditions obtained in compari-
son with literature methods, typically involved one or more of the following: increased
convenience, savings in time, higher yields, greater selectivity, the need for less cata-
lyst, or employment of a more environmentally benign solvent or reaction medium.
2.6.9
Kinetic Products
If a reaction can give multiple products, that obtained first (the kinetic product) is
not necessarily the most stable (the thermodynamic product). The kinetic product
will be in equilibrium with the starting materials but not necessarily with the ther-
modynamic product.
Where multiple products are possible, the CMR and MBR have been employed to
optimize conditions for formation of specific components of a reaction sequence. Ex-
amples discussed below, were obtained by heating organic substrates such as allyl
phenyl ether [46] and carvone [47] in water. Rearrangements, addition or elimination
of water and isomerizations occurred, with each transformation favored under
tightly defined conditions.
Heating of carvone in water at 210 8C for 10 min afforded 8-hydroxy-p-menth-6-
en-2-one, the equilibrium position favoring the starting material by a factor of 4 : 1.
The hydroxymenthenone, although in equilibrium with the starting material, also
underwent elimination of water and isomerization to carvacrol, giving a 1 : 1:1 mix-
ture of these three substances at 230 8C after 10 min. At 250 8C, for 10 min, carvone
isomerized to carvacrol almost quantitatively (Scheme 2.11) [47].
Scheme 2.12 Kinetic and thermodynamic products from allyl phenyl ether in water.
Allyl phenyl ether was heated with water in the MBR for 1 h at different tempera-
tures [46]. It underwent Claisen rearrangement to 2-allylphenol (56 % yield) at
200 8C, 2-(2-hydroxyprop-1-yl)phenol (37 % yield) at 230 8C and 2-methyl-2,3-dihydro-
furan (72 % yield) at 250 8C (Scheme 2.12). Support for the reaction sequence was ob-
tained through experiments with authentic intermediates.
2.6.10
Gaseous Reactants and Media
Even though the CMR and MBR operate under conditions in which pressure is de-
veloped, gaseous reactants or media often can be handled in these systems without
problems arising through over-pressure. Mannich reactions with dimethylamine,
Baylis±Hillman reactions with formaldehyde, aminoreductone formation with am-
monia, all proceeded without difficulty, as did Willgerodt reactions in which gases
are formed during the process.
In an extension, we recently reported the formation of cyclic carbonates respec-
tively from epichlorohydrin (Scheme 2.13) and styrene oxide, with carbon dioxide
under catalysis by KI in the presence of a crown ether [48].
2.7
High-temperature Water as a Medium or Solvent for Microwave-assisted
Organic Synthesis
At ambient temperature water is a poor solvent for most organic compounds. How-
ever, its ionic product increases one thousandfold between 25 8C and 240 8C, making
it a stronger acid and base. The dielectric constant decreases from 78 at 25 8C to 20
at 300 8C, indicating that the polarity is lowered with temperature increase. These
properties suggest that in organic reactions, water could have a role that varied with
temperature. The MBR and CMR were ideally suited to investigations into organic
synthesis in high-temperature water [6, 33, 47].
A diversity of reactions and high selectivities was obtained with seemingly minor
variations in the conditions. Scaleup also was achieved, including to continuous op-
eration with the CMR, confirming the potential of aqueous high-temperature media
for the development of clean processes [49]. Obvious advantages of water include
low-waste workup, low cost, negligible toxicity and safe handling and disposal. Some
examples have already been discussed and additional ones appear below.
2.7.1
Biomimetic Reactions
Naturally occurring monoterpene alcohols were heated in water without prior deriva-
tization with typical biological water-solubilizing groups such as phosphates or glyco-
sidic units. Biomimetic reactions that normally would be acid-catalyzed, proceeded
on the underivatized compounds in the absence of added acidulant. Cooling of the
mixtures rendered the products insoluble, readily isolable and the aqueous phase
did not require neutralization before workup.
Geraniol, nerol and linalool are practically insoluble in water at ambient tempera-
ture. Although acid labile, they do not readily react in water at moderate temperature
and neutral pH. In unacidified water at 220 8C in the MBR, they reacted within min-
utes. Geraniol rearranged to a-terpineol (18 %) and linalool (16 %) predominantly.
Lesser amounts of the monoterpene hydrocarbons were also obtained, including
myrcene, a-terpinene (10 %), limonene (11%), g-terpinene, the ocimenes, a-terpino-
lene and alloocimenes (Scheme 2.14) [50].
Nerol and linalool underwent considerably more elimination than did geraniol, to
give the same hydrocarbons. The major products and their relative proportions were
2.7.2
Indoles
In the first example of water as the reaction medium for Fischer indole synthesis,
2,3-dimethylindole was obtained in 67 % yield from phenylhydrazine and butan-2-
one, at 220 8C for 30 min (Scheme 2.15). Neither a preformed hydrazone nor addi-
tion of acid was required [33].
Interestingly, the Fischer indole synthesis does not easily proceed from acetalde-
hyde to afford indole. Usually, indole-2-carboxylic acid is prepared from phenylhy-
drazine with a pyruvate ester followed by hydrolysis. Traditional methods for decarb-
oxylation of indole-2-carboxylic acid to form indole are not environmentally benign.
They include pyrolysis or heating with copper±bronze powder, copper(I) chloride,
ªcopperº chromite, ªcopperº acetate or copper(II) oxide, in for example, heat-transfer
oils, glycerol, quinoline or 2-benzylpyridine. Decomposition of the product during
lengthy thermolysis or purification affects the yields.
In water at 255 8C, decarboxylation of indole-2-carboxylic acid was quantitative
within 20 min in the MBR (Scheme 2.16) [33].
A semisystematic study into the hydrolysis of ethyl indole-2-carboxylate in aqu-
eous media at high temperature, indicated that decarboxylation of the resultant acid
proceeded by an arenium ion mechanism and was inhibited by base. As base pro-
moted hydrolysis of the ester, it was possible to obtain either the acid or indole from
the ester merely by manipulating the equivalents of base present [51].
2.7.3
Reactions in Aqueous Acid and Base
2.7.4
Avoiding Salt Formation
2.7.5
Resin-based Adsorption Processes
From the foregoing, microwave heating in pressurized systems facilitates organic re-
actions in aqueous media. Normally, the products would be recovered by extraction
with organic solvent. The aqueous phase would become saturated with the organic
solvent (and the solvent with water), complicating disposal and offsetting environ-
mental benefits gained through the use of water as the reaction medium in the first
place.
To avoid this, we have employed hydrophobic resins for concentration and isola-
tion of the products from aqueous media [49]. Organics are retained on the resin
and subsequently can be desorbed with solvents such as ethanol, which is useful for
ªgreenº chemistry as it is readily recyclable, renewable and biodegradable. Nonex-
tractive processes offer convenience, can be conducted with high throughput and af-
ford low waste owing to ready disposal of the spent water, recyclability of the resin
and the solvent used for desorption.
2.8
Metal-catalyzed Processes
The first reports of the use of microwave heating to accelerate Heck, Suzuki and
Stille reactions on the solid phase [54] and in solution did not appear until 1996 [55].
Since then, many metal-catalyzed reactions have been performed within minutes by
microwave heating in pressurized systems, sometimes with high regio- and enan-
tioselectivity [56, 57].
Palladium, its complexes and salts can catalyze transformations including oxida-
tion, hydrogenation and rearrangement. One of the most useful applications of the
metal though, is for activation of C±H bonds towards coupling reactions. The Heck
reaction, which involves C±C coupling of an aryl or vinyl halide with an alkene in the
presence of palladium derivatives has been the subject of intensive study [58]. Syn-
thetic transformations of terminal alkynes via homo- or heterocouplings of the Gla-
ser, Eglington, or Chodkiewicz±Cadiot type have also attracted interest.
Stille coupling involves the use of tin reactants. Tin is both toxic and difficult to re-
move. In an elegant extension of the pioneering work of Horvath [59], Curran and
his coworkers prepared fluorous tin reactants that facilitated Stille reactions and en-
abled the convenient isolation and separation of products afterwards [60]. Probably
owing to low solubility of fluorine-containing compounds in organic solvents, the re-
actions normally required about one day at 80 8C. With microwave heating, they
were completed within minutes [61].
Typically, the Pd species for Heck couplings are homogeneous catalysts, stabilized
by air-sensitive ligands. They present economic and environmental problems regard-
ing separation, regeneration and reuse. These difficulties can be diminished with
heterogeneous catalysts that are more easily recoverable from the reaction mixture.
As mentioned in Sect. 2.6, a catalyst consisting of palladium metal deposited on por-
2.9 Pressurized Microwave Systems Developed by Others for Organic and Organometallic Chemistry 55
ous glass tubing was developed for C±C coupling reactions [35]. It was used for reac-
tions conducted continuously or batchwise and could be reused for repeat or differ-
ent reactions. Reactions were performed in the presence of air, with either conven-
tional heating or under microwave irradiation in the MBR.
Coupling of aryl halides with terminal acetylenes affords internal alkynes. Typi-
cally, high amounts of catalyst (1±5 mol % Pd) and coaddition of copper salts (also
1±5 mol % in Cu) are needed, thus reducing the industrial viability of such proce-
dures. With palladium on porous glass, copper salts or activating ligands were not
necessary. In contrast with Heck reactions involving halogen-containing reactants
(usually aryl bromides or iodides), homocoupling of terminal acetylenes occurred
readily and with excellent atom economy [35].
A new oxidation±dehydrogenation±Heck coupling catalyzed by Pd on porous glass
or Pd(OAc)2 also proceeded under microwave heating in the MBR (Scheme 2.18)
[36]. With an excess of PhI, saturated alcohols including 1-propanol and 3-phenylpro-
panol afforded 3,3-diphenylpropenal as the major product, and trans-2,3-diphenylpro-
penal by a concerted process taking 10 min at 220 8C. The yields of this remarkable
transformation were low, however, and further work into several aspects is required
to make them more respectable.
2.9
Pressurized Microwave Systems Developed by Others for Organic and Organometallic
Chemistry
This account has been written from a personal perspective and thus has focused al-
most exclusively on the development and applications of the CMR and MBR. In that
context, the numerous significant contributions of others are beyond the present
scope and have been discussed comprehensively by the author elsewhere [6, 33, 34].
Nonetheless, they shall not be allowed to pass here without mention.
56 2 Microwave-assisted Organic Chemistry in Pressurized Reactors
In their pioneering work, Gedye et al. used a domestic oven and commercially
available screw cap pressure vessels [7, 62]. The pressure was measured and tem-
perature was estimated through an infrared sensor, immediately after the power had
been turned off. The reactor of Baghurst and Mingos [63] employed a glass vessel
with a pressure-controlling device in a modified domestic oven. Pougnet and his
coworkers constructed a range of equipment mainly for analytical applications.
Among their developments was a pressure vessel [64] and a cylindrical microwave
cavity that housed containers for chemical reactions and distillation [65]. Abramo-
vitch et al. performed organic synthesis in a batch unit that incorporated a stirrer
and temperature control, but the details were not disclosed [66]. Majetich and Hicks
employed digestion bombs for 45 organic reactions in which a baristat was used to
indirectly maintain the temperature below 200 8C [67].
Chemat et al. have reported several microwave reactors, including systems that
can be used in tandem with other techniques such as sonication [68], and ultravio-
let radiation [69]. With the microwave±ultrasound reactor, the esterification of acetic
acid with n-propanol was studied along with the pyrolysis of urea. Improved results
were claimed compared with those from conventional and microwave heating [68].
The efficacy of the microwave±UV reactor was demonstrated through the rearran-
gement of 2-benzoyloxyacetophenone to 1-(2-hydroxyphenyl)-3-phenylpropan-1,3-
dione [69].
A specialized application of microwave-assisted organic synthesis involves the pre-
paration of radiopharmaceuticals labeled with short-lived radionuclides, particularly
for use in positron emission tomography [70±72]. This represented an excellent ap-
plication of microwave technology, where the products must be prepared quickly and
in high radiochemical yield, on a small scale.
Recent studies by Mingos and Whittacker into the optimum conditions have con-
firmed the benefits of pressurized systems for microwave chemistry [73].
2.10
Technical Considerations and Safety
As demonstrated above with examples, the CMR and MBR offer many advantages
for synthetic processes that benefit from rapid heating and cooling. These systems
are less useful and may be inappropriate when the reaction requires low tempera-
ture conditions throughout, when materials or reactions that are incompatible with
microwave energy (e. g. reactions involving predominantly nonpolar organics) are to
be employed or for reactions with ªdryº media.
With microwaves, high rates of heating are usually desired. A difficulty can arise,
however, when a material undergoing irradiation, possesses a dissipation factor that
increases with temperature. In contrast with conventional heating, microwave energy
is then absorbed more efficiently with temperature and the rate of temperature rise
accelerates. Although, this phenomenon can be beneficial for maintaining high tem-
peratures for catalysts, a thermal runaway can result unless the temperature is care-
fully monitored and the power controlled.
2.11 Conclusion 57
On the other hand, solvents usually show a decrease in dielectric constant with
temperature. Efficiency of microwave absorption diminishes with temperature rise
and can lead to poor matching of the microwave load, particularly as fluids approach
the supercritical state. Solvents and reaction temperatures should be selected with
these considerations in mind, as excess input microwave energy can lead to arcing.
If allowed to continue unchecked, arcing could result in vessel rupture or perhaps
an explosion, if flammable compounds are involved. Therefore it is important in mi-
crowave-assisted organic reactions, that the forward and reverse power can be moni-
tored and the energy input be reduced (or the load matching device adjusted) if the
reflected power becomes appreciable.
In the MBR, the applicator of plate steel was an important safety feature in the
possible event of vessel rupture or explosion. Temperature and pressure measure-
ments, stirring, infinitely variable control of microwave power input, the cold-finger,
as well as a pressure relief valve, have all contributed significantly to the safety and
reliability of the system.
Safety issues have been discussed comprehensively, in an earlier review [33]. At
time of writing, more than two thousand reactions have been conducted with a range
of solvents, times and temperatures in the CMR and MBR without accident or in-
jury.
2.11
Conclusion
Acknowledgments
The author is particularly indebted to the following dedicated coworkers, whose ef-
forts have helped to convert dreams into reality: A. F. Faux, T. Cablewski, R. W. Trai-
nor, K. D. Raner, J. S. Thorn, L. Bagnell, U. Kreher, L. Mokbel and F. Vyskoc.
1 Haggin, J., Chem. Eng. News 1996, 13 Stadler A., Kappe, C. O., J. Chem. Soc.,
74(23) 38. Perkin Trans. 2 2000, 1363.
2 Casumano, J. A., J. Chem. Educ. 1995, 14 Vidal, T., Petit, A., Loupy, A., Gedye,
72, 959. R. N., Tetrahedron 2000, 56, 5473.
3 Collins, T. J., J. Chem. Educ. 1995, 72, 15 Abramovitch, R. A., Org. Prep. Proced.
965. Int. 1991, 23, 683.
4 Tebo, P. V., Chemtech 1998, 28(3), 8. 16 Bose, A. K., Manhas, M. S., Banik,
5 Strauss, C. R., Scott, J. L., Chem. Ind. B. K., Robb, E. W., Res. Chem. Intermed.
(London) 2001, 610. 1994, 20, 1.
6 Strauss, C. R., Aust. J. Chem. 1999, 52, 17 Banik, B. K., Manhas, M. S., Kalu-
83. za, Z., Barakat, K. J., Bose, A. K., Tetra-
7 Gedye, R., Smith, F., Westaway, K., hedron Lett. 1992, 33, 3603.
Ali, H., Baldisera, L., Laberge L., 18 Bose, A. K., Manhas, M. S., Ghosh, M.,
Rousell, J., Tetrahedron Lett. 1986, 27, Shah, M., Raju,V. S., Bari, S. S., Newaz,
279. S. N., Banik, B. K., Chaudhary, A. G.,
8 Giguere, R. J., Bray, T. L., Duncan, S. Baraket, K. J., J. Org. Chem. 1991, 56,
M., Majetich, G., Tetrahedron Lett. 6968.
1986, 27, 4945. 19 Banik, B. K., Manhas, M. S., Newaz,
9 Loupy, A., Bram, G., Sansoulet, J., S. N., Bose, A. K., Bioorg. Med. Chem.
New J. Chem. 1992, 16, 233. Lett. 1993, 3, 2363.
10 Loupy, A., Petit, A., Hamelin, J., 20 Stull, D. R., Ind. Eng. Chem. 1947, 39,
Texier-Boullet, F., Jacquault, P., 517.
Math, D., Synthesis 1998, 1213. 21 Strauss, C. R., Chem. Aust. 1990, 57,
11 Langa, F., De la Cruz, P., De la 186.
Hoz, A., Diaz-Ortiz, A., Diez-Bar- 22 Cablewski, T., Faux, A. F., Strauss,
ra, E., Contemp. Org. Synth. 1997, 4, C. R., J. Org. Chem. 1994, 59, 3408.
373. 23 Peterson, C., New Scientist 1989,
12 Varma, R. S., Green Chem. 1999, 1, 43. 123(1681), 18.
References and Footnotes 59
24 Bagnell, L., Bliese, M., Cablewski, T., 43 Malwitz, D., Metzger, J. O., Chem.
Strauss, C. R., Tsanaktsidis, J., Aust. J. Ber. 1986, 119, 3558.
Chem. 1997, 50, 921. 44 Strauss, C. R., Trainor, R. W., Org.
25 Braun, I., Schulz-Ekloff, G., Prep. Proc. Int. 1995, 27, 552.
Wohrle, D., Lautenschlager,W., 45 Cablewski, T., Gurr, P. A., Pajalic,
Micropor. Mesopor. Mater. 1998, 23, 79. P. J., Strauss, C. R., Green Chem. 2000,
26 Raner, K. D., Strauss, C. R., Trainor, 2, 25.
R. W., Thorn, J. S., J. Org. Chem. 1995, 46 Bagnell, L., Cablewski, T., Strauss,
60, 2456. C. R., Trainor, R. W., J. Org. Chem.
27 Manufactured by Personal Chemistry, 1996, 61, 7355.
Uppsala, Sweden. 47 An, J., Bagnell, L., Cablewski, T.,
28 Metaxis, A. C., Meredith, R. J., Indus- Strauss, C. R., Trainor, R. W., J. Org.
trial Microwave Heating, Peregrinus, Chem. 1997, 62, 2505.
London, 1988. 48 Kreher, U., Strauss, C. R.,
29 Gabriel, C., Gabriel, S., Grant, E. H., Walther, D., ECSOC-5, Electronic Conf.
Halstead, B. S. J., Mingos, D. M. P., Synth. Org. Chem., Sept. 1±30, 2001;
Chem. Soc. Rev. 1998, 27, 213. www.mdpi.net/ecsoc-5 [4e0041].
30 Neas, E. D., Collins, M. J. in: Introduc- 49 Bagnell, L., Bliese, M., Cablewski, T.,
tion to Microwave Sample Preparation, Strauss, C. R., Tsanaktsidis, J., Aust. J.
Kingston, H. M., Jassie, L. B. (eds.), Chem. 1997, 50, 921.
American Chemical Society, Washing- 50 Strauss, C. R., Trainor, R. W. in: Bio-
ton, DC, 1988, Chapter 2, pp. 7±32. technology for Improved Foods and Fla-
31 Schiffmann, R. F. in: Encyclopedia of vors; ACS Symposium Series, 637,
Chemical Processing and Design,Vol. 30, Takeoka, G. R., Teranishi, R., Wil-
McKetta, J. J., Cunningham, W. A. liams, P. J. (eds.), ACS,Washington,
(eds.), Marcel Dekker, New York, 1989, 1996, pp. 272±281.
p. 220. 51 Strauss, C. R., Trainor, R. W., Aust. J.
32 Kingston, H. M., Jassie, L. B. (eds.), Chem. 1998, 51, 703.
Introduction to Microwave Sample Pre- 52 Sheldon, R. A., Chem. Ind. (London)
paration, American Chemical Society, 1997, 12.
Washington, DC, 1988. 53 Sheldon, R. A., Chemtech, 1994, 24(3),
33 Strauss, C. R., Trainor, R. W., Aust. J. 38.
Chem. 1995, 48, 1665. 54 Larhed, M., Lindeberg, G., Hall-
34 Strauss, C. R. in: Handbook of Green berg, A., Tetrahedron Lett. 1996, 37,
Chemistry and Technology, Clark, J., Mac- 8219.
quarrie, D. (eds.), Blackwell, London, 55 Larhed, M., Hallberg, A., J. Org.
2002. Chem. 1996, 61, 9582.
35 Li, J., Mau, A. W.-H., Strauss, C. R., 56 Vallin, K. S. A., Larhed, M., Johans-
Chem. Commun. 1997, 1275. son, K., Hallberg, A., J. Org. Chem.
36 Bagnell, L., Kreher, U., Strauss, 2000, 65, 4537.
C. R., Chem. Commun. 2001, 29. 57 Bremberg, U., Larhed, M., Mo-
37 Raner, K. D., Strauss, C. R., J. Org. berg, C., Hallberg, A., J. Org. Chem.
Chem. 1992, 57, 6231. 1999, 64, 1082.
38 Raner, K. D., Strauss, C. R.,Vyskoc, F., 58 Beletskaya, I. P., Cheprakov, A. V.,
Mokbel, L., J. Org. Chem. 1993, 58, 950. Chem. Rev. 2000, 100, 3009.
39 Danner, H., Braun, R., Chem. Soc. Rev. 59 Horvath, I. T., Acc. Chem. Res. 1998,
1999, 28, 395. 31, 641.
40 Cundy, C. S., Collect. Czech. Chem. 60 Curran, D. P., Hoshino, M., J. Org.
Commun. 1998, 63, 1699. Chem. 1996, 61, 6480.
41 Bagnell, L., Cablewski, T., Strauss, 61 Larhed, M.; Hoshino, M.; Hadi-
C. R., Chem. Commun. 1999, 283. da, S.; Curran, D. P., Hallberg, A.,
42 Bagnell, L., Strauss, C. R., Chem. J. Org. Chem. 1997, 62, 5583.
Commun. 1999, 287. 62 Gedye, R. N., Smith, F., Westaway,
K. C., Can. J. Chem. 1988, 66, 17.
60 2 Microwave-assisted Organic Chemistry in Pressurized Reactors
3
Nonthermal Effects of Microwaves in Organic Synthesis
Laurence Perreux and Andr Loupy
1200
1000
annually
number of publications
800 cumulative
600
400
200
0
86
87
88
89
90
91
92
93
94
95
96
97
98
99
00
19
19
19
19
19
19
19
19
19
19
19
19
19
19
20
year
and reliable data resulting from strict comparisons of reactions performed under si-
milar conditions (reaction medium, temperature, time, pressure) except for micro-
wave irradiation or conventional heating [3, 8, 9]. A monomode microwave reactor
should preferably be used, because this enables wave focusing (reliable homogeneity
of the electric field) and accurate control of the temperature (by use of optical fiber
or infrared detection) throughout the reaction [3, 9]. This makes it possible to use
both types of activation with similar temperature-profile increases. On the basis of
such strict comparisons it becomes possible to make an educated judgment about
the suitability, or otherwise, of microwave irradiation depending on reaction type
and experimental conditions.
3.1
Origin of Microwave Effects
a) b) c)
Scheme 3.1 Effects of the surrounding to a continuous electric field; and (c) sub-
electric field on the mutual orientation of di- mitted to an alternating high frequency elec-
poles: (a) without constraint; (b) submitted tric field.
3.2 Specific Microwave Effects 63
tain free conduction electrons. This phenomenon is essential for heating of solid par-
ticles, more or less magnetic, such as a variety of mineral oxides or metallic species.
For liquid products (solvents), only polar molecules selectively absorb microwaves,
because nonpolar molecules are inert to microwave dielectric loss. In this context of
efficient microwave absorption it has also been shown that boiling points can be
higher when solvents are subjected to microwave irradiation rather than conven-
tional heating. This effect, called the ªsuperheating effectº [13, 14] has been attribu-
ted to retardation of nucleation during microwave heating (Tab. 3.1).
Tab. 3.1 Boiling points (8C) of some polar solvents under the action of MW irradiation in the ab-
sence or presence of a nucleation regulator.
It is clearly connected with the effect of stirring and the presence of a nucleation
regulator [15]. It is also related to the microwave power. It has been shown that the
effect is eliminated when the experiments are performed on well-stirred mixtures
[16, 17] using low microwave power. It could be essentially a consequence of the ab-
sence of stirring, i. e. in closed vessels inside a domestic microwave oven.
3.2
Specific Microwave Effects
The origin of specific microwave effects is twofold ± those which are not purely ther-
mal and a special thermal effect connected with possible intervention of ªhot spotsº.
One of the few theoretical papers trying to explain acceleration under the action of
microwaves has recently been published by A. Miklavc [18]. He stated that large in-
creases in the rates of chemical reactions occur because of the effects of rotational ex-
citation on collision geometry. This could be cautiously considered when one has
knowledge of the quasi-nil energy involved by microwave interaction according to
Planck's law [E = hc/l = 0.3 cal/mol].
Non-purely thermal effects (other than simple dielectric heating) can be foreseen
to have multiple origins. These effects can be rationalized by consideration in terms
of the Arrhenius law [19, 20] and can result from modification of each of the terms
of this equation.
k A exp
DG7=RT
64 3 Nonthermal Effects of Microwaves in Organic Synthesis
Calculations have shown that faster diffusion rates might be explained by an in-
crease in the factor A with no change in activation energy.
A decrease in the activation energy DG7 is certainly a major effect. Because of the
contribution of enthalpy and entropy to the value of DG7 (= DH7 ± TDS7), it might
be predicted that the magnitude of the ±TDS7 term would increase in a microwave-
induced reaction, because organization is greater than with classical heating, as a
consequence of dipolar polarization.
Lewis et al. [22] presented experimental evidence for such an assumption after
measurements of rate constants at different temperatures for the unimolecular imi-
dization of a polyamic acid (Eq. (1), Fig. 3.2, and Tab. 3.2).
Eq. (1)
-2,00
-3,00
ln k
-4,00
-5,00 microwave
thermal
-6,00
-7,00
1000/T (K)
Fig. 3.2 First-order kinetic plots for microwave (MW) and ther-
mal (D) activation of the imidization reaction.
3.3 Effects of the Medium 65
MW 57 ± 5 13 ± 1
D 105 ± 14 24 ± 4
The apparent activation energy is evidently substantially reduced. The same expla-
nation, i. e. a decrease in DG7, was also proposed for the decomposition of sodium
hydrogen carbonate in aqueous solution [23].
Intervention of localized microscopic high temperatures is possible [8, 14, 24], as
advocated in sonochemistry to justify the sonochemical effect. There is an inevitable
lack of experimental evidence, because we can necessarily have access to macro-
scopic temperature only. It has been suggested [6, 19] that, in some examples, MW
activation could originate from hot spots generated by dielectric relaxation on a mo-
lecular scale.
3.3
Effects of the Medium
Microwave effects should also be treated according to the reaction medium. Solvent
effects are of particular importance [25, 26].
If polar solvents are used, either protic (e. g. alcohols) or aprotic (e. g. DMF,
CH3CN, DMSO etc), the main interaction might occur between MW and polar mole-
cules of the solvent. Energy transfer is from the solvent molecules (present in large
excess) to the reaction mixtures and the reactants, and it would be expected that any
specific MW effects on the reactants would be masked by solvent absorption of the
field. The reaction rates should, therefore, be nearly the same as those observed un-
der the action of conventional heating (D).
This is essentially true, as is evidenced by the rates of esterification in alcoholic
media of propan-1-ol with ethanoic acid [27] or of propan-2-ol with mesitoic acid [28].
The absence of a specific microwave effect became apparent from several experi-
ments carefully conducted in alcohols or in DMF under similar conditions but with
microwave or classical heating [7].
More recently [29] the microwave-mediated Biginelli dihydropyrimidine synthesis
(Eq. 2) was reinvestigated using a purpose-built commercial microwave reactor with
on-line temperature, pressure, and microwave power control. Transformations per-
formed with microwave heating at atmospheric pressure in ethanol solution resulted
in neither a rate increase nor an increase in yield when the temperature was identi-
cal to that used for conventional thermal heating. The only significant rate and yield
enhancements were found when the reaction was performed under solvent-free con-
ditions in an open system.
66 3 Nonthermal Effects of Microwaves in Organic Synthesis
Eq. (2)
A rapid and efficient procedure for flash heating by microwave irradiation has
been described for attachment of aromatic and aliphatic carboxylic acids to chloro-
methylated polystyrene resins via their cesium salts (Eq. 3) [17].
Eq. (3)
Significant rate accelerations and higher loadings are observed when the micro-
wave-assisted and conventional thermal procedures are compared. Reactions times
are reduced from 12±48 h with conventional heating at 80 8C to 5±15 min with mi-
crowave flash heating in NMP at temperatures up to 200 8C. Finally, kinetic compari-
son studies have shown that the observed rate enhancements can be attributed to
the rapid direct heating of the solvent (NMP) rather than to a specific nonthermal
microwave effect [17].
The synthesis of b-lactams from diazoketones and imines can be realized not only
by using photochemical reaction conditions but also under the action of microwave
irradiation. When the reaction was performed in o-dichlorobenzene at 180 8C, how-
ever, the rates of thermal and microwave-assisted formations of b-lactams were
shown to be identical within the limits of experimental error (80±85 % conversion
after 5 min) [30].
As described above, however, some rather small differences could be observed,
taking into account the superheating effect of the solvent under the action of micro-
waves in the absence of any stirring. This probably occurs in the isomerization of sa-
frole and eugenol in ethanol under reflux [31] (MW 1 h, D 5 h to obtain equivalent
yields).
Superheating of the solvent was believed to be responsible of the observed rate en-
hancement under microwave irradiation in the synthesis of 3,5-disubstituted 4-amino-
1,2,4-triazoles when conducted in 1,2-ethylene glycol as (polar) solvent (Eq. 4) [32].
Eq. (4)
The yields obtained by use of MW and D were clearly different for very short reac-
tion times and became similar after 15 min at 130 8C (Fig. 3.3).
3.3 Effects of the Medium 67
80
yield (%)
60
40
classical heating
microwave irradiation
20
0
0 10 20 30
time (min)
More interesting is the use of nonpolar solvents (e. g. xylene, toluene, carbon tetra-
chloride, hydrocarbons), because these are transparent to MW and absorb them only
weakly. They therefore enable specific absorption by the reactants. If these reactants
are polar, energy transfer occurs from the reactants to the solvent and the results
might be different under the action of MW and D. This effect seems to be clearly de-
pendent on the reaction and is, therefore, the subject of controversy. In xylene under
reflux, for example, no MW-specific effects were observed for the Diels±Alder reac-
tion [5] whereas important specific effects were described for aryldiazepinone synth-
esis [33].
These examples will be discussed and explained later, during discussion of the de-
pendence of MW effects on reaction mechanisms.
Clearly, the effect of the solvent seems to be of great importance to the possibility
of MW-specific effects. These could decrease when the polarity of the solvent is in-
creased. This effect was shown in at least two studies by Berlan et al. [19] and, more
recently, by Bogdal [34]. In the first study acceleration of a nonsymmetric Diels±Al-
der reaction under the action of MW was much more apparent in xylene than in the
more polar dibutyl ether (Fig. 3.4).
In the second investigation [34], involving a coumarin synthesis by Knoevenagel
condensation, supported by rate constant measurements and activation energy cal-
culations, it was found that the effect of MW was more important when the reaction
was conducted in xylene ± it was noticeably reduced in ethanol (Eq. (5) and Tab. 3.3).
Eq. (5)
Microwave effects are most likely to be observed under solvent-free reactions [3].
In addition to the preparative interest of these methods in terms of use, separation,
and economical, safe and clean procedures, absorption of microwave radiation
68 3 Nonthermal Effects of Microwaves in Organic Synthesis
Xylene
100
80
Yield (%)
60
40
classical heating
20 microwave
0
0 5 10 15 20 25 30
Time (h)
Dibutyl ether
100
80
Yield (%)
60
40 classical heating
microwave
20
0
0 5 10 15 20 25 30
Time (h)
Tab. 3.3 Rate constants (× 103) for the reaction depicted in Eq. (5).
a) b)
∆ MW
should now be limited only to the reactive species. The possible specific effects will
therefore be optimum, because they are not moderated or impeded by solvents.
They can be accomplished by three methods [3, 35]:
These procedures coupled with microwave activation have proven beneficial and
have led to many success stories which are described in several reviews [3, 4, 39±42].
Some apparently specific effects can, however, arise from the supports. Mineral
supports are usually poor heat conductors, i. e. significant temperature gradients can
develop inside the vessels under the action of conventional heating, whereas they be-
have as efficient absorbers of microwave energy with consequently more tempera-
ture homogeneity (Scheme 3.2).
3.4
Effects of Reaction Mechanisms
Microwave effects result from material±wave interactions and, because of the dipolar
polarization phenomenon, the greater the polarity of a molecule (such as the solvent)
the more pronounced the microwave effect when the rise in temperature [43] is con-
sidered. In terms of reactivity and kinetics the specific effect has therefore to be con-
sidered according to the reaction mechanism and, particularly, with regard to how the
70 3 Nonthermal Effects of Microwaves in Organic Synthesis
-
A + B A
δ+ δB
TS
∆G∆
∆GMW < ∆G ∆
∆GMW
GS
polarity of the system is altered during the progress of the reaction. These assump-
tions are evidently connected to the Hughes±Ingold model [44] universally adopted to
explain solvent effects [45] and especially the intervention of aprotic dipolar solvents.
Specific microwave effects can be expected for polar mechanisms, when the polar-
ity is increased during the reaction from the ground state towards the transition state
(as more or less implied by Abramovich in the conclusion of his review in 1991 [42]).
The outcome is essentially dependent on the medium and the reaction mechanism.
If stabilization of the transition state (TS) is more effective than that of the ground
state (GS), this results in enhancement of reactivity as a result of a decrease in the ac-
tivation energy (Fig. 3.5), because of electrostatic (dipole±dipole type) interactions of
polar molecules with the electric field.
3.4.1
Isopolar Transition-state Reactions
Isopolar activated complexes differ very little or not at all in charge separation or
charge distribution from the corresponding initial reactants. These complexes are
formed in pericyclic reactions such as Diels±Alder cycloadditions and the Cope rear-
rangement.
The polarity of ground and transition states are a priori identical, because no
charges are developed during the reaction path. Following this rule, specific micro-
wave effects would not be expected for these reactions, as has been verified when the
reactions were performed in a nonpolar solvent [5, 6]. Solvent effects in these reac-
tions are also small, or negligible, for the same reasons (Fig. 3.6) [46].
Such a conclusion is, nevertheless, connected with the synchronous character of
the mechanism. If a stepwise process is involved (nonsimultaneous formation of the
two new bonds), as for unsymmetric dienes and/or dienophiles or in hetero Diels±
Alder reactions, a specific microwave effect could intervene, because charges are de-
veloped in the transition state. This could certainly be so for several cycloadditions
[47, 48] and particularly for 1,3-dipolar cycloadditions [49]. Such an assumption has
3.4 Effects of Reaction Mechanisms 71
O O O
TS
GS ∆GMW ∆G∆
=
CH3NHCH2CO2H/HCHO
C 70 C70 N CH3 Eq. (6)
MW or ∆
Eq. (7)
Eq. (8)
72 3 Nonthermal Effects of Microwaves in Organic Synthesis
Tab. 3.4 Thermal or microwave activation for the cycloaddition depicted in Eq. (7).
CH3 P(O)(OEt)2 D 20 90 5
MW 20 90 78
H CO2Et D 5 100 70
MW 5 100 92
C6H5 CO2Et D 10 160 >98
MW 10 160 >98
H C6H5 D 30 120 40
MW 30 120 >98
H CH2OH D 30 100 40
MW 30 100 >98
a
The ratio of the amounts of the two isomers formed was identical for both activation conditions.
Tab 3.5 Thermal or microwave activation for the cycloaddition depicted in Eq. (8).
D Toluene 24 h 110 65
MW None 3 min 119 98
D None 3 min 119 64
D None 30 min 119 98
3.4.2
Bimolecular Reactions between Neutral Reactants Leading to Charged Products
δ+ δ-
N + R X N R X N+ R, X-
δ+ δ-
P + R X P R X P+ R, X-
δ+ δ- + -
N + O N O N C O
3.4.3
Anionic Bimolecular Reactions Involving Neutral Electrophiles
- + δ- δ-
Nu , M + R X Nu R X, M+
- -
δ δ
Nu -, M+ + O Nu O, M+
Scheme 3.4
74 3 Nonthermal Effects of Microwaves in Organic Synthesis
. If tight ion pairs (between two hard ions) are involved in the reaction the micro-
wave-accelerating effect then becomes more important, because of enhancement
of ionic dissociation during the course of the reaction as tight ion pairs (GS) are
transformed into more polar loose ion pairs (TS).
. If, on the other hand, loose ion pairs (between soft ions) are involved, microwave
acceleration is limited, because ionic interactions are only slightly modified from
GS to TS.
This duality in behavior of some SN2 reactions can be foreseen and observed (vide
infra) by comparing reactions involving hard or soft nucleophilic anionic reagents ac-
cording to the cation and the leaving group.
3.4.4
Unimolecular Reactions
3.5
Effect of the Position of the Transition State Along the Reaction Coordinates
The position of the transition state along the reaction coordinates in relation to the
well-known Hammond postulate [53] will now be considered. If the activation en-
ergy, DG7, of a reaction is only small the TS looks like the GS (it is depicted as a ªre-
actant-like transition stateº). Consequently, the polarity is only slightly modified be-
tween the GS and TS during the course of the reaction and only weak specific micro-
wave effects can be foreseen under these conditions.
δ+ δ-
R X R X
δ-
C X C X
+
N N δ
TS
TS
∆G
GS
∆G
(a) GS (b)
Scheme 3.6 (a) Small DG v early TS v little change in polarity
TS/GS v weak microwave effect. (b) Large DG v late TS v impor-
tant change in polarity TS/GS v large microwave effect.
3.6
Effect on Selectivity
The literature contains a few examples of increased selectivity [56±58] in which the
steric course and the chemo- or regioselectivity of reactions can be altered under the
action of microwave irradiation compared with conventional heating.
As a further consequence of these assumptions, it might be foreseen that micro-
wave effects could be important in determining the selectivity of some reactions.
When competitive reactions are involved, the GS is common for both processes. The
mechanism occurring via the more polar TS could, therefore, be favored under the
action of microwave radiation (Scheme 3.7).
Langa et al. [26, 59, 60], while conducting the cycloaddition of N-methylazo-
methine ylide with C70 fullerene, proposed a rather similar approach. Theoretical
calculations predict an asynchronous mechanism, suggesting that this phenomenon
can be explained by considering that, under kinetic control, ªmicrowave irradiation
will favor the more polar path corresponding to the hardest transition stateº.
3.7
Some Illustrative Examples
To illustrate these trends, we now present some typical illustrative examples. These
have been selected because strict comparisons of microwave and classical heating ac-
tivation were made under similar conditions (time, temperature, pressure, etc. ¼)
for the same reaction medium and using, preferably, a monomode system equipped
with stirring. They mostly involve reactions performed under solvent-free conditions
or, occasionally, in a nonpolar solvent, because these conditions are also favorable for
observation of microwave effects.
3.7.1
Bimolecular Reactions between Neutral Reactants
These reactions are among the most propitious for revealing specific microwave ef-
fects, because the polarity is evidently increased during the course of the reaction
from a neutral ground state to a dipolar transition state.
+ -
δ δ + -
N + O N O N O products
Eq. (9)
dipolar transition state
This example covers very classical processes such as syntheses of a wide variety of
compounds including imines, enamines, amides, oxazolines, hydrazones, etc ¼
Amines
R
R R - N R1
δ R2 = H
O O R'
R'
R'
+
δ
R1
N
R N R2
N H R2
H R2 R1
R2 H
R1 R'
Eq. (10)
Eq: 11
It has been shown that a mixture of an indoloquinazoline and anthranilic acid, ad-
sorbed on graphite, led to cyclization in good yields after 30 min at 140 8C, in less
time than for the purely thermal procedure under similar conditions, when a very
poor yield is obtained even after 24 h [63].
A large specific microwave effect was observed in the solvent-free synthesis of N-
sulfonylimines, a similar type of reaction [64] (Eq. 12).
Ar Ar δ-
O O
H Ar
H
+ N SO2Ar'
δ
N - H2 O H
N H SO 2Ar'
H SO2Ar' H
Eq:
12
H
Hydrazone synthesis
In a typical example, a mixture of benzophenone and hydrazine hydrate in toluene
resulted in 95 % yield of the hydrazone within 20 min [65] (Eq. 13).
78 3 Nonthermal Effects of Microwaves in Organic Synthesis
Ph Ph δ−
O O
Ph Ph
Ph
δ + N NHR
N - H2O Ph
N H NHR
H NHR H
Eq. (13)
H
The hydrazone was subsequently treated with KOH under the action of MW to un-
dergo Wolff±Kishner reduction (leading to PhCH2Ph) within 25±30 min in excellent
yields (95 %). As an extension, the reaction of neat 5- or 8-oxobenzopyran-2(1H)-ones
with a variety of aromatic and heteroaromatic hydrazines is substantially accelerated
by irradiation in the absence of any catalyst, solid support, or solvent [66] (Eq. 14).
NHR
O N
O O O O
RNHNH2
Eq. (14)
NHCOPh NHCOPh
R R −
δ
O O
HO
HO
+
PhCH2NH3 , RCO2- δ+
N
N H CH2Ph
H CH 2Ph H
H
R
O
- H2O PhCH2 N
H
Eq. (15 a)
3.7 Some Illustrative Examples 79
Tab. 3.6 Reaction of benzylamine with carboxylic acids at 150 8C for 30 min.
Ph 1:1 10 10
1.5 : 1 75 17
1 : 1.5 80 8
PhCH2 1:1 80 63
1.5 : 1 93 40
1 : 1.5 92 72
CH3(CH2)8 1:1 85 49
If it is considered that at 150 8C water can be removed equally under both types of
activation, the noticeable difference in yields is clearly indicative of an improvement
in the nucleophilic addition of the amine to the carbonyl group when performed un-
der the action of microwave irradiation, with important specific effects.
The preparation of aliphatic, aromatic, or functionalized tartramides directly from
tartaric acid and amines under solvent-free conditions and microwave activation was
very recently described [67 b].
Eq. (15 b)
Synthesis of 2-oxazolines
Oxazolines can be readily synthesized by means of a noncatalyzed solvent-free proce-
dure by two successive nucleophilic additions on a carbonyl group with the forma-
tion of an amide as an intermediate [68] (Eq. 16).
80 3 Nonthermal Effects of Microwaves in Organic Synthesis
Ar
O Ar δ−
O O
HO HO Ar
+ HN OH
N Nδ OH - H2O
OH
H H OH
H HO
H OH
OH
HO HO
δ−
O Ar
Ar O δ+
C OH
HN
N - H2O
OH
OH
HO
OH
Synthesis of aminotoluenesulfonamides
These compounds were prepared by reacting aromatic aldehydes with sulfonamides
under the action of microwaves in the presence of a few drops of DMF to enable bet-
ter energy transfer [69] (Eq. (17) and Tab. 3.7).
Ar
Ar δ− SO2NH 2
O O
H
H H 3C
δ+ - H2O
N SO2 NH2
N SO2NH 2 N
H Ar
H H
H CH3
CH3 Eq. (17)
p-NO2C6H4 1 40 98 5
o-ClC6H4 2 22 90 6
5-NO2-2-furyl 2 20 96 5
3.7 Some Illustrative Examples 81
R1 R1
O + HCONH2 + HCO2H C NHCHO
R2 R2
R1 R1 δ-
O O
R2
R2
products
δ+
N
N H CHO
H CHO H
Eq. (18)
H
Ph Ph 202 MW >98
202 D 2
p-CH3OPh p-CH3OPh 193 MW 95
193 D 3
Ph CH2Ph 210 MW 95
210 D 12
S S
additive (3eq.)
RCHO + H N N H + S8 N N
R R
Eq. (19)
DMF 5min 100-102°C MW 85% ∆ 48%
EG 5min 105-107°C MW 80% ∆ 40%
Alcohols
CH3 CH3
HOTs (10%)
H 35C 17CO2H + HO-(CH2)2 CH C 17H35CO2-(CH2)2 CH
CH3 CH3
Eq. (20)
1min 20 90°C MW 98%
1min 20 120°C ∆ 13%
10min 120°C ∆ 57%
The very important specific microwave effect is consistent with the mechanism
which involves the formation of a dipolar TS from neutral molecules (Scheme 3.8).
H H
+
R O R Oδ
R' -
R' δ
O O
H O H O
Scheme 3.8
δ−
S O S O - H2O S O
HO HO RO
+
O Oδ
R H R H Eq. (21)
CH3
+ 130°C
CH3COCH2CO2Et
N OH N O O
MW 12 min 62 %
∆ 390 min 62 %
Eq. (22)
Eq. (23)
84 3 Nonthermal Effects of Microwaves in Organic Synthesis
Tab. 3.9 Synthesis of ketal from cineole ketone and propylene glycol (R1 = CH3 R2= H).
Alumina MW 30 78
D 300 27
Toluene MW 15 90
D 360 30
N N
+ CO2Et
N N
H CH2 CH2CO2Et
Eq. (24)
montmorillonite Li+ 1 min 40 °C MW 40% ∆ 0%
5 min 75 °C MW 72% ∆ 2 7%
It was shown that microwave irradiation accelerated the 1,4 Michael addition of
primary and cyclic secondary amines to acrylic esters, leading to several b-amino
acid derivatives in good yields within short reaction times [78] (Eq. 25).
H
N δ+ OR
CO 2R + N H N
δ- O CO2R
δ+ δ-
N R Br N R Br N+
N N R
N Br -
H H H
R = C 6H5CH2 CH2
Eq. (26)
H
N
OH
N
N N δ+
Ph MW 3 min Ph N
Ph
O H via
N Oδ −
N OH
N dipolar TS
MW 6 min Ph
N
Eq. (27)
N-alkylation of 2-halopyridines
A microwave- (focused waves) assisted procedure for N-alkylation of 2-halopyridines
has been described; the noticeable microwave effect was indicative of a polar TS [82]
(Eq. 28).
δ+
N Cl N Cl
+N Cl
R Cl -
R CH2
CH2 CH2
R
δ−
Cl Cl Eq. (28)
X
X
+
δ
X NH Cl
N + HCl
R Cl + δ
−
N
H
R
R
Eq. (29)
R = CHO, NO2 X = CH2 , O
Tab. 3.10 SNAr reaction between p-chlorotoluene and piperidine (X = CH2, R = CH3).
K2CO3±EtOH D 16 h Reflux 60
MW 6 min Reflux 70
Basic alumina MW 75 s ±a 92
a
Undetermined, but certainly a very high temperature because the vessel was placed inside an alumina
bath (prone to microwave absorption).
+
Ph −
δ δ + -
PPh3 + PhCH2 Cl Ph 3P C Cl Ph 3PCH2Ph, Cl
H H
100 78 24
150 94 91
Although the microwave effect is not appreciable at 150 8C, it becomes clearly ap-
parent when the temperature is decreased to 100 8C. When delineating microwave
effects, careful attention needs to be paid to the temperature level. If this is too high,
the microwave effect will be masked and the temperature has to be minimized in or-
der to start from a low yield under the action of D and therefore have the possibility
of observing microwave enhancements.
This conclusion is in agreement with the kinetic results from Radoiu et al. [86] ob-
tained for the transformation of 2- and 4-tbutyl phenols in the liquid phase in the
presence of montmorillonite KSF as catalyst either under the action of MW or D
(Eq. (31) and Tab. 3.12).
OH OH OH OH
KSF
+ + +
MW or ∆
Eq. (31)
Tab. 3.12 Rate constants (r8) for the transformation of 2-tbutyl phenols under the action of MW or
D according to temperature.
3.7.2
Bimolecular Reactions with One Charged Reactant
The TS for anionic SN2 reactions involves loose ion pairs as in a charge delocalized
(soft) anion. On the another hand, the GS could involve a neutral electrophile and
either tight or loose ion pairs depending on the anion structure (hard or soft) (Eq. 32).
δ- δ-
Nu-, M+ + R X Nu R X, M+ Nu R + M+, X-
Eq. (32)
88 3 Nonthermal Effects of Microwaves in Organic Synthesis
Eq. (33)
Tab. 3.13 Reaction of KOtBu with 2-ethoxyanisole in the presence of 18-crown-6 and, optionally,
ethylene glycol.
± 20 min MW 120 7 0 90
± 20 min D 120 48 0 50
E.G. 1h MW 180 0 72 23
E.G. 1h D 180 98 0 0
Demethylation results from the SN2 reaction whereas de-ethylation occurred via
the E2 mechanism (Scheme 3.9).
Eq. (34)
Tab. 3.14 Yield (%) from dialkylation of dianhydrohexitols under PTC conditions.
RX t (min), T (pC) A B C
MW D MW D MW D
PhCH2Cl 5, 125 98 13 98 15 97 20
nC8H17Br 5, 140 96 10 74 10 95 10
These observations are consistent with the reactive species being constituted from
tight ion pairs between cations and the alkoxide anions resulting from abstraction of
hydrogen atoms in A, B and C (Scheme 3.11).
90 3 Nonthermal Effects of Microwaves in Organic Synthesis
Scheme 3.11
Eq. (35)
Tab. 3.15 Yield (% E) from reaction of monobenzylated isosorbide (D) with ditosylates for 15 min.
(CH2)8 95 91 96 39
(CH2)6 91 90 96 45
CH2CH2OCH2CH2 92 92 91 36
Eq. (36)
3.7 Some Illustrative Examples 91
H 8 138 96 <2
Et 15 160 94 <2
n-Bu 20 167 89 <2
n-Hex 20 186 87 <2
n-Hex 60 186 ± 22
n-Hex 180 186 ± 60
O O H
-
δ δ-
+
O C Br, M
R
H CH3
Scheme 3.12 Transition state for the Krapcho
reaction.
Anionic b-elimination
Ketene acetal synthesis by b-elimination of haloacids from halogenated acetals under
well controlled conditions using thermal activation (D), ultrasound (US) or micro-
wave irradiation [92] (MW) has been described. From a mechanistic point of view, as
the TS is more charge delocalized than the GS and the polarity is enhanced during
the course of the reaction, a favorable microwave effect can therefore be observed
(Eqs. (37) and (38) and Scheme 3.13).
Ph Ph
O KO tBu (2 eq) O
Eq. (37)
O Br NBu4Br (5%) O
nBu S t
KO Bu (2 eq) nBu S
Eq. (38)
nBu S NBu4Br (5%) nBu S
-
δ
Br Br
-
δ
H H +
M
-
- + δ
tBuO M tBuO
K 2CO3
Z CO2H + nC 8H17Br Z CO2nC8H17
NBu4Br
Eq. (39)
Z = CN 2 min 202 °C 95 %
Conversely, when n-octyl bromide was used with the less reactive terephthalate
species, which constitutes a ªslow-reacting systemº, the yield was raised from 20 to
84 % under the action of microwaves compared with D, which can be attributed to a
later TS along the reaction coordinates (Eq. 40).
K2CO 3
HO2C CO 2H + RBr RO2C CO2R
NBu4Br
Eq. (40)
R = nC8H17 6 min 175 °C MW 84% ∆ 20%
3.7 Some Illustrative Examples 93
PhCH2CH2Br N
N
N + KOH - N N
- H2O N
H CH2CH2Ph
K+
Eq. (41)
loose ion pair
8 min 145 °C MW 64 % ∆ 61 %
This effect could be predicted when considering the weak evolution of polarity be-
tween the GS and TS as the reactive species consist of loose ion pairs (involving a
soft anion).
Eq. (42)
94 3 Nonthermal Effects of Microwaves in Organic Synthesis
Tab. 3.17 C-alkylation of b-naphthol in the presence of lithiated base under solvent-free conditions.
R R Yield (%)
MW D
Ph Me 92 73
Ph n-Oct 98 86
Me 90 48
n-Oct 97 39
O O
Eq. (44)
Tab. 3.19 Transesterification in 15 min at 160 8C in basic medium with methyl benzoate and
laurate (monomode reactor, relative amounts 1 : 2 : 2).
R Yield (%)
MW D
Ph 96 21
CH3(CH2)10 88 0
The reactive species under these conditions consist of tight ion pairs involving the
alkoxide anion from the carbohydrate (charge localized anion). The less reactive long
chain methyl laurate leads to a later TS along the reaction coordinates and the mag-
nitude of the microwave effect is therefore increased.
Eq. (45)
Ph ± 0 0
KOtBu 80 22
KOtBu 87 70
336 87 70
PhCH2 ± 63 6
KOtBu + Aliquat 336 63 36
96 3 Nonthermal Effects of Microwaves in Organic Synthesis
Eq. (46)
As expected, the tighter ion pair (RNH±, K+) exhibits the larger microwave effect.
As an extension of this work, and in order to gain further insight, the effect of amine
substituents was studied [99] in the reaction with ethyl benzoate (Eq. (47) and
Tab. 3.21).
Eq. (47)
Tab. 3.21 Ethyl benzoate aminolysis with different amines at 150 8C for 10 min (relative amounts
PhCO2Et : RNH2 : KOtBu = 1.5 : 1 : 2).
R Yield (%)
Without NR4Cl With NR4Cl
MW D MW D
Ph 88 73 90 83
PhCH2 84 42 98 85
n-Oct 80 22 87 70
The microwave effects are clearly substituent dependent and, as in the former ex-
ample, disappear by adding a phase transfer agent. When the substituent R is able to
delocalize the negative charge on the amide anion (R = Ph), the ion pairs RNH±, M+
exist in a looser association. Consequently, a decrease in microwave effect is expected
as the evolution from the GS v TS occurs with only a slight modification of polarity
in the ion pairs. Conversely, the microwave effect is optimal with the tighter ion pair
(nOctN±, K+).
In a parallel study [100], it was shown that formamide, primary and secondary
amines react with esters in the presence of potassium tert-butoxide under the action
of microwave irradiation. Substituted amides are formed in yields (generally more
than 70 %) much higher than under the action of conventional heating (Eq. (48) and
Tab. 3.22).
3.7 Some Illustrative Examples 97
Eq. (48)
Tab. 3.22 Ester aminolysis with different amines under the action of microwave irradiation during
3 min and with conventional heating.
FeCl 3 (5%)
MeO + PhCOCl COPh + HCl
MW or ∆ MeO
Eq. (49)
The reactive species is the acylium ion resulting from abstraction of a chloride an-
ion from benzoyl chloride (Eq. 50). This reagent comprises an ion pair formed be-
tween two large (soft) ions which are therefore associated as loose ion pairs. Accord-
ing to these assumptions, the absence of a microwave effect should be expected as
the polarity evolution is very weak between the GS and TS (two loose ion pairs of si-
milar polarities).
Eq. (50)
98 3 Nonthermal Effects of Microwaves in Organic Synthesis
This conclusion is also in agreement with the results obtained for the sulfonyla-
tion of mesitylene in the presence of metallic catalysts such as FeCl3 [102] (5 %)
(Eq. 51)
SO2Ph
FeCl 3 (5%)
+ PhSO2Cl + HCl
Eq. (51)
A nonthermal microwave effect was not observed when identical temperature gra-
dients were produced by classical heating and microwave irradiation and if the reac-
tion temperature was strictly controlled.
CHO
POCl3-DMF
Cl COCH3 Cl
SiO2
Et
Eq. (52)
Cl
Cl
Cl P O δ-
Cl P O
Cl
Cl
O
O
+
N δ
N
3.7 Some Illustrative Examples 99
Ph
+ δ+ δ+
Ph3P + PhCH 2 NEt3, Cl- Ph3P C NEt3 , Cl-
H H Eq. (53)
This effect is readily attributable to the very loose structure of the ion pairs in the
TS (additionally involving delocalization in the phenyl groups of the phosphine) and
which are therefore far more polar than the initial ion pairs in the GS.
3.7.3
Unimolecular Reactions
Scheme 3.14
100 3 Nonthermal Effects of Microwaves in Organic Synthesis
Analysis of the kinetic parameters showed that the apparent activation energy for
the reaction was reduced from 105 to 57 kJ mol±1 (Tab. 3.2). This observation is con-
sistent with the polar mechanism of this reaction implying the development of a di-
pole in the transition state (Fig. 3.8) even when the reaction was performed in a polar
solvent.
O
O
N δ+
H NH
HO C HO
O δ−
O
H
R1 NHCOCF 3 K10 clay (R2)R1 N
CF 3 Eq. (54)
R2 NH 2 (R1)R2 N
H H 125 87 23
H CH3 127 84 19
NO2 H 134 95 28
This observation is consistent with the assumptions of the authors predicting mi-
crowave effects when the polarity is enhanced in a dipolar TS. The kinetic rate-deter-
mining step consists of an intramolecular attack of the nitrogen lone pair on the car-
bon atom of the carbonyl moiety (Scheme 3.15).
3.7 Some Illustrative Examples 101
H H
-
δ
N O N O
CF 3 N δ
+ CF 3
N H H
H H
GS TS
Scheme 3.15 Mechanism of cyclocondensation.
CHO
RNHNH2
N Eq. (55)
cat PTSA N
N Cl N
R
CHO
RNHNH 2 NHR
N N
rapid slow N
N Cl N Cl N
(SN Ar)
R
N N +
δ
NHR δ
-
NHR
N Cl N Cl
GS dipolar TS
R = H, Ph
Scheme 3.16 Mechanism of internal SNAr.
102 3 Nonthermal Effects of Microwaves in Organic Synthesis
Eq. (56)
+
N Ar Nδ Ar
δ-
Scheme 3.17 Mechanism of
O intramolecular Michael addition
O
of an amino group.
A similar study has been reported with ortho-hydroxychalcones in dry media on si-
lica gel [107]. Conventional thermal cyclization, under the same conditions as for mi-
crowave irradiation, required a much longer reaction period (Eq. (57), Tab. 3.24, and
Scheme 3.18).
Eq. (57)
Tab. 3.24 Intramolecular Michael addition of o-hydroxychalcones during 20 min at 140 8C.
H H MW 82
D 44
CH3O H MW 61
D 22
3.8 Illustrative Examples of the Effects of Selectivity 103
H
δ+
R1 OH Ph R1 O Ph
-
δ
R2 R2
O O
GS dipolar TS
O R1 K10 O
Eq. (58)
Ph O R2 Ph OH
This effect can also get again be rationalized via a mechanism with the interven-
tion of a dipolar TS (Scheme 3.19).
O O
δ- δ+
Ph O Ph O
GS dipolar TS
Scheme 3.19 Mechanism of deprotection of allyl esters.
3.8
Illustrative Examples of the Effects of Selectivity
Very few results on selectivity effects are available due to a lack of strict comparisons
between microwave and D activation and in which kinetic details of reactions have
been described. Reports of the effect of microwaves on selectivity up to 1997 have
been reviewed by Langa et al. [57].
104 3 Nonthermal Effects of Microwaves in Organic Synthesis
3.8.1
Benzylation of 2-Pyridone
R R R R
RX
+ + +
N O N O N O N O N O
H R H H H
N-alkyl
Eq. (59)
C-alkyl
To justify these results, it may be assumed that the TS leading to C-alkylation will
be more polar than that responsible for N-alkylation. This assumption, however, pre-
sumes the existence of kinetic control (which is not ensured in this case).
3.8.2
Addition of Vinylpyrazoles to Imine Systems
Eq. (60)
3.8 Illustrative Examples of the Effects of Selectivity 105
The use of microwave irradiation as an energy source is crucial to conduct the re-
action and to avoid the decomposition or dimerization of the starting pyrazole,
which are observed in the absence of microwave irradiation.
A similar conclusion has been drawn during an examination of the Diels±Alder re-
action of 6-demethoxy-b-dihydrothebaine with methylvinylketone using microwave
irradiation [110]. When performed under conventional heating conditions, extensive
polymerization of the dienophile was observed whereas reaction is much more clea-
ner under microwave activation (Eq. 61).
Eq. (61)
3.8.3
Stereo Control of b-Lactam Formation [111, 112]
Formation of b-lactams by the reaction of an acid chloride, a Schiff base and a ter-
tiary amine (Eq. 62) appears to involve multiple pathways, some of which are very
fast at higher temperatures. When conducted in open vessels in unmodified micro-
wave ovens, high level irradiation leads to preferential formation of the trans b-lac-
tams (55 %) whereas, at low power, the cis isomer was obtained as the only product
(84 %). The failure of the cis isomer to isomerize to the trans compounds is an exam-
ple of induced selectivity.
Eq. (62)
This effect has recently been explained by considering that under the action of mi-
crowave irradiation the route involving direct reaction between the acyl chloride and
the imine competes efficiently with the ketone±imine reaction pathway, a situation
highlighted by theoretical calculations [113]. In the other words and according to our
assumptions, it can be stated that the transition state leading to the trans isomer
therefore seems to be more polar than that leading to the cis compound.
106 3 Nonthermal Effects of Microwaves in Organic Synthesis
3.8.4
Cycloaddition to C70 Fullerene
CH3 CH3
CH3 N N
N
Tab. 3.25 Yields of monoadducts and isomer distribution for cycloaddition (Eq. 63) in ODCB at
180 8C.
MW 120 W 30 39 50 50 ±
180 W 30 37 45 55 ±
300 W 15 37 47 53 ±
D 120 32 46 46 8
3.8.5
Selective Alkylation of 1,2,4-Triazole
PhH 2C Br-
N N N
+
This observation may be explained by the increased efficiency of the first benzyla-
tion (SN2 reaction between two neutral reagents proceeding via a dipolar TS) under
microwave conditions.
As an extension towards azolic fungicides, phenacylation was next examined. Un-
der the action of microwave irradiation, exclusive reaction in position 1 (or equiva-
lent 2) occurred whereas mixtures of N1, N4 and N1, 4 products were obtained by D
under the same conditions [115] (Eq. (65) and Tab. 3.26).
Eq. (65)
Cl 25 140 MW 90 100 ± ±
D 98 33 29 38
Cl 20 140 MW 95 100 ± ±
D 98 38 27 37
Br 24 170 MW 90 100 ± ±
D 98 38 28 34
As kinetic control was ensured, this clear microwave effect could possibly be due to
a difference in polarity of the transition states, with apparently a more polar TS being
formed when p attack by the nitrogen atom in position 2 is concerned. Theoretical cal-
culations are in progress at the present time to try to confirm this assumption.
108 3 Nonthermal Effects of Microwaves in Organic Synthesis
3.8.6
Rearrangement of Ammonium Ylides
CN
+ 2 min
Ph N CN Ph +
N N
CN
A (1,2) B (2,3)
Eq. (66)
MW yield 69% A : B = 70 : 30
∆ yield 56% A : B = 10 : 90
Under similar profiles of raising in temperature, it was shown that the selectivity
favoring 1,2 Stevens rearrangement is exemplified under the action of microwaves.
A tentative explanation can be to consider that, under the action of radiation, the
more polar mechanism (1,2 ionic shift) is favored when compared to less polar one
(2,3 radical shift). Maybe this result is indicative of a competition between ionic and
radical pathways.
3.9
Concerning the Absence of Microwave Effects
The absence of a microwave effect can result from at least three different origins:
(a) a similar polarity of the transition state when compared to the ground state. This
is the situation for synchronous mechanisms in some pericyclic reactions when
performed in nonpolar solvents [5, 6] or in neat liquids (Eqs. 67 and 68).
CO2Et
CO2Et
CO 2Et
toluene Eq. (67)
+
MW or ∆
CO 2Et
3.10 Conclusions 109
(b) a very early transition state along the reaction coordinates (cf. Hammond postu-
late) which cannot allow the development of polarity between the GS and TS (re-
actant-like). This will occur when the reactions are rather easy and do not require
classically harsh conditions. This would be true for phthalimide syntheses by re-
acting phthalic anhydride and amino compounds [117], and chalcone syntheses
by reacting aromatic aldehydes and acetophenones [118], etc. ¼ Slight differ-
ences can appear when performing the reaction in the presence of a solvent due
to a superheating effect if no stirring is used.
(c) a too high temperature level, which may produce good yields in short reaction
times under the action of conventional heating. In order to find evidence of mi-
crowave effects, it is necessary to reduce the temperature under conventional
conditions in order to start from a rather poor yield (< 30±40 %) to appreciate pos-
sible microwave activation. These cases have been revealed in some studies
where a microwave effect appeared at relatively low temperatures but are masked
at higher temperatures where yields of conventionally heated reactions are ele-
vated [85, 86, 89].
3.10
Conclusions
of ketones from aldehydes and terminal alkenes by Wilkinson complex [Rh (I) com-
plex], a serious improvement was evidenced under the action of microwave when
compared to classical heating (Eq. (69) and Tab. 3.27).
Eq. (69)
C6H5 n-C8H17 30 95 30
C6H5 CH2C6H5 10 84 30
n-C6H13 n-C8H17 10 61 30
This microwave effect is consistent with the fact that the rate-determining step
may be certainly the generation of aldimine A from aldehydes and 2-amino 3-pico-
line. It is in agreement with the transition state develops a dipole and, consequently
more polar than the ground state. To ascertain this hypothesis, it was effectively next
shown that the reaction involving reaction with preformed aldimine A, no revealed
any microwave effect [119].
References
4
Organic Synthesis using Microwaves in Homogeneous Media
Richard N. Gedye
4.1
Introduction
Although the ability of microwaves (MW) to heat water and other polar materials has
been known for half a century or more, it was not until 1986 that two groups of re-
searchers independently reported the application of MW heating to organic synth-
esis. Gedye et al. [1] found that several organic reactions in polar solvents could be
performed rapidly and conveniently in closed Teflon vessels in a domestic MW oven.
These reactions included the hydrolysis of amides and esters to carboxylic acids, es-
terification of carboxylic acids with alcohols, oxidation of alkyl benzenes to aromatic
carboxylic acids and the conversion of alkyl halides to ethers.
Most of these reactions occurred much faster using MW irradiation (at increased
pressure and temperature) than under conventional heating (ambient pressure, re-
flux), with rate enhancements ranging from 5 to 1200 times.
Shortly after these results were published, Giguere and coworkers reported dra-
matic reductions in reaction times in other MW-assisted syntheses, including Diels±
Alder, Claisen rearrangements and ene reactions [2]. These reactions were also per-
formed at elevated pressures, but sealed glass vessels (inside a bath packed with ver-
miculite) were used rather than Teflon.
It is to be noted that the reactions mentioned so far were performed under homo-
geneous conditions and, in most cases, using polar solvents, which are efficient ab-
sorbers of MW energy. Rate enhancements were attributed to the superheating of
the solvent due to the elevated pressures generated in the closed vessels.
During the 15 years since these reports, a large number of synthetic applications
of MW in organic. synthesis have been reported and these have been summarized in
a number of reviews [3±10].
The use of dry media (solvent-free) conditions, in which the reactants are ab-
sorbed on inert solid supports, in MW-heated reactions, has received a consider-
able amount of attention recently and has been used in the synthesis of a wide
range of compounds [11±16]. These reactions generally occur rapidly and the
method avoids hazards, such as explosions, associated with reactions in solvents
in sealed vessels in which high pressures may be generated. Also the removal of
solvents and their disposal are not required making the technique more environ-
mentally friendly.
MW heated reactions in homogeneous media, using either neat reagents or in the
presence of solvents, may also be performed at atmospheric pressure. This approach
has been used particularly by Bose et al. [17]. (MORE Chemistry), who reported, for
example, the rapid synthesis of heterocycles [18] in open vessels. Another approach,
which avoids hazards due to the flammability of solvents, is to perform the reactions
under reflux in a MW oven, which is modified to allow the reaction vessel to be at-
tached to a reflux condenser outside the MW oven [7, 19]. It should be pointed out,
however, that most of the available evidence shows that rate enhancements of MW
heated reactions in homogeneous media at atmospheric pressure are small or non-
existent [19]. This will be discussed in more detail later in this review (see also
Chapt. 5 of this book).
4.2
Reactions at Elevated Pressures
with pressure release valves, designed to vent when the pressure approaches an un-
safe level.
A continuous MW reactor (CMR), which operates by passing a reaction mixture
through a pressurized tubular microwave-transparent coil and a MW batch reactor
(MBR), have been developed by CSIRO in Australia and are used for organic synth-
esis on the laboratory scale [8]. The CMR can be operated at pressures up to
1400 kPa and temperatures up to 200 8C and the MBR at pressures and temperatures
up to 10 MPa and 260 8C.
Milestone [23] have produced a range of MW reactor systems for organic synth-
esis, including a quartz or ceramic MW reactor (MRS) for high pressure (up to
4 MPa) and temperature reactions, designed for large volume batch synthesis and a
multiple batch reactor MPR/HPR for up to 12 vessels, with volumes 2±270 mL for
operation at 3.5±10 MPa.
These instruments, designed by CSIRO and Milestone, include, in addition to
pressure and temperature measurement and control, a number of other features al-
lowing for greater safety and reproducibility of reaction conditions, such as stirring
to minimize temperature gradients, rapid cool-down at the end of the heating period
and energy shut-down if temperatures or pressures exceed safe levels.
The pressure generated in a reaction vessel, and hence the rate enhancement, de-
pends on a number of factors including the MW power level, the volatility of the sol-
vent, the dielectric loss of the reaction mixture, the size of the vessel and the volume
of the reaction mixture [7, 20]. Gedye et al. [20] found that, in the esterification of
benzoic acid with a series of aliphatic alcohols (Scheme 4.1) in closed Teflon vessels,
the most dramatic rate enhancements were observed with methanol (the most vola-
tile solvent).
The rate enhancement for the esterification of benzoic acid with methanol was
close to 100, when compared with the classical heating under reflux. On the other
hand, the rate enhancement for the esterification with n-pentanol, using the same
power level (560 W) was only 1.3. The approximate reaction temperature was almost
the same for the two alcohols (134 8C and 137 8C respectively). It should be noted,
however, that the rate enhancement for the esterification in pentanol increased to
6 times when a higher power level (630 W) was used, the reaction temperature being
higher (162 8C).
It is interesting to note that in the case of the reaction with pentanol (boiling point
138 8C), when the temperature of the MW reaction and the conventional reflux reac-
tion were about the same, there was very little difference in the reaction rates, sug-
gesting that the reaction rate depends on the reaction temperature rather than the
method of heating.
The effect of the size of the reaction vessel on the rate of MW-heated SN2 reaction of
4-cyanophenoxide ion 1 and benzyl chloride 2 in methanol (Scheme 4.2) was investi-
gated by reacting identical amounts of reagents in Teflon vessels of different sizes.
It was found that there was an inverse relationship between the size of the vessel
and the reaction rate. In each case the time required for the reaction to go to 65 % com-
pletion was determined. It was shown that the rate of the reaction increases as the
pressure increases, since the pressure is inversely related to the volume of the vessel.
The heating rate, and hence the rate of pressure increase, also depends on the vo-
lume of the reaction mixture [20]. When the volume is small, the pressure increases
as the volume of the reaction mixture increases. However at a certain volume this
trend is reversed and a larger volume heats more slowly. For example when water
was heated in a 150-mL Teflon vessel, the greatest rate of pressure increase occurs
when 15 mL water are heated using a power of 560 W. The volume at which this
maximum heating rate, however, varies for different solvents. For example it occurs
at 20 mL for 1-propanol.
Using this information, it was possible to optimize the reaction conditions to
achieve a particularly high rate enhancement. The rate enhancement of the esterifi-
cation of benzoic acid with 1-propanol (Scheme 4.1) was increased from 18 to
60 times when the volume was increased from 10 mL to 20 mL at 560 W and in-
creased further to 180 times by increasing the power level to 630 W.
In this early work, pressure was measured by connecting the cap of a pressure-re-
lease Savillex [24] Teflon vessel to a pressure gauge outside the MW oven and the fi-
nal temperature estimated by pointing an infrared sensor at the mixture in the vessel
immediately after the heating was completed.
MW heating under pressure has been used to accelerate the rates of a wide variety
of reactions, which occur only slowly under conventional heating. This technique
has been particularly useful in accelerating Diels±Alder and related reactions. For
example the reaction of anthracene 3 with dimethyl fumarate 4 in p-xylene
(Scheme 4.3), gave an 87 % yield of the adduct in only 10 min by MW heating using
a sealed tube as the reaction vessel [2]. In contrast, the control reaction, performed
under reflux in p-xylene (bp 138 8C required 4 h to give a 67 % yield. It is to be noted
that the reaction temperature of the MW heated reaction was estimated to be in the
range 325±361 8C, approximately 200 8C higher than the control reaction.
Dimethyl formamide (DMF) has been found to be a particularly useful solvent in
these reactions because it absorbs MW irradiation strongly, has a relatively high boil-
4.2 Reactions at Elevated Pressures 119
ing point and is miscible with water, making it relatively easy to remove from the re-
action mixture. Majetich and Hicks [6] showed that the reaction of 1,4-diphenylbuta-
diene 5 with diethyl acetylenedicarboxylate 6 under reflux in DMF (Scheme 4.4) at
153 8C gave a 67 % yield of the Diels±Alder adduct 7 in 6 h. When the same reaction
was performed by use of MW heating in a CEM microwave system using Teflon ves-
sels with pressure and temperature monitoring (196 8C and 200 kPa), a comparable
yield of 7 was obtained in only 20 min.
The ortho-Claisen rearrangement of phenyl vinyl ether 8 to 2-vinylphenol 9
(Scheme 4.5) is a very slow reaction, requiring 80 h of reflux in DMF to give a yield
of only 34 %. Using the CEM microwave system an 80 % yield was obtained in 5 h at
196 8C and 200 kPa [6].
The Alder±Bong reaction of cyclohexene with diethyl acetylenedicarboxylate 6 in
DMF under reflux (Scheme 4.6) gives a 14 % yield of the tricyclic compound 10 in
Br OMe
MeOH
acting the bromide with methanol using the MBR at 170 8C gave a 75 % yield of 16
in 20 min [8].
4.3
Reactions at Atmospheric Pressure
4.4
Effect of Microwaves on the Rates of Homogeneous Reactions in Open Vessels
There have been a number of reports which claim that MW heating causes signifi-
cant rate increases in some reactions in homogeneous media in open vessels, based
on comparisons between rates of reactions performed by MW heating and conven-
tional heating at the same temperature. This has led to the suggestion that specific
or nonthermal activating effects are responsible for these rate enhancements. How-
ever, some of these reactions have been re-examined and shown to occur at the same
rate, or only slightly faster, under MW irradiation.
4.4.1
Diels±Alder reactions
The first report suggesting specific activation of an organic reaction by MW was that
of Berlan et al. [28] who observed that some Diels±Alder reactions occurred more ra-
pidly on MW heating than under conventional heating at the same temperature
(95 8C). The reactions were performed in two different solvents, xylene and dibutyl
ether and the rate enhancements were slightly higher in xylene, the less polar sol-
vent. For example the rate enhancement of the reaction of 2,3-dimethyl-1,3-buta-
diene 21 with methyl vinyl ketone 22 was 8 times in xylene and 2.3 times in dibutyl
ether, based on the half lives of the reactions. Reaction of anthracene 3 with diethyl
maleate 23 in xylene (Scheme 4.12) resulted in an approximately fourfold rate in-
crease.
124 4 Organic Synthesis using Microwaves in Homogeneous Media
In a subsequent paper [32], however, Berlan himself cast doubt on the existence of
nonthermal effects, attributing the observed rate increases to localized hot-spots in
the reaction mixture or to superheating of the solvent above its boiling point. He
also mentioned the difficulty of measuring the temperature accurately in MW cav-
ities. Furthermore, kinetic studies by Raner et al. [33], showed that the Diels±Alder
reaction of 3 with 23 (Scheme 4.12) occurred at virtually the same rate under MW
and conventional heating at the same temperature.
4.4.2
Reactions of Biologically Important Molecules
Sun et al. [34] reported that the rate of hydrolysis of the biomolecule ATP under MW
irradiation was 25 times faster than under classical heating at similar temperatures.
However, the same research group [35] later observed that, with more accurate tem-
perature control, the hydrolysis rates were in fact almost identical.
There have also been reports [36, 37] that racemization of amino acids occurs
more rapidly using MW heating than conventional heating at the same temperature.
Chen et al. [36] observed that racemization of amino acids in acetic acid the presence
of benzaldehyde was accelerated by MW heating. Lubec et al. [37] reported that some
d-proline and cis-4-hydroxy-d-proline were found in samples of infant milk formula
when they were heated in a MW oven. On the other hand, conventionally heated
samples did not contain these unnatural d-amino acids. This report caused concern,
and received media attention because d-proline is neurotoxic and suggested that
MW heating of some foods could have deleterious effects on their nutritional value
and the health of the consumer.
However, Marchelli et al. [38]. were unable to detect any d-proline when milk for-
mula containing l-proline was heated in open vessels in a MW oven. Westaway and
Gedye [30] showed that very small amounts of d-proline (0.1±0.2 %) were produced
when an aqueous solution of l-proline was heated under reflux in a MW oven for
4.4 Effect of Microwaves on the Rates of Homogeneous Reactions in Open Vessels 125
15 min. Although d-proline was not detected when the same solution was heated un-
der reflux conventionally for the same time, approximately 1% of d-proline was pro-
duced after 24 h of conventional reflux. MW did not have a significant effect on the
rate of racemization of l-proline in acetic acid in the presence of benzaldehyde [30].
The small increase in racemization rate observed when an aqueous solution of l-pro-
line was heated under reflux on a MW oven at atmospheric pressure could be attribu-
ted to localized superheating or a generalized superheating of the solvent. It is
known that water superheats by 4±10 8C when boiled in a MW oven [39, 40].
4.4.3
Other Reactions in Polar Solvents
(hot spots) may also contribute. For this reason, these modest rate enhancements
should not necessarily be taken as evidence for nonthermal effects since rate in-
creases of less up to about 20 times can be explained by superheating, assuming re-
action rates double for every 10 8 rise in temperature. Mingos [47] has suggested that
rate enhancements should only be taken as evidence for nonthermal effects if they
are 100 times or more, but this includes the longer time required to reach the reac-
tion temperature by conventional heating.
Comparison between reaction rates using MW and conventional heating should
preferably be performed using kinetic studies [19, 33, 48], where rate constants can
be evaluated or plots of yield of product versus reaction time for the conventional
and MW reactions can be compared.
Bose et al. [18] reported that the Knoevenagel reaction between salicylaldehyde 27
and ethyl malonate 28 in the presence of piperidine gave a high yield of 3-carbethox-
ycoumarin 29 in 3 min in an open vessel in a MW oven (Scheme 4.15).
In their initial studies on this reaction, Westaway and Gedye [30] observed slightly
increased rates both in the presence of ethanol and in the absence of a solvent, when
MW heating was used in an open vessel. Since it was difficult to measure the tem-
perature accurately in the MW experiment, the reaction in ethanol was later per-
formed under reflux [19]. The yield of the product was found to be 73 % in 8 min
using MW reflux compared with a yield of 71% in 20 min under conventional reflux,
representing a rate enhancement of approximately 2.5 times.
The rate of the Knoevenagel reaction of acetophenone 30 with ethyl cyanoacetate
31 (Scheme 4.16) was observed to be considerably slower and it was therefore found
to be more suitable for rate studies [19].
A mixture of 30 and 31 and a small amount of acetic acid in ethanol was heated to
reflux in a domestic MW oven. A catalytic amount of piperidine was then added and
the mixture heated under reflux at low power (180 W). Heating was interrupted at
regular time intervals and aliquots taken for G.C. analysis. A plot of yield of the pro-
duct versus reaction time showed that the same yield of the mixture of isomers 32
and 33 (17 %) was obtained after heating under reflux for 120 min in the MW oven
and 320 min of conventional reflux, and hence the rate enhancement was 2.5±3.0
times.
The isomer ratio E : Z (33 : 32) was 1.2±1.6 over the course of the reaction and was
not affected by the heating method, so it was concluded that MW irradiation did not
affect the stereochemical outcome of this reaction, presumably due to the reaction
being under thermodynamic control.
It is interesting to note that when the same reaction was performed using a vari-
able frequency MW system [49] with temperature control at 80 8C in the absence of a
solvent, it occurred at the same rate as a similar reaction heated conventionally at
the same temperature. The use of variable frequency provides very uniform heating,
minimizing the possibility of hot spots. Thus it can be concluded that the modest
rate enhancement observed in ethanol under reflux was because of hot spots or to a
general superheating of the solvent. Again, it should be emphasized that these mod-
est MW rate enhancements should not be taken as hard evidence for nonthermal
MW effects.
Substantial MW rate enhancements have been reported in the Biginelli synthesis
of dihydropyrimidines [50, 51] under homogeneous conditions. The synthesis in-
volves a one-pot cyclocondensation of a b-ketoester with an aryl aldehyde and urea or
thiourea in the presence of a catalytic amount of HCl in ethanol solution. An exam-
ple of this synthesis is shown in Scheme 4.17.
The reactions were performed in an open beaker using a domestic MW oven, and
reaction times were reduced from 2±24 h of conventional reflux to 3±11 min under
MW irradiation.
Stadler and Kappe [52] re-examined this synthesis using a Milestone ETHOS 1600
series MW reactor [23] with on-line temperature and pressure control in order to in-
vestigate the existence of a nonthermal MW effect. They found, however, that when
the reactions were performed under the action of MW heating using a reflux con-
denser no significant rate increase occurred over conventional reflux at the same
temperature (80 8C). They confirmed that there was an increase in rate and yield
when the reaction was performed in an open beaker, but this could readily be ex-
plained by evaporation of the solvent, causing an increase in temperature and con-
centration solution. It is to be noted that the same explanation applies for the appar-
4.4.4
Reactions in Nonpolar Solvents
There have been a few reports recently of substantial MW rate enhancements of re-
actions of polar reactants in nonpolar solvents [53, 54]. Soufiaoui et al. [53] have
synthesized a series of 1,5-aryldiazepin-2-ones 36 in high yield in only 10 min by the
condensation of o-aryldiamines 34 with b-ketoesters 35 in xylene under MW irradia-
tion in open vessels (Scheme 4.18). The temperature at the end of these reactions
was shown to be 136±139 8C. Surprisingly, they observed that no reaction occurred
when the same reactions were heated conventionally for 10 min at the same tem-
perature. These results could be taken as evidence for a specific MW effect.
Reddy et al. [54] performed a rather similar synthesis of trifluoromethyl aryl-
diazepines 38 and 39 by the reaction of o-aryldiamines 35 with the synthon 37
(Scheme 4.19).
Under MW irradiation (980 W) in xylene yields of 73±93 % were obtained in 10±
20 min, whereas under conventional conditions a complex mixture of products was
obtained, which was unsuitable for preparative work. Thus it would appear that these
reactions not only occur much faster under microwaves but also give cleaner pro-
ducts, indicating greater selectivity. Loupy [44, 45] has speculated that these reac-
tions, among others performed in nonpolar solvents or in the absence of solvent,
might be accelerated under MW irradiation due to the involvement of relatively polar
transition states, leading to lower activation energies. Specific effects would be more
likely when nonpolar solvents are used, since only the reactants would absorb MW
energy, whereas in polar solvents any such effects would be masked due to absorp-
tion of MW by the solvent.
Since there appeared to be strong evidence for a nonthermal effect in this type of re-
action, we repeated the reaction of o-phenylenediamine 34 (Scheme 4.13, R1 = R2 = H)
with ethyl acetoacetate 35 (R = CH3) [19], which was one of the reactions reported by
Soufiaoui [53] to give the diazepine only on MW heating. However, when the same re-
action mixtures were heated for10 min with the same temperature profile, almost
identical yields of the diazepines were obtained by MW and classical heating. Later,
this was also found to be the case in the reaction of 34 with ethyl benzoylacetate 35
(R = Ph).
However, the possibility of the participation of nonthermal effects in MW-assisted
reactions in nonpolar solvents is still an open question. Loupy et al. [55] observed an
increase in yield and purity of the diazepine 36, in the reaction of ethyl acetoacetate
with o-phenylenediamine using monomode MW reactor with focused MW heating,
when compared with conventional heating with the same temperature profile.
Recently Bogdal [48] observed, using kinetic studies, greater MW rate enhance-
ments when the Knoevenagel reaction of salicylaldehyde with ethyl malonate (vide
supra, Scheme 4.15) was performed in toluene than when ethanol was used as the
solvent. The calculated rate constants in toluene solution were more than three
times higher under MW irradiation than under conventional conditions, whereas
the rate constants of the reaction in ethanol were the same, within experimental er-
ror, under both heating methods.
In a joint paper, Loupy and Gedye [45] reported an investigation of the influence
of MW activation on the synthesis of phthalimides by the reaction of phthalic anhy-
dride with several amino compounds (Scheme 4.20) both in the absence of solvent
and in the presence of a nonpolar solvent.
This study was prompted by a report that phthalimidoacetic acid (R = CH2CO2H,
see also Scheme 4.11, vide supra) could be synthesized by the reaction of phthalic an-
hydride with glycine in the absence of solvent, which involves the reaction between
two solids [56]. However, in this study [45], it was established that the synthesis of
phthalimides under solvent-free reactions requires at least one liquid reactant in order
to occur. This was possible when reacting a liquid amine (e. g. R = CH2Ph) or a solid
with a sufficiently low melting point to melt rapidly under MW (e. g. R = (CH2)6OH,
m.p. 56±58). In these cases, the reaction temperature was typically over 135 8C after 2
or 3 min of MW heating resulting in dissolution of phthalic anhydride in the molten
amine, thus producing a homogeneous mixture. Evidence was found in some cases
for a faster reaction under MW irradiation. For example the reaction of phthalic anhy-
dride with 6-amino-1-hexanol (R = (CH2)6OH) in mesitylene (a nonpolar solvent) gave
a 60 % yield of the phthalimide on heating for 10 min at 160 8C using an oil bath. The
yield increased to 97 % when the same reaction was performed using monomode MW
heating for 10 min at the same temperature. The same reaction also occurred faster in
the absence of a solvent when MW heating was used.
If no liquid phase is present, that is, for reactions between two solids, the reaction
will not take place and requires the use of a high-boiling solvent. Thus, contrary to
the previous report [56], the reaction of phthalic anhydride and glycine [R =
CH2CO2H, m.p. 240 8C] did not occur in the absence of a solvent, but when per-
formed in either DMF or xylenes gave a good yield (90 %) of the phthalimide after
only 10 min of MW heating. Finally, in all cases, excellent yields of phthalimides
(about 90 %) were obtained in short reaction times (5±10 min) giving products of
high purity.
Also, as mentioned earlier, Berlan et al. [28] had observed that the use of a nonpo-
lar solvent resulted in a more significant rate enhancement in a Diels±Alder reaction
(vide supra, Scheme 4.12), although later kinetic studies by Raner et al. [33] cast
some doubt on this result.
4.4.5
Reactions in Homogeneous Media Showing no MW Rate Enhancement
4.4.6
Reactions in Homogeneous Media Showing MW Rate Enhancement
at various temperatures, were in the range 20±30 times It was suggested, however,
that these rate enhancements were due to nonuniform heating on a molecular scale
rather than to nonthermal effects.
Wei et al. [65, 66] compared the MW and thermal cure of epoxy resins for stoichio-
metric mixtures of diglycidyl ether of bisphenol A (DGEBA) and a curing agent of
high dipole moment, diaminodiphenylsulfone (DDS), and for DGEBA and a curing
agent of low dipole moment, m-phenylenediamine (mPDA) at the same tempera-
ture. They found significant increases in reaction rates in the MW cured DGEBA/
DDS samples but only slight rate increases for the MW cured DGEBA/mPDA sam-
ples. A possible explanation for the higher rate enhancements of the DGEBA/DDS
system is that when polymers are irradiated with MW, the energy is directly coupled
to the polar functional groups and it takes a finite time for the absorbed energy to
distribute itself evenly with the rest of the system. If this distribution time is signifi-
cantly longer than the reciprocal of the frequency of the MW irradiation, this will re-
sult in a localized higher molecular temperature in the case of the more polar DDS
than in the less polar mPDA [66]. However it was also suggested that nonthermal ef-
fects could be involved since the MW cured samples showed a higher glass transition
temperature, Tg than the thermally cured samples after gelation. It was argued that if
the effects were purely thermal, the Tg should be the same for the MW and thermally
cured samples at the same extent of cure [66].
Scola et al. [67] studied the kinetics of the MW cure of a phenylethyl-terminated
imide polymer model compound and an oligomer using a variable frequency MW
source and found that the activation energy of the MW cures were 68 % and 51% of
the thermal cure respectively. It should be noted that the reactions were performed
in the liquid phase in the absence of solvent.
4.5 Selectivity in MW-assisted Reactions 135
4.4.7
Possible Explanations of MW Acceleration
Because observed rate enhancements are usually small, or zero, nonthermal ef-
fects do not seem to be important in MW heated reactions in homogeneous media,
except possibly in some reactions of polymers and reactions in nonpolar solvents.
Relatively few studies have been conducted on MW-assisted reactions of polar reac-
tants in nonpolar solvents. Also, since there is some disagreement as to whether or
not these reactions are accelerated significantly by MW, in comparison with conven-
tionally heated reactions at the same temperature, more research on the effect of
MW irradiation on the rates of these reactions is required. Nonthermal effects may,
however, explain the more substantial MW rate enhancements in solvent-free reac-
tions on solid supports [44] (see Chapt. 5) and solid state reactions [68, 69].
4.5
Selectivity in MW-assisted Reactions
meric material was formed when the reaction was performed in a modified MW
oven under reflux conditions, resulting in a 32 % yield of the isomeric mixture of the
mixture of adducts.
In some MW-assisted reactions, in which a mixture of products is formed, the ra-
tio of products may be influenced by the heating rate or by the polarity of the sol-
vent.
Stuerga et al. [74] showed that, in the sulfonation of naphthalene, the ratio of the
products, naphthalene-1- and 2-sulfonic acids, depends on the MW power applied.
Higher powers, which cause higher heating rates, increase the proportion of the
thermodynamically more stable product, naphthalene 2-sulfonic acid, in the product
mixture.
Alvarez-Builla et al. [75] have shown that the stereochemical outcome of the MW-
assisted addition of 2-aminothiophenol 50 to the glycidic ester 51 (Scheme 4.27) is
affected by the nature of the solvent and by the MW power output. When the nonpo-
lar solvent toluene was used, the cis isomer of the product 52 predominated, whereas
in the presence of acetic acid the trans isomer was the chief product. Increasing the
power level caused an increase in the proportion of the trans isomer when toluene
was used as the solvent. The reactions were performed in open vessels in a domestic
MW oven.
Also, Bose et al. [76] have shown that the steric course of b-lactam formation can
be influenced by the MW heating rate. For example, in the reaction of the benzoylox-
yacetyl chloride 53 with the Schiff base 54 (Scheme 4.28) the cis adduct 55 is the
main product at low irradiation power whereas high power favors the formation of
the trans adduct 56. Lactams of this type can serve as intermediates for the side chain
of taxol and its analogs.
Bose et al. [77] compared a similar reaction of the acid chloride of tetrachlor-
ophthaloyl glycine 57 with a Schiff base under MW irradiation and conventional
heating (Scheme 4.29). When MW heating was used, the trans isomer 58 was
formed almost exclusively, whereas a mixture of cis and trans isomers 59 and 58 was
obtained under conventional heating. It was noted that the temperature at which the
acid chloride was added to the imine solution was crucial for trans selectivity. The
imine solution was initially heated in the MW oven to about 100 8C and then the acid
chloride added and the mixture heated at a low power level for 3 min.
Pagnotta et al. [78] reported that the mutarotation of a-d-glucose 60 to b-d-glucose
61 in aqueous ethanol (Scheme 4.30) not only occurred more rapidly under MW
heating than conventional heating but also gave a different equilibrium composition,
with a larger concentration of a-d-glucose being obtained using MW irradiation.
Because this change in selectivity is difficult to explain by a classical heating effect,
Langa et al. [9] consider that it is one of the most convincing examples of a possible
specific microwave effect.
Langa et al. [79] recently observed that the regioselectivity of the cycloaddition of
N-methylazomethane to C70 fullerene could be modified by MW irradiation. The ex-
tent of the modification in selectivity, compared with conventional heating, was
shown to be greater in the presence of the polar solvent, o-dichlorobenzene, than in
the nonpolar solvent, toluene. It was suggested that the effect on selectivity in this re-
action could be because of (1) the presence of hot-spots, (2) the greater heating rate
induced by MW or (3) a selective interaction of the electromagnetic field with the
transition state, which gives rise to the different isomers if they have different polari-
ties.
Loupy et al. [70] have studied the effect of solvent on the selectivity of the alkyla-
tion of 1,2,4-triazole 62 by 2,2',4'-trichloroacetophenone 63 (Scheme 4.31) under
MW and conventional heating. The reactions were all performed at 140 8C.
4.5 Selectivity in MW-assisted Reactions 139
In contrast with the results of Langa et al. on the cycloaddition reaction to C70,
MW irradiation had no effect on the regioselectivity of the reactions in polar solvents,
but a substantial effect was observed both in the nonpolar solvent, xylene and under
solvent-free conditions. In polar solvents (pentanol and DMF) the ratio of products
64, 65, and 66 was 95 : 5 : 0 under both MW and conventional heating. In xylene and
in the absence of solvent the ratio of isomers changed from 32 : 28 : 40 (xylene) and
36 : 27 : 27 (no solvent) under conventional heating to 100 : 0 : 0, i. e. totally regioselec-
tive, under MW activation.
The effect of solvent on regioselectivity was attributed to nonthermal effects,
which are favored in nonpolar solvents and under solvent-free conditions, where pro-
ducts formed via more polar transition states would be expected to predominate.
Recently, Jankowski et al. [80] reported the application of MW to the Diels±Alder
synthesis of N-methyloctahydroisoquinoline adducts, which are important inter-
mediates in the synthesis of medicinally important compounds, such as HIV prote-
ase inhibitor isoquinoline carboxylate pharmaceuticals. The reaction of the arecoline
67 or its isomer 70 with Danishefsky's diene 68 in toluene (Scheme 4.32) was studied
under both conventional and MW heating in sealed tubes, i. e. at elevated pressure.
The MW-assisted reaction was performed in a focused MW extraction system (Pro-
labo Soxwave-10) and occurred not only much more rapidly than the conventionally
heated reaction, but also resulted in higher yields and purer products. Under con-
ventional heating (160 8C for 5 days) the yields of the adducts 69 and 71 from 68
with 67 and 70 respectively, were 30 % and 50 %. In contrast, the MW method re-
sulted in yields of 73 % of 69 and 63 % of 71 in only 5 min. The stereochemistry of
the products was not affected by the heating mode. Under conventional reaction con-
ditions a significant level of diene decomposition and polymerization as well as the
formation of dienophile dimers was observed. Under MW activation, however, the re-
action was cleaner and amount of dimerization was reduced. It is interesting to note
that in the case of the MW reaction of 70 with 68 a new ab-unsaturated pyridyl ke-
tone 72 was also produced. It was suggested that this product was formed via the iso-
merization of the double bond of the dienophile 70 from the conjugated position 3
to position 2 (presumably due to the higher temperature of the MW reaction) which
then reacts with the diene to afford compound 72.
Most of the changes in selectivity due to MW heating in homogeneous media dis-
cussed here may be explained by higher reaction temperatures or greater heating
rates in the MW-assisted reactions. However, the observations of Pagnotta et al. [78]
140 4 Organic Synthesis using Microwaves in Homogeneous Media
and Loupy et al. [70] suggest that nonthermal effects may be involved in some cases.
More detailed comparisons between the selectivity of MW and conventionally heated
reactions at the same temperatures and in solvents of different polarity need to be
conducted before definite conclusions can be reached (see also Chapt. 5).
4.6
Comparison of Homogeneous and Heterogeneous Conditions
to styrene were then 11 and 3 times respectively. Again, the higher MW rate en-
hancements under heterogeneous conditions can be accounted for by the increased
temperature of the solid catalyst. It has been reported that CuCl absorbs MW
strongly and heats at a rate of 20 8 s±1 in a 2.45 GHz MW cavity [84].
Recently, Hajek and Radoiu observed that MW not only increase the rate of hetero-
geneous catalytic reactions, but also affect the product selectivity [85]. The results
were explained in terms of MW-induced polarization, involving the absorption of
MW by highly polarized reagent molecules on the active site of the catalyst. On the
other hand there is little, if any, activation of homogeneous catalytic reactions in po-
lar solvents [86], presumably due to the high absorbent power of MW irradiation by
the solvent.
4.7
Advantages and Limitations of MW Heating in Organic Synthesis
The most obvious advantage of MW heating over conventional heating is that it al-
lows many reactions to be performed more rapidly. In addition there have been
many reports of increased yields and fewer side-reactions when MW heating is used.
Rate increases of two or three orders of magnitude are achievable by performing re-
actions in sealed containers and hence at increased pressures. Earlier applications of
this technique were limited to small-scale reactions at pressures of up to about
700 Kpa to avoid the possibility of vessel rupture or explosions [1±3, 20]. Computer-
ized MW systems with temperature and pressure control are now available allowing
reactions to be performed safely and rapidly [8, 21, 23]. Hence there can be consider-
able time savings in a synthesis, which may have one or more steps requiring long
reflux periods of conventional heating. Unfortunately, such systems are expensive at
present, and are not, for example, usually available for routine use in university re-
search laboratories. It can be expected, however, that this situation will change in the
next few years as organic chemists become more aware of their potential to drasti-
cally reduce the time required for slow reactions as well as for obtaining products of
greater purity.
MW-assisted reactions are often performed in comparatively inexpensive domestic
MW ovens at ambient pressure. The ovens are sometimes modified to allow the reac-
tions to be performed under reflux [7, 19], thus preventing rapid evaporation of the
solvent which could otherwise present a fire hazard. Alternatively, the MORE tech-
nique [17] can be used, where the reactants are dissolved in a high-boiling solvent
such as DMF or may be performed in the absence of a solvent. Although actual rate
enhancements are small and sometimes nonexistent, reaction mixtures can be
heated more rapidly than under classical heating leading to some time-saving. MW
heating avoids the use of dangerously hot oil baths, which also require a considerably
longer time to reach the required temperature [58].
The variability of domestic MW ovens, nonhomogeneity of the MW field, and diffi-
culties in temperature measurement often make it difficult to obtain reproducible
rates and yields [10], and this may partly explain why MW heating has not become
References 143
more widely used by synthetic organic chemists. These problems also apply to sol-
vent-free reactions on solid supports, when performed in domestic microwave ovens,
but they can be avoided by using focused microwaves in monomode systems [43].
Again, the expense of these systems has so far limited their use to specialized re-
search laboratories. However, with the impressive advances in technology for the de-
sign of MW systems for organic synthesis in the last few years, there is little doubt
that these systems will soon become much more widely used by organic chemists.
The MW technique offers considerable advantages over conventional heating for sol-
vent-free reactions on solid supports because of rapid heating, substantial rate en-
hancements and cleaner reactions.
For these reasons, and because avoidance of solvents makes the technique more
environmentally friendly, the majority of recent papers on MW-assisted organic
synthesis have reported the use of solvent-free reactions.
MW heating will, of course, only be effective if the reaction mixture absorbs MW
irradiation and so is limited to reactions of polar molecules, either in polar solvents
or in the absence of a solvent. In some cases, reactions of polar compounds in non-
polar or slightly polar solvents can be heated with MW [19, 53], but the solutions
must be must be sufficiently concentrated in order to heat effectively. Relatively high
MW powers are also required. For this reason, some reactions requiring several
hours of classical reflux in solvents of low polarity can be performed more rapidly in
a MW oven by using a more polar solvent [17].
Although most applications have been in small-scale syntheses, recent technology
has led to the design of continuous flow and batch systems [8, 23] making it feasible
to synthesize products on the kilogram scale. Many noteworthy applications of MW-
assisted synthesis have been reported including the synthesis of radiopharmaceuti-
cals containing atoms with short half-lives [27], the synthesis of organometallic com-
pounds [7] and the synthesis and cross-linking of polymers [64, 87, 88]. Recently, the
great potential of rapid MW-assisted organic synthesis for combinatorial chemistry
and drug discovery is receiving considerable interest and has been recognized in two
reviews [89, 90].
Finally, there are a number of reports of modifications in selectivity and this aspect
of MW synthesis [9, 44, 70] promises to receive much attention in the future.
References
5
Microwave and Phase-transfer Catalysis
Andr Loupy, Alain Petit, and Dariusz Bogdal
5.1
Phase-transfer Catalysis
In the last three decades there have been many reports of organic reactions per-
formed under phase-transfer catalytic conditions (PTC) [1±4]. PTC has found numer-
ous applications in essentially all fields of organic syntheses, industrial chemistry,
biotechnology, and material sciences. It can be encountered in the manufacture of
advanced pharmaceuticals, fragrances, crop-protection chemicals, highly advanced
engineering plastics, materials for semiconductors, and electro-optical and data sto-
rage devices [5]. It is worth remarking that in the 1990s sales of products made by
use of phase-transfer catalysis exceeded 10 USD billion year±1.
The concept of PTC was developed mainly because of the work of independent re-
search groups led by Makosza [1] and, later, Starks [2], Brandstrom [3], and Dehm-
low [4]. They introduced techniques in which the reactants were situated in two se-
parate phases, e. g. liquid±liquid or solid±liquid. Because the phases were mutually
insoluble, ionic reagents (i. e. salts, bases, or acids) were dissolved in the aqueous
phase while the substrate remained in the organic phase (liquid±liquid PTC). In so-
lid±liquid PTC, on the other hand, ionic reagents can be used in their solid state as
a suspension in the organic medium. Transport of the anions from the aqueous or
solid phase to the organic phase, where the reaction occurred, was ensured by use
of catalytic amounts of lipophilic agents, usually quaternary onium salts or cation-
complexing agents (e. g. crown ethers). Because the reactions proceeded very slowly,
or not at all, in the absence of a catalyst, phase-transfer catalysts have been found to
be of utmost importance in the extraction of reaction species between phases so
that the reaction can proceed, thus increasing yields and rates of reactions substan-
tially.
In general, under PTC conditions, three types of catalytic procedure can be con-
sidered (Scheme 5.1).
Scheme 5.1
organic phase as loose lipophilic ion pairs, because of the intervention of the
phase-transfer catalyst, which continuously transfers the anionic species from the
aqueous phase to the organic phase, in which the reaction occurs (Scheme 5.1;
path a);
2. Liquid±liquid PTC conditions in which weak organic acids (e. g. carboanions) re-
act in the presence of concentrated aqueous sodium or potassium hydroxide
which is in contact with the organic phase containing an anion precursor and or-
ganic reactants; the anions are created on the phase boundary and continuously
introduced, with the cations of the catalyst, into the organic phase, in which
further reactions occur (Scheme 5.1; path b).
3. Solid±liquid PTC conditions in which the nucleophilic salts (organic or mineral)
are transferred from the solid state (as they are insoluble) to the organic phase by
means of a phase-transfer agent. Most often the organic nucleophilic species can
be formed by reaction of their conjugated acids with solid bases (sodium or potas-
sium hydroxides, or potassium carbonate) (Scheme 5.1; path b). Another pro-
posed mechanism suggests that interfacial reactions occur as a result of absorp-
tion of the liquid phase on the surface of the solid.
The organic phase can be a nonpolar organic solvent (e. g., benzene, toluene, hex-
ane, dichloromethane, chloroform, etc.) or a neat liquid substrate, usually the elec-
trophilic reagent, which acts both as a reactive substrate and the liquid phase.
In chemical syntheses under the action of microwave irradiation the most success-
ful applications are necessarily found to be the use of solvent-free systems [6]. In
these systems, microwaves interact directly with the reagents and can, therefore,
drive chemical reactions more efficiently. The possible acceleration of such reactions
might be optimum, because they are not moderated or impeded by solvents. Reac-
tions on solid mineral supports and, in turn, the interaction of microwaves with the
reagents on the solid phase boundary, which can substantially increase the rate of
the reactions, are of particular interest [7].
PTC reactions are perfectly tailored for microwave activation, and combination of
solid±liquid PTC and microwave irradiation gives the best results in this area [8]:
1. after ion-pair exchange with the catalyst, the nucleophilic ion pair [Q+Nu±] is a
highly polar species especially prone to interaction with microwaves;
5.1 Phase-transfer Catalysis 149
was very modest (Fig. 5.1, curve c). In the presence of small amounts of xylene
(Fig. 5.1, curve d), a nonpolar solvent (i. e. inert to MW irradiation), a large tempera-
ture increase provides evidence of the formation of tetrabutylammonium benzoate
in the liquid phase, from which the positive thermal effect results (Fig. 5.1, curve d).
It must be stressed that a liquid component can be substituted with an efficient
absorber of microwave irradiation together with a low-melting component. The use
of most typical PTC solvents (nonpolar aromatic or aliphatic hydrocarbons, or highly
chlorinated hydrocarbons) is most interesting for microwave activation, because
such solvents are transparent or absorb microwaves only weakly. They can, therefore,
enable specific absorption of microwave irradiation by the reagents, and the results
or product distributions might be different under microwave and conventional con-
ditions [7].
Hence, by coupling microwave technology and solid±liquid phase-transfer conditions, we
create a clean, selective, and efficient methodology for performing organic reactions,
with substantial improvements in terms of conditions and simplicity in operating
procedures; this is essentially useful for poorly reactive systems involving, for in-
stance, hindered electrophiles or long-chain halides.
Numerous reactions in organic synthesis can be achieved under solid±liquid PTC
and with microwave irradiation in the absence of solvent, generally under normal
pressure in open vessels. Increased amounts of reactants can be used to ensure bet-
ter compatibility between the in-depth penetrability of materials and the radiation
wavelength.
Because microwave activation is rather a new technique, the number of examples
might seem limited at the present time; it is, however, increasing rapidly.
5.2
Synthetic Applications of Phase-transfer Processes
5.2.1
O-Alkylations
Alkyl acetates
Potassium acetate can be readily alkylated in a domestic microwave oven by use of
equivalent amounts of salt and alkylating agent in the presence of Aliquat 336
(10 mol%). Some important results, exemplified by Eq. (1), are given in Tab. 5.1
[10, 11].
+ -
CH 3COO K
- +
+ R X CH 3 COOR + KX Eq. (1)
5.2 Synthetic Applications of Phase-transfer Processes 151
Tab. 5.1 Alkylation of CH3COO±K+ under MW + PTC conditions (domestic oven, 600 W).
n-C8H17Br 1 187 98
n-C8H17Cl 1 162 98
n-C8H17I 2 165 92
n-C16H33Br 1 169 98
Yields are always almost quantitative within 1±2 min, irrespective of chain length
and the nature of the halide leaving groups.
This procedure was scaled up from 50 mmol to the 2 mol scale (i. e. from 15.6 to
622.4 g total starting materials) in a larger batch reactor (Synthewave 1000) [12].
Yields were equivalent to those obtained under the original conditions (5 min,
160 8C) in the laboratory-scale experiment (Synthewave 402) (Tab. 5.2).
Tab. 5.2 Synthesis of n-octyl acetate under MW + PTC conditions (Synthewave, 5 min, 160 8C).
Xu et al. have obtained similar results with n-butyl bromide using TBAB
(10 mol%) and alumina (4 : 1 w/w) as the catalyst [13]. Benzyl acetate was also conve-
niently prepared from sodium acetate and benzyl halide by use of microwave irradia-
tion and PTC in synergy [14].
Long-chain esters
As a generalization of the above method, stearyl stearate was synthesized in 1 min in
quantitative yield (Eq. 2) [11].
- +
Aliquat 10%
C17 H35 COO K + n -C18 H37 Br C17 H35 COO n -C18 H37
MW (600 W) 1 min. Eq:
2
99%
It has been shown that reaction of carboxylic acids with benzyl halides, which
does not occur when heated conventionally, could be performed efficiently under the
action of MW irradiation in the presence of a quaternary ammonium salt as a cata-
lyst (Eq. 3) [15]. Typical results are given in Tab. 5.3.
+ -
Q X
Ar CH 2X + n C5H11 COOH Ar CH 2 O CO C 5H11 + HBr Eq:
3
152 5 Microwave and Phase-transfer Catalysis
Tab. 5.3 Reaction of benzyl halides with hexanoic acid under MW + PTC conditions (10 min,
560 W, 10 % PhCH2N+Me3Cl± ).
Benzyl bromide 72
Benzyl iodide 90
a-Bromo-p-xylene 76
a-Iodo-p-xylene 92
a-Bromomesitylene 81
Aromatic esters
It is possible to alkylate benzoic acids directly, without the need to prepare reactive
potassium salts in a separate step, because they can be generated in situ by reacting
the acid with a base (potassium carbonate or hydroxide) in the presence of a phase-
transfer catalyst.
As an illustration of this principle, a volatile polar molecule is a byproduct, elimi-
nated as a result of the microwave heating (Eq. 4), and the equilibrium is shifted to
completion. The second effect of irradiation is activation of the alkylation step itself
(Eq. 5). All the reagents can be used in the theoretical stoichiometry. Some indicative
results are given in Tab. 5.4 [9].
Aliquat 10%
ArCOOH + KOH ArCOO-K+ + H2O Eq:
4
Aliquat 10%
ArCOO K
- +
+ R X ArCOOR + K X
+ -
Eq:
5
Tab. 5.4 Alkylation of potassium 4-Z-benzoate under MW + PTC conditions (domestic oven,
600 W, 10 % Aliquat).
H 2.5 99 99
NMe2 3 97 100
OMe 2 82 98
CN 3 80 95
NO2 2 81 95
A striking example in this series is the alkylation of terephthalic acid (Eq. 6). The
specific effect of the microwaves appears clearly in this example, because, other fac-
tors being equal, the yields are unambiguously much higher.
5.2 Synthetic Applications of Phase-transfer Processes 153
Aliphatic ethers
Yuan et al. studied two types of condition for this reaction ± use of either the alcohols
or the corresponding halides as starting materials [16, 17]. In the presence of qua-
ternary ammonium salts the reaction shown in Eq. (7) is complete within a few min-
utes. Typical results are given in Tab. 5.5.
+ -
PhCH 2N Me3 Cl
ROH + R'X R-O-R' + HX Eq:
7
Tab. 5.5 Synthesis of ethers under MW + PTC conditions (domestic oven, 560 W).
Et PhCH2Cl 5 85
n-Bu PhCH2Cl 10 78
n-Oct PhCH2Cl 10 88
n-Oct n-BuBr 10 78
PhCH2 n-BuBr 10 92
More recently, this method has been extensively applied to a wide range of Wil-
liamson syntheses in dry media with K2CO3 and KOH as bases, TBAB as phase-
transfer agent, and a variety of aliphatic alcohols (e. g. n-octanol and n-decanol, yields
75±92 %) [18].
KOH
HOH 2C CH 2 OH + R X ROH2C CH 2OR Eq. (8)
O Aliquat O
Microwaves Monomode Reactor (30 W) 5 min. 180°C 93%
Conventional heating (Oil Bath) 5 min. 180°C 41%
154 5 Microwave and Phase-transfer Catalysis
The diethers were synthesized in high yield within short reaction times. When
compared with classical heating, under otherwise comparable conditions, reactions
times were improved by microwave activation.
Diethers from dianhydrohexitols
A series of new ethers has been obtained by alkylation of dianhydrohexitols (i. e. iso-
sorbide, isomannide, isoidide) under the action of microwave irradiation and PTC
conditions. Yields exceeded 90 %, a dramatically improvement compared with those
from conventional heating, despite similar temperature profiles. The best yields, for
example from isosorbide, were obtained in the presence of a small amount of xylene
and TBAB as catalyst at 140 8C (Eq. 9 and Tab. 5.6) [20].
Eq. (9)
Tab. 5.6 Reaction of isosorbide with several alkylating agents in the presence of KOH and TBAB
(relative amounts 1 : 3 : 3 : 0.1) under the action of microwave irradiation in a xylene solution.
n-C8H17Br 5 140 96 10
C6H5CH2Cl 5 125 98 13
3-Cl-C6H4CH2Cl 5 125 95 15
4-Cl-C6H4CH2Cl 5 125 96 14
3-F-C6H4CH2Cl 5 125 95 15
CH3CH2OCH2CH2Br 30 100 78 45
Eq. (10)
5.2 Synthetic Applications of Phase-transfer Processes 155
Tab. 5.7 Polyethers from isosorbide and 1,8-dibromo- or dimesyloctane; yield and distribution
data for the microwave procedure (Mn and Mw are, respectively, the number average and weight
average molecular weights, the ratio Mw/Mn being the polydispersity index).
Phenolic ethers
Gram quantities of aryl-2-(N,N-diethylamino)ethyl ethers, compounds of biological
interest, have been prepared in a household microwave oven, with potassium hydro-
xide and glyme as the transfer agent, according to Eq. (11) [22].
OH O
KOH, glyme NEt2 Eq. (11)
ClCH 2CH 2 NEt2
R R
MW (600 W) 2 min. 92-99%
Under the action of microwave irradiation several phenol reacts remarkably fast in
dry media with primary alkyl halides to give aromatic ethers (Eqs. 12 and 13) [23, 24].
Eq. (13)
Catechol was reacted with b-methallyl chloride under the action of MW and PTC
conditions; yields of 2-methallyloxyphenols varied from 59 to 68 % under liquid±li-
quid conditions (Tab. 5.8), whereas no reaction was observed in a solid±liquid PTC
procedure (Eq. 14) [25].
156 5 Microwave and Phase-transfer Catalysis
CH3
OH CH 3 OH O CH 2 C
base, PTC CH2
+ CH 2 C + CH3
CH 3
OH CH2 Cl H2 O, MW 1-2 min O CH 2 C O CH 2 C
CH 2 CH 2
mono di
Eq. (14)
Tab. 5.8 Reaction of catechol with b-methallyl chloride (8 mmol) in the presence of base (2 mmol)
and phase-transfer catalyst (0.2 mmol) under microwave irradiation.
HO O
ICH2I / TBAB / CaCO3 Eq. (15)
HO CHO Support / MW O CHO
3 min 157°C 95%
O KOH O
C OH + C16H33Br C O C16H33 Eq. (16)
H Aliquat H
15 min 140°C
MW 96% ∆ 22%
O KOH + K 2CO3 O
C OH + RBr C OR Eq. (17)
H TBAB H
OH MW, 10 min 130°C OR
Tab. 5.9 Alkylation of 3,4-dihydroxybenzaldehyde with long chain alkyl bromides under the action
of MW + PTC conditions.
Eq. (18)
Desyl ethers are key intermediates in the synthesis of biologically active furanopy-
rones. MW-assisted synthesis of these has been performed in an open vessel under
PTC solvent-free conditions (Eq. 19) by alkylation of 7-hydroxy-4-methyl coumarin
with desyl chloride [30].
H3C O O O Ph H3C O O O
K 2CO3/NBu 4HSO4
+ Cl CH COPh
MW 2 min
OH O CH COPh
98% Ph
Eq. (19)
Eq. (21)
CHO CHO
K 2CO3, TBAB
+ Cl CO2R CO2R
OH 8-10 min O O
X X CO2R X
Y Y Y
(1) (2)
Eq. (22)
It was found that the first step was rate determining. When, moreover, the reac-
tion was run with the same reaction-temperature profiles under both conventional
(oil) and microwave (monomode cavity) conditions, different distributions of the in-
termediate (1) and final (2) products were obtained (Tab. 5.10). Indeed, the product
distribution was strongly affected by microwaves when the reaction was run at 85 8C
rather than 110 8C, and addition of a small amount of a polar or nonpolar solvent
also affected the product distribution. In this work two solvents capable of extensive
coupling (i. e. ethanol) and not coupling (i. e. cyclohexane) with microwaves were
used. Addition of ethanol strongly shifted the product distribution towards the final
product (2), whereas addition of cyclohexane resulted in much lower yield of 2 [34].
Tab. 5.10 Distribution of the intermediate (1) and final product (2) in the synthesis of benzo[b]fur-
ans under both conventional and microwave conditions.
110 10/MW ± 12 66
110 20/MW ± ± 100
110 20/D ± 4 96
85 20/MW ± 15 85
85 20/D ± 90 10
85 30/MW Cyclohexane/1.5 62 38
85 30/MW EtOH(100 %)/1.5 0 100
85 30/MW EtOH(96%)/1.5 0 100
85 30/D EtOH(100 %)/1.5 47 53
Phenolic polyethers
Polycondensations of 3,3-bis(chloromethyl)oxetane and a variety of bisphenols
were studied using the microwave-PTC technique (Eq. 23) [35]. The results ob-
tained showed the advantages of microwaves in terms of the molecular weights
for crystalline polymer, as reflected in higher values of the transition temperature
(Tg) and melting point (Tm) but also in reduction of reaction times for all types of
structure.
5.2 Synthetic Applications of Phase-transfer Processes 159
Eq. (23)
5.2.2
N-Alkylations
O O
Silica gel
N-Na+ + R Br N R Eq. (24)
S TEBA
S
O O O O
Tab. 5.11 N-Alkylation of saccharin with bromides under MW + PTC conditions (750 W).
PhCH2 None 4 8
SiO2 4 75
SiO2 + TEBA 4 92
n-C16H33 SiO2 6 82
SiO2 + TEBA 6 97
Y NaOH, TBAB Y
+ R X Eq. (25)
N O Silica gel N O
H R
Y = O, S MW 8-10 min. 72-90%
The Michael reaction of the same compounds with acrylonitrile and methyl acry-
late was promoted under the same conditions within 9±10 min [38].
Et Et Et Et
+ -
O O K2CO3, Me 3N C16 H33 , Br O O
+ R X Eq. (26)
N N MW (400 W) 6-18 min. N N
H H R R
O O
Ex : PhCH2Br, 10 min., 99% 86-99%
O NaOH, TBAB O
H3 C + R X H3C Eq. (27)
NHPh MW N Ph
R
Ex : R-X = n -BuI 80 sec. 88%
R2 K CO , KOH, TBAB R2
1 2 3 1
R X + R N R N
O MW, 55-120 sec O
H R
R Eq. (28)
NH K2 CO3 , KOH, TBAB N
R X + (CH2 )n (CH2)n
MW, 80-150 sec
O O
Microwaves 55-120 sec 45-92%
with 50 % excess alkyl halide and catalytic TBAB, which were later absorbed on po-
tassium carbonate. Irradiation of the reaction mixtures in a domestic oven led to the
desired phthalimide derivatives (Eq. 29) in a short time (4±10 min) [42].
O O
K2 CO3 , TBAB
N H + R X N R Eq. (29)
MW (450 W) 4-10 min.
O O
49-95%
Ex : R-X = PhCH 2 Cl 4 min. 93%
n -BuBr 4 min. 73%
n -C12 H 25 I 10 min. 95%
The same reaction (R±X = n-OctBr) was studied using TBAB and several basic
supports. Na2SO4 and CaCO3 (yields 93 and 95 %, respectively, within 3 min of MW
irradiation) were shown to be more efficient than K2CO3 (71%) [43].
Et
NaOH / CTAB
PhNHEt + PhCH 2Cl Ph N CH 2Ph Eq. (30)
H2O
MW 25 min 90%
∆ 16h 90%
Ph Ph
K 2 CO 3 , TBAB
N H + R X N R
MW (780 W) 10 min., 80°C
Eq. (31)
Ex : R-X = PhCH2 Br 70%
CH 2 =CH2 -CH2 Br 39%
n -C 8 H 17 Br 31%
The synergy between the dry media and microwave irradiation was convincingly
demonstrated in this work. For instance, with the allyl compound, the yield is only
16 % after 24 h in toluene under reflux, and no reaction occurs after 10 min at
100 8C (classical heating), thus revealing an important specific microwave effect.
162 5 Microwave and Phase-transfer Catalysis
N K 2 CO 3 / KOH, TBAB N
+ R X Eq. (32)
N MW (300 W) 30 sec.-1 min. N
H R
Ex : PhCH 2 Cl, 40 sec., 89% 73-89%
K2CO3, TBAB
+ R X Eq. (33)
MW (450 W) 4-10 min. N
N
H R
79-95%
Ex : R-X = PhCH2Cl 4 min. 70%
n -BuBr 4 min. 73%
n-C10 H21 Cl 10 min. 95%
O O
HN RX / DMF HN
Eq. (34)
Na 2 CO3 S N NH2
S N NH2
nBu4NI
H R
p-NO2C6H4CH2Cl 6 70± 75 91
p-BrC6H4COCH2Br 6 70± 75 95
Br-(CH2)5-Br 10 80± 85 87
n-C16H33Br 9 105±110 91
5.2 Synthetic Applications of Phase-transfer Processes 163
5.2.3
C-Alkylations of Active Methylenes [48±57]
n-Bu4N+X- R COOEt
R COOEt + R' X + Base Eq. (35)
R'
PhSO2 PhCH2Cl 3 76
n-BuBr 3 83 49
n-OctBr 3 79
PhCH=N PhCH2Cl 1 63 50
n-BuBr 2 55
PhS PhCH2Cl 4.5 83 51, 52
n-BuBr 4.5 59
CH3CO PhCH2Cl 3 81 53
n-BuBr 4.5 61
COOEt PhCH2Cl 2 72
n-BuBr 2 86 54
AllylBr 2 75
p-NO2-C6H4 Ph-(CH2)3-I 3.5 55 55
Ph-(CH2)8-I 7 79
p-NO2-C6H4 Ph-S-(CH2)6-Br 7 50 55
Ph-S-(CH2)8-Br 7 59
Rapid monoalkylations are achieved in good yield compared with classical methods.
Of particular interest is the synthesis of a-amino acids by alkylation of aldimines with
microwave activation. Subsequent acidic hydrolysis of the alkylated imine provides
leucine, serine, or phenylalanine in preparatively useful yields within 1±5 min [50].
Alkylation of phenylacetonitrile was performed by solid±liquid PTC in 1±3 min
under microwave irradiation (Eq. 36 and Tab. 5.14). The nitriles obtained can subse-
quently be quickly hydrolyzed in a microwave oven to yield the corresponding
amides or acids [56].
R
NaOH, TEBACl
PhCH 2 CN + R-X Ph CH CN + Ph C CN Eq. (36)
H2 O R R
164 5 Microwave and Phase-transfer Catalysis
Tab. 5.14 Alkylation of phenylacetonitrile under MW + PTC conditions (NaOH 6 equiv., TEBA
15 %).
O O X O X
X
K 2CO3 / Aliquat Y -H2 O Y
O + CH 2 OH
Y Eq. (37)
Tab. 5.15 Epoxyisophorone ring opening by reaction with several active methylene compounds un-
der MW + PTC conditions.
5.2.4
Alkylations with Dihalogenoalkanes
5.2.4.1 O-Alkylations
Ethers from alkylation of furfuryl alcohol with dihalides were obtained under solvent
PTC + MW conditions with quasi-quantitative yields (Eq. 38) [19].
Eq. (38)
5.2 Synthetic Applications of Phase-transfer Processes 165
OH OH O R O
O PhCH2 Br, KOH O X-R-X O O
Tab. 5.16 Reaction of monobenzyloxy isosorbide with X±R±X under microwave irradiation.
5.2.4.2 S-Alkylations
6-Mercaptobenzimidazo[1,2-c]quinazoline reacts with dibromo derivatives under
PTC conditions (K2CO3/TBAB). Microwave irradiation enabled a striking reduction
in reaction times (12 h compared with 15 min) with similar yields (86 and 81%)
when compared with conventional heating (Eq. 40) [61].
N
N
N S
K2CO3, TBAB, DMF
+ Br-(CH2)n-Br (CH2)n Eq. (40)
N MW, 15 min, 60°C S
n = 1, 2
N N
HS N
N
n=1 86% n=2 81%
166 5 Microwave and Phase-transfer Catalysis
5.2.5
Nucleophilic Additions to Carbonyl Compounds
A second example of aldolization (Eq. 42) is the ªdryº reaction of ferrocene carbal-
dehyde with carbonyl compounds in the presence of potassium hydroxide, and Ali-
quat as catalyst [63]. Reactions which are too slow at room temperature are efficiently
accelerated by use of microwaves, giving good yields within a few minutes.
O
CHO
R2
O R1
KOH, Aliquat
Fe + R 1 CH 2 R2 Fe Eq. (42)
-H2O
Microwave activation and solvent-free PTC have been shown to be of prime effi-
ciency for the synthesis of new benzylidene cineole derivatives (UV sunscreens) by
the Knoevenagel reaction. When performed classically by use of KOH in ethanol at
room temperature for 12 h (Eqs. 43 and 44) the yield was 30 %. This was improved
to 90±94 % within 2±6 min under PTC + MW conditions (Tabs 5.17 and 5.18) [27, 28].
O O
O
KOH (1,5 eq) CH R
+ H C R
Aliquat (5%)
no-solvent
O O Eq. (43)
CH3 2 180 90 20
N(CH3)2 2 180 94 40
OC16H33 6 200 94 12
5.2 Synthetic Applications of Phase-transfer Processes 167
O O
O
K 2 CO3 + KOH CH OR
+ H C OR
TBAB OR
OR
O O Eq. (44)
Tab. 5.18 Benzylidene derivatives of 2-cineolylol from 3,4-dialkoxybenzaldehydes [28].
C12H25 60 140 80 13
C14H29 60 140 78
C16H33 60 140 81
C18H37 60 140 87
1. KOH, Aliquat
R COOR' R COOH Eq. (45)
2. HCl
From the few examples summarized in Tab. 5.19, three important conclusions can
be drawn:
. Rapid and easy reactions occur, even with the most hindered mesitoic esters,
which are otherwise practically nonsaponifiable under classical conditions [65].
. The advantage of using a monomode reactor rather than a domestic oven is clearly
apparent.
. More interesting fundamentally is the very strong specific nonthermal effect of
microwaves, as evidenced by comparison with classical heating. This effect grows
as ester reactivity falls.
168 5 Microwave and Phase-transfer Catalysis
In a more detailed study, it was shown that MW effects are strongly dependent on
the temperature and the nature of the cation associated with hydroxide anion [66]
(for example Eq. (46) and Tab. 5.20).
Aliquat
CO2CH3 + M OH
+ -
CO2H + CH3OH Eq. (46)
then HCl
K+ 2 200 5 94 90 98 98
K+ 1 70 60 77 42 67 65
Na+ 2 200 5 92 77 90 73
Base
ArCOOMe + n -C8H17 OH ArCOOn -C8 H17 + MeOH Eq. (47)
Aliquat
C6H5 KOH 3 40
KOt-Bu 3 42
K2CO3 2.5 90
2,4,6-Me3C6H2 KOH 10 64
KOt-Bu 10 68
K2CO3 10 89
This study was next extended to the synthesis of benzoyl and dodecanoyl deriva-
tives from protected carbohydrates [67]. Microwave-assisted PTC transesterifications
with methyl benzoate or dodecanoate were studied for several carbohydrates. Small
amounts of dimethylformamide (DMF) were shown to be necessary to provide good
yields (76±96 %) within 15 min. Rate enhancements when compared to conventional
heating (D) and specific microwave activation were especially noticeable when less
reactive fatty compounds were involved (Eq. 48).
5.2 Synthetic Applications of Phase-transfer Processes 169
O O
O OH K2 CO3, TBAB O OR
O O
+ RCO2 CH3 Eq. (48)
O 15 min., 160°C O
O O
R = Ph MW 44%
DMF MW 96% ∆ 21%
R = CH 3(CH 2)10 DMF MW 88% ∆ 0% (id. 12h)
5.2.6
Deprotonations
OMe OMe
HO HO
t BuOK, Aliquat
Eq. (49)
MW (45 W) 18 min.
94%
NaOH ..
CHCl 3 CCl 2 Eq. (50)
CTAB
Cl Cl
5.2.6.3 b-Elimination
Bromoacetals in basic media can be converted to cyclic ketene acetals (Eq. 51). These
b-eliminations, previously performed under solid±liquid PTC without solvent and
with sonication [70], were further improved by microwave irradiation (Tab. 5.22) [71].
170 5 Microwave and Phase-transfer Catalysis
Br O Base O
Eq. (51)
O TBAB O
Tab 5.22 b-Elimination from bromoacetals in a monomode reactor (75 W); comparison with sonochemical condi-
tions and classical heating.
With potassium t-butoxide and TBAB, higher yields are obtained more rapidly
than under sonication conditions or with conventional heating.
5.2.7
Miscellaneous Reactions
Cl OEt
Catalyst 10%
Z + EtOH + NaOH Z Eq. (52)
-NaCl, -H 2 O
Tab. 5.23 Aromatic nucleophilic substitution under MW + PTC conditions (420±700 W).
4-NO2 None 2 14
PhCH2N+Me3Cl± 2 51 72
PEG 400 2 99
2-OH None 2 63
PhCH2N+Me3Cl± 2 70 73
PEG 400 2 82
5.2 Synthetic Applications of Phase-transfer Processes 171
Tab. 5.24 Solvent-free PTC SNAr reactions under MW + PTC conditions (potassium salts).
F NO 2 + PhO - ± 30 150 98 73
Cl NO 2 + CH 3 O- 18-crown-6 20 170 90 37
Cl
+ CH3 O- 18-crown-6 60 100 71 27
F
+ CH3 O- 18-crown-6 60 80 94 57
O O
COOEt + - R
LiBr, Bu 4N Br
R Eq. (53)
H2O
H 8 30 138 96
Et 15 30 160 94
n-Bu 20 45 167 89
n-Hex 20 90 186 87
172 5 Microwave and Phase-transfer Catalysis
Eq. (54)
The cycloaddition can be performed almost quantitatively within 6 min under the
action of microwaves, with KF as base, and Aliquat. For the sake of illustration, re-
sults for substituted chalcones are listed in Tab. 5.26. When the same reactions are
performed all other factors being equal (time, temperature), no reaction occurs un-
der classical thermal conditions. This behavior once more confirms a specific radia-
tion effect.
Tab. 5.26 1,3-Dipolar cycloaddition of DNPI to chalcones in a monomode reactor (30 W).
H 6 170 90
Br 5 170 93
Cl 6 174 95
Me 6 168 87
OMe 6 175 89
H
Ph Ph
Me OSiMe 3 KF - 18-crown-6 Ph N
N + Eq. (55)
H Ph H OMe MW (300 W) 7 min. H
Me O
93% (anti/syn = 65/35)
OEt OEt OH
OMe KOtBu + 18-crown-6 OH OMe
+ Eq. (56)
(Ethylene glycol)
R R R
R = CH=CH-CH3 , H 1 2
± MW 20 120 7 ± 90
± D 20 120 48 ± 50
EG MW 75 180 ± 72 23
EG D 75 180 98 ± ±
This procedure was extended to the hundred-gram scale with success because, un-
der rather similar conditions, yields remained excellent (82 %) requiring only a slight
modification in temperature (140 instead of 120 8C) (Tab. 5.28) [12].
Tab. 5.28 Deethylation of 2-ethoxyanisole within 20 min under MW + PTC conditions (KOtBu 2
equiv.; TDA-1 10 %).
Reactor Temperature (pC) Amounts of materials [g (mmol)] Total amount (g) Yield (%)
Ethoxyanisole KOtBu TDA-1
Tab. 5.29 Dibenzyl diselenides under MW + PTC conditions (15 min, 750 W; 15 equiv. NaOH;
5 % PEG 400).
R X Yield (%)
H Cl 75
H Br 95
4-Br Br 85
4-CH3 Br 91
4-NO2 Br 72
The same authors have more recently described the synthesis of dibenzoyl disele-
nides by reaction of selenium with sodium hydroxide under PTC and MW irradia-
tion conditions to afford sodium diselenides which react further with benzoyl chlor-
ide at 0 8C [81].
O
aq NaOH - PEG 400
R C N R C NH 2 Eq. (58)
MW
Ph2CH 90 94
Ph 45 83
PhCH2 90 71
C17H35 60 78
2-Pyridyl 40 52
5.2 Synthetic Applications of Phase-transfer Processes 175
Eq. (59)
Tab. 5.31 Synthesis of diaryl-a-tetralones from isophorone and chalcone under MW + PTC condi-
tions.
C6H5 5 110 86 51
p-CH3O-C6H4 5 108 86 51
p-CH3-C6H4 6 120 89 52
o-Cl-C6H4 5 110 88 54
p-Cl-C6H4 6 110 86 56
COOEt
X X COOEt
K2CO3, I 2 H
O O Eq. (60)
Y Z
Y O Aliquat 336
toluene O
Z
Microwave (boiling toluene) 10-16 min 65-67%
Conventional heating (boiling toluene) 25-30 min 65-94%
Room temperature (toluene) 6-13 h 65-94%
176 5 Microwave and Phase-transfer Catalysis
O
Na2WO4, TBAHS
R OH + H2O2
MW, 20 min R OH
OH O
Na2 WO4 , TBAHS
+ H2O2
R1 R2 MW, 10 min R1 R2
Eq. (62)
OH O
Na2WO4, TBAHS
+ H2O2
MW, 10 min
OH OH
Microwaves 10-20 min 60-97%
Conventional heating at 90°C 4h 83-96%
5.3 Conclusion 177
Hydrogen peroxide has also been used with microwave irradiation for the epoxida-
tion of simple or cyclic alkenes. The reactions were accomplished under liquid±li-
quid PTC conditions in ethylene chloride solution in the presence of Na2WO4 and
Aliquat 336 as catalysts. The best results were obtained at 70 8C when the concentra-
tion of hydrogen peroxide was set to 8 % and the pH of aqueous phase was kept be-
low 2 (Eq. 63) [90].
n + H2 O2 n Eq. (63)
MW, 70°C
ethylene chloride
Microwaves 100 min 91-98%
Conventional heating 100 min 54-65%
KSCN / TBAB
R-Br R-SCN R-NCS Eq. (64)
4 min - MW
support % R-SCN % R-NCS Σ Yield
SiO2 92 8 36
K10 95 5 30
Graphite 71 29 68
NaCl 52 48 94
NC R R
NaOH / PEG 400
Eq. (65)
2 min - MW
5.3
Conclusion
The use of microwave irradiation to provide the activation energy for synthetic che-
mistry certainly leads to faster and cleaner reactions when compared to conventional
178 5 Microwave and Phase-transfer Catalysis
References
1 a) Makosza, M.; Fedorynski, M. in: Petit, A. Org. Proc. Res. Dev. 2000, 4,
Handbook of Phase-Transfer Catalysis. 498±504.
Blackie Academic and Professional, 13 Xu,W.G.; Liu, F.A.; Yu,Y.X.; Jin, S.X.;
London, 1997, Chapt. 4, pp. 135±167; Liu, J.; Jin, Q.H. Chem. Res. Chin.
b) Makosza, M.; Fedorynski, M. Polish Univ. 1992, 8, 324±326; Chem. Abstr.
J. Chem. 1996, 70, 1093; c) Makosza, M. 1993, 118, 233448x.
Survey Prog. Chem. 1980, 9, 1. 14 Jiang,Y.L.; Yuan,Y.C.; Sun,Y.H. Chem.
2 Starks, C.M.; Liotta, C.L.; Hal- Res. Chin. Univ. 1994, 10, 159±162;
pern, M. in: Phase Transfer Catalysis. Chem. Abstr. 1994, 121, 300548g.
Chapman and Hall, New York, 1994. 15 Yuncheng,Y.; Julin,Y.; Dabin, G.
3 Brandstrom, A. in: Preparative Ion Synth. Commun. 1992, 22, 3109±3114.
Pair Extraction, an Introduction to Theory 16 Yuan,Y.C.; Jiang,Y.L.; Gao, D.B. Chin.
and Practice. Apotekarsocieteten-Hassle Chem. Lett. 1992, 3, 613±614; Chem.
Lakemedel, Stockholm, 1974. Abstr. 1993, 118, 38524s.
4 Dehmlow, E.V.; Dehmlow, S.S. in: 17 Yuan,Y.C.; Jiang,Y.L.; Pang, J.;
Phase Transfer Catalysis, 3rd edn.,Verlag Zhang, X.H.; Yang, C.G. Gazz. Chim.
Chemie, Weinheim, 1993. Ital. 1993, 123, 519±520.
5 Sasson,Y.; Neumann, R. in: Handbook 18 Bogdal, D.; Pielichowski, J.; Jas-
of Phase Transfer Catalysis, Chapman kot, K. Org. Prep. Proced. Int. 1998, 30,
and Hall, London, 1997. 427±432.
6 Loupy, A.; Petit, A.; Hamelin, J.; 19 Majdoub, M.; Loupy, A.; Petit, A.,
Texier-Boullet, F.; Jacquault, P.; Roudesli, S. Tetrahedron 1996, 52, 617±
Math, D. Synthesis 1998, 1213±1234. 628.
7 Perreux, L.; Loupy, A. Tetrahedron 20 Chatti, S.; Bortolussi, M.; Loupy, A.
2001, 57, 9199±9223. Tetrahedron Lett. 2000, 41, 3367±3370.
8 Deshayes, S.; Liagre, M.; Loupy, A.; 21 Chatti, S.; Bortolussi, M.; Loupy, A.;
Luche, J.L.; Petit, A. Tetrahedron 1999, Blais, J.C.; Bogdal, D.; Majdoub, M.
55, 10851±10870. Eur. J. Chem. 2002, 38, 1851±1861.
9 Loupy, A.; Pigeon, P.; Ramdani, M. 22 Campbell, L.J.; Borges, L.F.; Held-
Tetrahedron 1996, 52, 6705±6712. rich, F.J. Biomed. Chem. Lett. 1994, 4,
10 Bram, G.; Loupy, A.; Majdoub, M. 2627±2630.
Synth. Commun. 1990, 20, 125±129. 23 a) Bogdal, D.; Pielichowski, J.;
11 Loupy, A.; Petit, A.; Ramdani, M.; Yva- Boron, A. Synth. Commun. 1998, 28,
naef, C.; Majdoub, M.; Labiad, B.; Vil- 3029±3039; b) Bratulescu, G.; Le Bi-
lemin, D. Can. J. Chem. 1993, 71, 90± got,Y.; Delmas, M.; Pogany, I. Rev.
95. Roum. Chim. 1998, 43, 321±326.
12 Clophax, J.; Liagre, M.; Loupy, A.; 24 Jiang,Y.L.; Hu,Y.Q.; Pang, J.;
References 179
Yuan,Y.C. J. Am. Oil Chem. Soc. 1996, guez, M.C. Tetrahedron Lett. 1998, 39,
73, 847±850. 6053±6056.
25 Li, J.; Pang, J.; Cao, G.; Xi, Z. Synth. 45 Bogdal, D.; Pielichowski, J.; Jas-
Commun. 2000, 30, 1337±1342. kot, K. Heterocycles 1997, 45, 715±722.
26 Varma, R.S. Green Chem. 1999, 1, 43± 46 Bogdal, D.; Pielichowski, J.; Jas-
55. kot, K. Synth. Commun. 1997, 27,
27 Mariani, E.; Genta, M.T.; Bargagna, 1553±1560.
A.; Neuhoff, C.; Loupy, A.; Petit, A. 47 Rodriguez, H.; Prez, R.; Suarez, M.;
in: Application of the Microwave Technol- Lam, A.; Cabrales, N.; Loupy, A.
ogy to Synthesis and Material Processing, Heterocycles 2001, 55, 291±301.
E. Mucchi (ed.), Modena, 157±167, 48 Jiang,Y.; Wang,Y.; Deng, R.; Mi, A.
2000. A.C.S. Symp. Ser. 1997, 659, 203±213.
28 Villa, C.; Genta, M.T.; Bargagna, A.; 49 Wang,Y.L.; Jiang,Y.Z. Synth. Commun.
Mariani, E.; Loupy, A. Green Chem. 1992, 22, 2287±2291.
2001, 3, 196±200. 50 Deng, R.H.; Mi, A.Q.; Jiang,Y.Z. Chin.
29 Wang, J.X.; Zhang, M.; Hu,Y. Synth. Chem. Lett. 1993, 4, 381±384; Chem.
Commun. 1998, 28, 2407±2413. Abstr. 1993, 119, 271670s.
30 Reddy,Y.T.; Rao, M.K.; Rajitha, B. In- 51 Deng, R.H.; Jiang,Y.Z. Hecheng
dian J. Heterocycl. Chem. 2000, 10, 73±74. Huaxue 1994, 2, 83±85; Chem. Abstr.
31 Wang, J.X.; Zhang, M.; Huang, D.; 1994, 121, 108127c.
Hu,Y. J. Chem. Res. (S) 1998, 216±217. 52 Deng, R.H.; Wang,Y.L.; Jiang,Y.Z.
32 Pchelka, B.; Plenkiewicz, J. Org. Synth. Commun. 1994, 24, 1917±1921.
Prep. Proced. Int. 1998, 30, 87±90. 53 Deng, R.H.; Wang,Y.L.; Jiang,Y.Z.
33 Bogdal, D.; Warzala, M. Tetrahedron Synth. Commun. 1994, 24, 111±115.
2000, 56, 8769±8773. 54 Wang,Y.; Deng, R.; Mi, A.; Jiang,Y.
34 Bogdal, D.; Warzala, M., unpublished Synth. Commun. 1995, 25, 1761±1764.
data. 55 Abramovich, R.A.; Shi, Q.; Bog-
35 Hurduc, N.; Abdelylah, D.; Buisi- dal, D. Synth. Commun. 1995, 25, 1±7.
ne, J.M.; Decock, P.; Surpateanu, G. 56 Barbry, D.; Pasquier, C.; Faven, C.
Eur. Polym. J. 1997, 33, 187±190. Synth. Commun. 1995, 25, 3007±3013.
36 Ding, J.; Gu, H.; Wen, J.; Lin, C. 57 Rissafi, B.; Rachiqi, N.; El Louzi, A.;
Synth. Commun. 1994, 24, 301±303. Loupy, A.; Petit, A.; Fkih-Tetouani, S.
37 Huang, Z.Z.; Zu, L.S. Org. Prep. Tetrahedron 2001, 57, 2761±2768.
Proced. Int. 1996, 28, 121±123. 58 AbenhaÒm, D.; Loupy, A.; Mun-
38 Wu, L.L.; Huang, X. Hecheng Huaxue nier, L.; Tamion, R.; Marsais, F.; Qu-
1997, 5, 179±181; Chem. Abstr. 1998, guiner, G. Carbohydr. Res. 1994, 261,
128, 204852a. 255±266.
39 Ding, J.; Yang, J.; Fu, M. Hecheng 59 Chatti, S.; Bortolussi, M.; Loupy, A.
Huaxue 1997, 5, 309±310; Chem. Abstr. Tetrahedron 2000, 56, 5877±5883.
1998, 128, 230343e. 60 Chatti, S.; Bortolussi, M.; Loupy, A.
40 Zhang, X.H.; You,Y.E.; Guo, M. Tetrahedron 2001, 57, 4365±4370.
Hecheng Huaxue 1998, 2, 220±222; 61 Soukri, M.; Guillaumet, G.; Bes-
Chem. Abstr. 1998, 129, 230514g. son, T.; Aziane, D.; Aadil, M.; El Es-
41 Bogdal, D. Molecules 1999, 3, 333. sassi, M.; Aksira, M. Tetrahedron Lett.
42 Bogdal, D.; Pielichowski, J.; 2000, 41, 5857±5860.
Boron, A. Synlett 1996, 873±874. 62 AbenhaÒm, D.; Chu Pham Ngoc Son;
43 Vass, A.; Toth, J.; Pallai-Varsanyi, E. Loupy, A.; Nguyen Ba Hiep Synth.
in: Effect of Inorganic Solid Support for Commun. 1994, 24, 1199±1205.
Microwave Assisted Organic Reactions, 63 Villemin, D.; Martin, B.; Puciova, M.;
OR 19, presented at the Int. Conf. Mi- Toma, S. J. Organomet. Chem. 1994, 484,
crowave Chemistry, Prague, Czech Re- 27±31.
public, Sept. 6±11, 1998. 64 Loupy, A.; Pigeon, P.; Ramdani, M.;
44 de la Cruz, P.; de la Hoz, A.; Jacquault, P. Synth. Commun. 1994,
Font, L.M.; Langa, F.; Prez-Rodri- 24, 159±165.
180 5 Microwave and Phase-transfer Catalysis
65 Dietrich, B.; Lehn, J.M. Tetrahedron 79 Oussaid, A.; Le Ngoc, T.; Loupy, A.
Lett. 1973, 14, 1225±1228. Tetrahedron Lett. 1996, 52, 2451±2454.
66 Perreux, L.; Loupy, A. Tetrahedron 80 Wang, J.X.; Bai, L.; Li,W.; Hu,Y. Synth.
2001, 57, 9199±9223. Commun. 2000, 30, 325±332.
67 Limousin, C.; Clophax, J.; Loupy, A.; 81 Wang, J.X.; Bai, L.; Liu, Z. Synth. Com-
Petit, A. Tetrahedron 1998, 54, 13567± mun. 2000, 30, 971±977.
13578. 82 Bendale, P.M.; Khadilkar, B.M.
68 Loupy, A.; Le Ngoc, T. Synth. Commun. Synth. Commun. 2000, 30, 1713±1718.
1993, 23, 2571±2577. 83 Rissafi, B.; El Louzi, A.; Loupy, A.;
69 Chen, X.; Hong, P.J.; Dai, S. Huaxue Petit, A.; Soufiaoui, M.; Fkih-T-
Tongbao 1993, 11, 29±30; Chem. Abstr. touani, S. Eur. J. Org. Chem. 2002,
1994, 121, 56753g. 2518±2523.
70 Diaz-Ortiz, A.; Diez-Barra, E.; de la 84 Toke, L.; Hell, G.T.; Szabo, G.;
Hoz, A.; Prieto, P. Synth. Commun. Toth, G.; Bihari, M.; Rockenbauer, A.
1993, 23, 1935±1942. Tetrahedron 1993, 49, 5133±5146.
71 Diaz-Ortiz, A.; Prieto, P.; Aben- 85 Finta, Z.; Hell, L.; Toke, L. J. Chem.
haÒm, A.; Loupy, A. Tetrahedron Lett. Res. (S) 2000, 242±244.
1996, 52, 6705±6708. 86 Larhed, M.; Lindenberg, G.; Hall-
72 Yuan,Y.C.; Gao, D.B.; Jiang,Y.L. Synth. berg, A. Tetrahedron Lett. 1996, 37,
Commun. 1992, 22, 2117±2119. 8219±8222.
73 Pang, J.; Xi, Z.; Cao, G. Synth. Com- 87 Wang, J.X.; Liu, Z.; Hu,Y.; Wie, B.;
mun. 1996, 26, 3425±3429. Bai, L. J. Chem. Res. (S) 2000, 484±485.
74 Chaouchi, M.; Loupy, A.; Marque, S.; 88 Sato, K.; Aoki, M.; Takagi, R.;
Petit, A. Eur. J. Org. Chem. 2002, Noyori, R. J. Am. Chem. Soc. 1997, 119,
1278±1283. 12386±12387.
75 Loupy, A.; Pigeon, P.; Ramdani, M.; 89 Bogdal, D.; Lukasiewicz, M. Synlett
Jacquault P. J. Chem. Res. (S) 1993, 2000, 143±145.
36±37. 90 Bogdal, D.; Lukasiewicz, M., in pre-
76 Barnier, J.P.; Loupy, A.; Pigeon, P.; paration.
Ramdani, M.; Jacquault, P. J. Chem. 91 Khadilkar, B.M.; Bendale, P.M. in
Soc. Perkin Trans I 1993, 397±398. Microwave assisted reductive decyanation
77 Bougrin, K.; Soufiaoui, M.; Loupy, A.; of alkyldiphenylmethanes, OR 24, pre-
Jacquault, P. New J. Chem. 1995, 19, sented at the Int. Conf. Microwave
213±219. Chemistry, Prague, Czech Republic,
78 Texier-Boullet, F.; Latouche, R.; Ha- Sept. 6±11, 1998.
melin, J. Tetrahedron Lett. 1993, 34,
2123±2126.
181
6
Organic Synthesis Using Microwaves and Supported Reagents
Rajender S. Varma
6.1
Introduction
In the electromagnetic radiation region, microwaves (0.3 GHz±300 GHz) lie be-
tween radiowave (Rf) and infrared (IR) frequencies with relatively large wavelengths
(1 mm±1 m). Microwaves, nonionizing radiation incapable of breaking bonds, are a
form of energy that manifest as heat through their interaction with the medium or
materials wherein they can be reflected (metals), transmitted (good insulators that
will not heat) or absorbed (decreasing the available microwave energy and rapidly
heating the sample). This unconventional microwave (MW) energy source has been
used for heating food materials for almost 50 years [1] and is now being utilized for a
variety of chemical applications including organic synthesis [2±11] wherein chemical
reactions are accelerated because of selective absorption of MW energy by polar mo-
lecules, nonpolar molecules being inert to the MW dielectric loss [12]. The initial ex-
periments with microwave heating exploited the use of high dielectric solvents such
as dimethyl sulfoxide (DMSO) and dimethylformamide (DMF) in a household
kitchen MW oven. The rate enhancements in such reactions are now believed to be
due to rapid superheating of the polar solvents and pressure effects [12]. However, in
these solution-phase reactions, the development of high pressures, and the use of
specialized sealed vessels are some of the limitations, although they have been cir-
cumvented by the introduction of commercial MW instruments with appropriate
temperature and pressure controls.
Heterogeneous reactions facilitated by supported reagents on inorganic oxide sur-
faces have received attention in recent years, both in the laboratory as well as in in-
dustry. Although the first description of the surface-mediated chemical transforma-
tion dates back to 1924 [13], it was not until almost half a century later that the tech-
nique received extensive attention with the appearance of several reviews, books and
account articles [14±22].
A related development that had profound impact on heterogeneous reactions is
the use of microwave (MW) irradiation techniques for the acceleration of organic re-
actions. Since the appearance of initial reports on the application of microwaves for
chemical synthesis in polar solvents [11], the approach has blossomed into a useful
6.2
Microwave-accelerated Solvent-free Organic Reactions
6.2.1
Protection±Deprotection Reactions
OH O
R
OH Clay, RCHO O
O O
O O
HO MW, 10 min, 60–66% HO
OH OH
Scheme 6.1 Formation of acetal derivatives of 1-galactono-1,4-lactone.
R1 OH R1 O
KSF or PTSA
C O +
R OH MW, 10–30 min R O
Scheme 6.2 Formation of dioxolanes.
184 6 Organic Synthesis Using Microwaves and Supported Reagents
Methanesulphonothioate,
R1 KF-Alumina R1 SMe
CH2 C
R MW R SMe
tained with the microwave method are better than those obtained using conventional
heating in an oil bath [32].
Thioacetals have been prepared using essentially a similar technique [33]. The ac-
tive methylene compounds are adsorbed on KF-alumina, admixed with methanesul-
fonothioate and are irradiated in microwave oven to produce thioacetals in good
yields (Scheme 6.3).
HO
Alumina MW Alumina
MW
CH(OAc)2 CHO
Neutral alumina
Interestingly, acylal formation has been accomplished with acetic anhydride [37]
on K 10 clay (75±98 %) as well as deacylation [38].
CH2CH2CO2R NH2CHCO2R
CH2OH
3 min/Neutral, (92%)
(10 min)
OH
10 min/Acidic, (92%)
NH2CH2CO2R
4 min/Neutral, (95%)
(10 min)
Where R= CH2C6H5 and time in parentheses refer to deprotection in oil bath at the same
temperature
R1
Lewis acid R1
N O NH
R2 But MW R2
Scheme 6.7 Deprotection of N-tert-
O butoxycarbonyl group.
OR OR OR
(10 min, 93%) (11 min, 78%)
(10 min, 91%)
O NH2
CH3 N
HN N
RO O N RO N N
O O
OH OH
CH3
CH3
R1 S X R1
Clayfen
C C O
R2 S Y Room Temp. or MW R2
(87–98%)
Where R1 = Ph, p-CH3 C6H4, p-NO2C6H4; R2 = H ; X–Y = –(CH2)2–
R 1 = R 2 = C 2H5 ; X–Y = –(CH2)2– ; R1 = R2 = Ph ; X = Y = C2H5
R 1 = Ph ; R2 = CH3 ; X–Y = –(CH2)2–
R 1–R2 = isoflavanolyl , 2-Methylcyclohexyl; X–Y = –(CH 2 )2–
R1
R1 Silica-Ammonium persulfate
C O
C NOH
MW, 1-2 min. R2
R2 (59-83%)
has continued with the availability of a wide range of such agents namely, Raney
nickel, pyridinium chlorochromate, pyridinium chlorochromate±H2O2, triethylam-
monium chlorochromate, dinitrogen tetroxide, trimethylsilyl chlorochromate,
Dowex-50, dimethyl dioxirane, H2O2 over titanium silicalite-1, zirconium sulfophe-
nyl phosphonate, N-haloamides, and bismuth chloride [47, 48].
The quest for a solvent-free deprotection procedure has led to the use of relatively
benign reagent, ammonium persulfate on silica, for regeneration of carbonyl com-
pounds (Scheme 6.10) [48]. Neat oximes are simply mixed with solid supported re-
agent and the contents are irradiated in a MW oven to regenerate free aldehydes or ke-
tones in a process that is applicable to both, aldoximes and ketoximes. The critical
role of surface needs to be emphasized since the same reagent supported on clay sur-
face delivers predominantly the Beckmann rearrangement products, the amides [49].
R1
R1
Silica-Moist NaIO4
C O
C NOH
MW, 1-2.5 min. R2
R2 (68-93%)
Where R 1 = CH3 ; R2 = Ph, p-Cl C 6H4, p-Br C6H4, p-CH3 C 6H4, p-CH3O
C6H4; p-NH2 C 6H4,
R1 = R2 = Ph; R1 = R2 = Cyclohexyl, tetrahydronaphthyl,
and R 1 = C 2H5 , R2 = n-Butyl
R1 (NH4)2S2O8 - Clay R1
C N NH R C O
R2 MW or ))))) R2
(65-94%)
Where R 1 = C 4H9, Ph, p-Cl C6H4, p-CH3 C 6H4, p-OH C6H4;
R 2 = CH3; C 2H5
and R = CONH2 , Ph
Scheme 6.12 Regeneration of carbonyls from semicarbazone and
phenylhydrazone derivatives.
6.2.1.11 Dethiocarbonylation
Dethiocarbonylation, transformation of thiocarbonyls to carbonyls, has been accom-
plished with several reagents namely, trifluoroacetic anhydride, CuCl/MeOH/
NaOH, tetrabutylammonium hydrogen sulfate/NaOH, clay/ferric nitrate, NOBF4,
bromate and iodide solutions, alkaline hydrogen peroxide, sodium peroxide, thio-
phosgene, trimethyloxonium fluoroborate, tellurium based oxidants, dimethyl selen-
oxide, benzeneseleninic anhydride, benzoyl peroxide and halogen-catalyzed alkox-
ides under phase transfer conditions [55]. However, these methods have certain lim-
itations such as the use of the stoichiometric amounts of the oxidants that are often
inherently toxic or require longer reaction time or involve tedious procedures. In a
process that is accelerated by microwave irradiation,Varma et al. have demonstrated
efficient dethiocarbonylation process under solvent-free conditions using clayfen or
clayan (Schs. 6.13 and 6.14) [55].
S
O
R2
Clayfen or (Clayan) R2
1-1.5 Min
R1
R1 (90-95 %)
O R1 O R1
Clayfen or (Clayan)
1.5-2 Min
R R
S (85-91 %) O
MW
R OMPM R OH
Clayan
R = alkyl, alkenyl, alkynyl, aryl, acetate, ester, benzyl or silyl ether groups
Scheme 6.15 Cleavage of methoxyphenyl methyl (MPM) ethers using clayan.
O
KF-Al2O3 R1
R1 KF-Al2O3 R1
NH X S Ph CH OH
R2 R2 R2
O (83–90%)
(76–85%)
Where X = N, CH-O-
Scheme 6.16 Cleavage of sulfonates and sulfonamides on a basic KF-alumina surface.
6.2 Microwave-accelerated Solvent-free Organic Reactions 191
6.2.2
Condensation Reactions
A wide variety of MW-assisted aldol [59, 60] and Knoevenagel condensation reactions
have been accomplished using relatively benign reagents such as ammonium acetate
[61], including the Gabriel synthesis of phthalides with potassium acetate [62].
(Ph)3 P=CHCOOEt
Me R Me
N N
MW H
NH + RCHO NH
O N O N
H H
Scheme 6.18 Knoevenagel condensation reaction of creatinine with alde-
hydes.
192 6 Organic Synthesis Using Microwaves and Supported Reagents
CHO R3
R3 Base (Piperidine)
+
CO2Et MW
R1 OH R1 O O
R2 R2
CH
CH O H2 N K 10 Clay or EPZG N
+
MW, 1–3 min.
R R
(90–97%)
Y Y
O ( )n ( ) n2
Y EPZG or K 10 Clay 2
N H2 O N
( )n
1
+ ( )n
2 MW, 2–6 min.
OH
N ( )n ( )n
H 1 1
CHO NO2
NH4OAc
+ RCH2NO2
MW, 2.5–8 min. R
X X (80–92%)
Where R = H, X = H, p-OH, m,p-(OMe)2, m-OMe-p-OH, 1-naphthyl, 2-naphthyl
R = Me, X = H, p-OH, p-OMe, m,p-(OMe)2, m-OMe-p-OH
Scheme 6.22 MW-assisted preparation of a,b-unsaturated nitroalkenes.
R1 R1
CHO CaCO3/K 10 Clay CH=N SO2
R R
+
CH(OMe) 3, MW, 30-70 W
H2NSO2 (52 - 91%)
R2 R2
R R
O O O O
R1 NHNH2 R
R
MW
NHCOPh NHCOPh
O N
R1 NH
MW irradiation conditions [80]. More recently, however, Varma and Kocevar's group
have shown that a solvent-free and catalyst-free reaction of hydrazines with carbonyl
compounds is possible upon MW irradiation (Scheme 6.24) [81]. Interestingly, the
general reaction proceeds smoothly even for solid reactants and is completed below
the melting points of the two reactants possibly via the formation of a eutectic. The
reactions have been conducted in household MW oven and the control experiments
are conducted concurrently in separate open beakers; the reactions can be essentially
followed by visual observation when a melt is obtained [82].
An interesting solid-state synthesis of amides has also been reported in a MW
oven that uses potassium tert-butoxide and easily accessible reagents, nonenolizable
esters and amines [83].
The kinetics of the acid-catalyzed esterification reaction of 2,4,6-trimethylbenzoic
acid in i-PrOH under microwave irradiation have been investigated [84]. A simple
and practical technique for MW-assisted synthesis of esters has been reported
wherein the reactions are conducted either on solid mineral supports or by using a
phase transfer catalyst (PTC) in the absence of organic solvents [85]. The esterifica-
tion of enols with acetic anhydride and iodine has also been recorded [86].
The detailed account of condensation reactions as applied to heterocyclic chemis-
try is found in subsequent section (Sect. 6.2.6), Chapt. 8 [87] and for cycloaddition re-
actions in Chapt. 9 [88].
6.2.3
Isomerization and Rearrangement Reactions
Me Me
Me OH Al 3+ –Montmorillonite Me O
MW, 15 min Me
HO Me Me
Me
(98-99%)
Scheme 6.25 Pinacol±pinacolone rearrangement on Al3+-montmorillonite
K 10 clay.
SC6H5 SC6H5
Cl AgBF4–Al2O3
Cl
Cl MW, 10 min
(75%)
Scheme 6.26 MW-assisted ring expansion reaction on alumina.
R1 O
Montmorillonite K 10 clay
C N OH R1 C NH R2
R2 MW, 7-10 min.
Where R1= CH3 or Ph and R 2 = Ph or substituted phenyl
Scheme 6.27 Beckmann rearrangement of oximes on clay.
196 6 Organic Synthesis Using Microwaves and Supported Reagents
6.2.4
Oxidation Reactions ± Oxidation of Alcohols and Sulfides
Metal-based reagents have been extensively used in organic synthesis. Peracids, per-
oxides, potassium permanganate (KMnO4), manganese dioxide (MnO2), chromium
trioxide (CrO3), potassium dichromate (K2Cr2O7) and potassium chromate (K2CrO4)
are some of the common oxidizing reagents employed for organic functional groups
[94, 95].
The utility of such reagents in the oxidation processes is compromised due to their
inherent toxicity, cumbersome preparation, potential danger in handling of metal
complexes, difficulties encountered in product isolation and waste disposal pro-
blems. Immobilization of metallic reagents on solid supports has circumvented
some of these drawbacks and provided an attractive alternative in organic synthesis
because of the selectivity and associated ease of manipulation. Further, the localiza-
tion of metals on the mineral oxide surfaces reduces the possibility of their leaching
into the environment.
R1 MnO2-Silica R1
CH OH C O
R2 MW, 20-60 sec. R2
(67-96%)
R1 CrO3–Wet Al2O3 R1
CH OH C O
R2 MW, 40 sec. R2
(73-90%)
R1 = Ph, p-MeC6H4, p-MeOC 6H4, p-NO2C6H4 ; R2 = H
R1 = Ph ; R2 = Me, Ph, PhCO; R1 = R2 = ,
Scheme 6.29 Oxidation of alcohols by chromium trioxide supported on
premoistened alumina.
R1 Clayfen R1
CH OH C O
R2 MW, < 1 Min. R2
(87-96%)
R1 H2O2-Claycop R1
CH R3 C O
R2 MW R2
(71-85%)
OH O
R2 Oxone –Al2O3 or CuSO4–Al2O3 R2
R1 MW, 2-3.5 min. R1
O O
(71-96%)
Under these solvent-free conditions, the oxidation of primary alcohols (e. g. benzyl
alcohol) and secondary alcohols (e. g. 1-phenyl-1-propanol) is rather sluggish and
poor and is of little practical utility. Consequently, the process is applicable only to
a-hydroxyketones as exemplified by various examples including a mixed benzylic/ali-
phatic a-hydroxyketone, 2-hydroxypropiophenone that delivers the corresponding
vicinal diketone [106, 107].
O
20% NaIO 4-Silica (3.0 eq.) 20% NaIO 4-Silica (1.7 eq.)
R1-SO2-R2 R 1-S-R2 R1-S-R2
MW, 1-3 min. MW, 0.5-2.5 min.
(72-93%) (76-85%)
[109]. A relatively reduced amount of the active oxidizing agent is employed which is
safer and easier to handle.
Several refractory thiophenes, that are often not reductively removable by conven-
tional refining processes, can be oxidized under these conditions, e. g. benzothio-
phenes are oxidized to the corresponding sulfoxides and sulfones using ultrasonic
and microwave irradiation, respectively, in the presence of NaIO4±silica [109]. A note-
worthy feature of the procedure is its applicability to long chain fatty sulfides that
are insoluble in most solvents and are consequently difficult to oxidize.
O
IBD-Alumina
R1—S—R2 R—S—R2
MW, 40-60 sec.
(80-90%)
where R1 = R2 = i-Pr, n-Bu, Ph, PhCH2 ; R 1 = Ph ; R2 = Me, PhCH2
R1 = n-C 12H25 ; R1 = Me ; R1 = R2 = , O
Scheme 6.35 Oxidation of sulfides to sulfoxides by alumina-supported
iodobenzene diacetate.
O
O
N R1 KMnO4- Al2O3 R1
N + O
H R2 MW
R2
CHO
Scheme 6.36 Oxidation of enamines with alumina-supported potassium
permanganate.
6.2 Microwave-accelerated Solvent-free Organic Reactions 201
6.2.5
Reduction Reactions
In MW-assisted chemistry reduction reactions were the last to appear on the scene;
the use of ammonium formate and catalytic transfer hydrogenation were initial ex-
amples [23 b].
R1 Me R1 Me
Aluminum alkoxide
O + CHOH CHOH + O
R2 Me MW R2 Me
Scheme 6.37 Solvent-free reduction of carbonyls using aluminum alkoxides.
O OH
NaBH4-Al2O3
R1 C R2 R1 CH R2
MW, 0.5–2 min
(62–93%)
The useful chemoselective feature of the reaction is apparent from the reduction
of trans-cinnamaldehyde (cinnamaldehyde/NaBH4±alumina, 1 : 1 mol equivalent);
olefinic moiety remains intact and only the aldehyde functionality is reduced in a fa-
cile reaction.
No side products are formed and the reaction does not proceed in the absence of
alumina. Further, the reaction rate improves in the presence of moisture. The moist-
ure is absorbed by alumina during the recovery of the product. The alumina support
can be recycled and reused for subsequent reduction, repeatedly, by mixing with
fresh borohydride without any loss in activity. In terms of safety, the air used for cool-
ing the magnetron ventilates the microwave cavity, thus preventing any ensuing hy-
drogen from reaching explosive concentrations. The technique has been elegantly
utilized for MW-enhanced solid-state deuteration using sodium borodeuteride im-
pregnated alumina [120]. Further extension of this work in the specific labeling of
molecules has been explored [121] and is discussed elsewhere in this book [122].
Ba(OH)2 . 8H2O
RCHO + (CH2O)n RCH2OH + RCOOH
MW or oil bath
(80-99%) (1-20%)
Scheme 6.40 Solvent-free crossed Cannizzaro reaction using paraformaldehyde.
204 6 Organic Synthesis Using Microwaves and Supported Reagents
6.2.6
Synthesis of Heterocyclic Compounds
6.2.6.1 Flavones
Naturally occurring flavonoids are widely distributed oxygen heterocyclic com-
pounds in plant kingdom, the most abundant being the flavones. Members of this
class display a wide variety of biological activities and have been useful in the treat-
ment of various diseases [135, 136]. Flavones have been prepared by a variety of
methods such as Allan±Robinson synthesis, and synthesis from chalcones via an in-
tramolecular Wittig strategy [137]. The commonly followed approach, however, in-
volves the Baker±Venkataraman rearrangement, wherein o-hydroxyacetophenone is
benzoylated to form the benzoyl ester followed by the treatment with base (KOH/pyr-
idine) to effect acyl migration, forming a 1,3-diketone [138, 139]. The ensuing dike-
tone is then cyclized under strongly acidic conditions with sulfuric acid and acetic
acid to afford flavone.
A solvent-free synthesis of flavones has been achieved that simply involves the
MW irradiation of o-hydroxydibenzoylmethanes adsorbed on montmorillonite K 10
clay for 1±1.5 min. A rapid and exclusive formation of cyclized flavones occurs in
good yields (Scheme 6.41) [140]. The intramolecular Michael addition of o-hydroxy-
chalcones on silica gel surface has also been reported [141].
R2
OH R2 O
K 10 Clay
K 10 Clay
C C O + HN R C C N R
MW
H R R
(95–98%) Enamines
R4 R4
OH R R3 O R
R3 NH4OAc
+ H
R2 MW, 2–8 min R2 Ph
CHO Ph
R1 R1
(73–88%)
Where R = morpholinyl, piperidinyl or pyrrolidinyl
and R 1 = R3 = R4 = H ; R2 = H, Cl, NO2
Scheme 6.42 One-pot synthesis of 2-substituted isoflav-3-enes from in situ- generated
enamines.
wherein the in situ generated enamine derivatives have been subsequently reacted
with o-hydroxyaldehydes in the same pot (Scheme 6.42) [145].
R1
O R2
OTs N
K 10 clay
+
MW H S
R1
R2
H2N S (88-96%)
Where R1 = H, Me, OMe, Cl and R2 = Cl, OMe
Scheme 6.43 Synthesis of substituted thiazoles from in situ-generated a-tosyloxyketones.
O OTS
H
K 10 clay N
+ HN NH
MW N
S
S
R1
R1 (85-92%)
Where R1 = H, Me, OMe, Cl
Scheme 6.44 Synthesis of bridgehead thiazoles from a-tosyloxyketones.
O O
X
Y
X Y R2 Lawesson's reagent
S
MW, 3-8 min R2
H H
R1 R1
(83-94%)
Where R1 = H, OMe, Br; R2 = Ph, O-alkyl
and X = Y = CH or N
Scheme 6.45 MW-assisted synthesis of S-heterocycle-containing liquid crystalline compounds.
O
TsO
R1 CHO R1 R2
KF-Al2O3
+ MW
OH O
R2 (89-96%) O
R1 R1
H
NH2 N
R2 K 10 clay R2
R
NH2 N
R CHO R1
N X Microwave N N X
+ Y Clay
H
Y
R1 NC
X= Y= C X= Y= C
X= C, Y= N X= C, Y= N
X= N, Y=C X= N, Y=C
Scheme 6.48 Synthesis of imidazo[1,2-a]annulated N-heterocycles via the Ugi reaction.
A similar strategy has been used for the Biginelli condensation reaction to synthe-
size a set of pyrimidinones (65±95 %) in a household MW oven [152]. This MW ap-
proach has been successfully applied to combinatorial synthesis [153]. Yet another
example is the convenient synthesis of pyrroles (60±72 %) on silica gel using readily
available enones, amines and nitro compounds [154].
6.2.7
Miscellaneous Reactions
CHO CN
NH2OH.HCl / K 10 clay
MW, 1-1.5 min
R2 R2
R1 R1
Where R1 = H, R2 = H, OCH3, NO2, OH, Br, CH3 (89-95 %)
R1 = R2 = OCH3
Scheme 6.49 Transformation of arylaldehydes to nitriles by hydroxylamine hydrochloride±clay.
6.2 Microwave-accelerated Solvent-free Organic Reactions 209
NO 2
CHO
Clayfen (Clayan)
+
MW or oil bath
(minor)
R R R
Where R = H, Cl, CH3, OCH 3
Scheme 6.50 Preparation b-nitrostyrenes using clay-supported nitrate salts.
H R
PdL2X2 + -
C C + RX + Base C C + BaseH X
Scheme 6.51 Palladium-catalyzed carbon±carbon bond forming reactions using microwaves.
210 6 Organic Synthesis Using Microwaves and Supported Reagents
R2
I
Pd-CuI-PPh3
KF-Al 2O3
+ R2 H
MW
R1 R1
D
_ _
NH3+ Cl ND3+Cl ND3+ Cl
_
D2 O D2O
MW
CH3 CH3
D CH3
Scheme 6.53 Deuteration and tritiation reactions using MW irradiation.
6.2 Microwave-accelerated Solvent-free Organic Reactions 211
Me Me Me Me
- + + -
2TCOOH + N(CH2)2N 2TCOO H N(CH2)2N H O2CT
Me Me
Me Me
Scheme 6.54 Hydrogenation reactions using formates.
OH OH
H H OCOR'
Me Lipase
Me Me
+ R'CO2R +
N N
MW R X
+
PF6 - NH4 PF6
X - NH4 BF4 -
BF4
+ +
N N N N N N
R R MW
R
The approach precludes the usage of volatile organic solvents, is relatively much
faster, efficient, and eco-friendly. Significant rate enhancements are reported in the
1,3-dipolar cycloaddition reactions including the use of covalently grafted dipolaro-
philes on the ionic liquids [189].
6.3
Conclusions
Microwave heating, being specific and instantaneous, is unique and has found a
place in the expeditious chemical synthesis domain. Specifically, the solvent-free re-
References 213
actions are convenient to perform and have demonstrated clear advantage over the
conventional heating procedures as summarized in this chapter. Numerous selective
organic functional group transformations have been accomplished more efficiently
and expeditiously using a variety of supported reagents on mineral oxides as cata-
lysts. Although a large body of work has been performed around the world using an
unmodified household microwave oven (multimode applicator), more recent work
does point out the advantages of using commercial systems wherein not only the im-
proved temperature/power control is possible but also the relatively large-scale reac-
tions can be conducted [190, 191] with additional options for a continuous operation.
The engineering and scale-up aspects for the chemical process development have
also been discussed [192].
There are distinct advantages of these solvent-free procedures in instances where
catalytic amounts of reagents or supported agents are used since they provide reduc-
tion or elimination of solvents, thus preventing pollution ªat sourceº. Although not
delineated completely, the reaction rate enhancements achieved in these methods
may be ascribable to nonthermal effects. The rationalization of microwave effects
and mechanistic considerations are discussed in detail elsewhere in this book [25,
193]. A dramatic increase in the number of publications [23 c], patents [194±203],
a growing interest from pharmaceutical industry, with special emphasis on combina-
torial chemistry, and development of newer microwave systems bodes well for micro-
wave-enhanced chemical syntheses.
References
1 C. R. Buffler, Microwave Cooking and 8 R. S. Varma, Pure Appl. Chem. 2001, 73,
Processing. Van Nostrand Reinhold, 193.
New York, 1993, pp. 1±68. 9 F. Langa, P. de la Cruz, A. de la Hoz,
2 R. S. Varma in: ACS Symposium Series A. DÌaz-Ortiz, E. DÌez-Barra,
No. 767/Green Chemical Syntheses and Contemp. Org. Chem. 1997, 4, 373.
Processes. P. T. Anastas, L. Heine, 10 A. Loupy, A. Petit, J. Hamelin,
T. Williamson, (eds.), American Che- F. Texier-Boullet, P. Jacquault,
mical Society, Washington DC, 2000, D. Math, Synthesis 1998, 1213.
Chapt. 23, pp. 292±313. 11 a) R. Gedye, F. Smith, K. Westaway,
3 R. S. Varma in: Green Chemistry: Chal- H. Ali, L. Baldisera, L. Laberge,
lenging Perspectives. P. Tundo, P. T. Ana- J. Rousell, Tetrahedron Lett. 1986, 27,
stas, (eds.), Oxford University Press, 279; b) R. J. Giguere, A. M. Namen,
Oxford, 2000, pp. 221±244. B. O. LÕpez, A. Arepally, D. E. Ramos,
4 R. S. Varma, Green Chem. 1999, 1, 43. G. Majetich, J. Defauw, Tetrahedron
5 R. S. Varma, Clean Products Proc. 1999, Lett. 1987, 28, 6553.
1, 132. 12 C. Gabriel, S. Gabriel, E. H. Grant,
6 R. S. Varma in: Microwaves: Theory and B. S. J. Halstead, D. M. P. Mingos,
Application in Material Processing IV. Chem. Soc. Rev. 1998, 27, 213.
D. E. Clark,W. H. Sutton, D. A. Le- 13 Using chemical reagents on porous car-
wis, (eds.), American Ceramic Society, riers, Akt.-Ges. Fur Chemiewerte. Brit.
Westerville, Ohio, 1997, pp. 357±365. Pat. 1924, 231,901. (Chem. Abstr. 1925,
7 a) R. A. Abramovich, Org. Prep. Proc. 19, 3571).
Int. 1991, 23, 683; b) S. Caddick, Tetra- 14 A. McKillop, K. W. Young, Synthesis
hedron 1995, 51, 10403. 1979, 401 and 481.
214 6 Organic Synthesis Using Microwaves and Supported Reagents
7
Microwave-assisted Reactions on Graphite
Andr Laporterie, Julien Marqui, and Jacques Dubac
7.1
Introduction
During the last 15 years numerous papers dealing with the use of microwave (MW)
irradiation, rather than conventional heating, in organic and inorganic chemistry
have reported dramatic reductions of reaction time and significant enhancement of
yields and purity of the products. Despite the possibility of operating with pressur-
ized reactors [1], however, MW irradiation of chemical reactions involving low boil-
ing reagents and/or products can involve serious safety problems. Consequently,
MW-assisted solvent-free reactions (ªdry mediaº) have been widely investigated in or-
ganic synthesis [2]. Among the materials most often used as supports are alumina,
silica, clays, and zeolites, which are sometimes also used as catalysts. When properly
dried, however, these materials are weak MW absorbers and poor thermal conduc-
tors. For reactions which require high temperatures the idea of using a reaction sup-
port which takes advantage both of strong MW coupling and strong adsorption of or-
ganic molecules has stimulated great interest. Because most organic compounds do
not interact appreciably with MW radiation, such a support could be an ideal ªsensi-
tizerº, able to absorb, convert, and transfer energy provided by a MW source to the
chemical reagents.
Most forms of carbon, except diamond, which are renowned as supports for pre-
cious metal catalysts in certain applications [3], interact strongly with MW [4]. Amor-
phous carbon and graphite, in their powdered form, irradiated at 2.45 GHz, rapidly
(within 1 min) reach very high temperatures (>1300 K). This property has been used
to explain MW-assisted syntheses of inorganic solids [5]. In these syntheses carbon is
either a ªsecondary susceptorº which assists the initial heating but does not react
with other reactants, or is one of the reactants, e. g. in the synthesis of metal car-
bides. MW±carbon coupling has also been widely developed:
. by Wan et al. for gas-phase reactions; for example, in the synthesis of hydrogen cy-
anide from ammonia and carbon or methane [6], in the MW-induced catalytic re-
action of water and carbon [7], and in the removal and/or destruction of acid gas-
eous pollutants such as SO2 and NOx [7, 8]; and
. for the processing of polymers and composites in which carbon black or graphite
particles or fibers are included in the material [9].
7.2
Graphite as a Sensitizer
7.2.1
Diels±Alder Reactions
Many Diels±Alder (DA) cycloadditions have been studied under MW irradiation [19].
The use of a ªdry processº, as in GS/MW coupling, is of great interest for difficult re-
actions which need high temperatures, particularly for those involving poor MW-ab-
sorbing reagents. Some reactions which normally require use of an autoclave can,
moreover, occur in an open reactor, owing to retention of a possibly volatile reagent
by the graphite.
Among the dienes known as weakly reactive are anthracene (1), metacrolein di-
methylhydrazone (2) and 3,6-diphenyl-1,2,4,5-tetrazine (3). DA cycloadditions with
these dienes require long reaction times under classical heating conditions (Tab. 7.1).
Three reactions of 1 successively with diethyl fumarate, maleic anhydride, and di-
methyl acetylenedicarboxylate (DMAD) are highly representative of the variety of ex-
perimental conditions used in the GS/MW process [26, 27]. Continuous MW irradia-
tion (CMWI) with an incident power of 120 W for 1 min led to a high increase in
temperature (Tmax > 300 8C). Adduct 4 was obtained almost quantitatively (Tab. 7.1,
entry 1), whereas only traces of adducts 5 and 6 were detected. When the incident
power was reduced (30 W) and sequential MW irradiation (SMWI) was used, adducts
5 and 6 were obtained in good yield (Tab. 7.1, entries 3 and 4). This controlled irra-
diation enabled the temperature to be limited (Tmax < 200 8C) and avoided the retro-
DA reaction. In the reaction between 1 and diethyl fumarate similar SMWI condi-
tions also gave the adduct 4 in high yield (Tab. 7.1, entry 2).
Other DA reactions of 1 (and some of its derivatives) in SMWI processes have
been reported [28]. Under powerful irradiation (Tmax > 300 8C), all products decom-
posed by the retro-DA reaction.
222
Tab. 7.1 Diels±Alder reactions of dienes 1±3 using the GS/MW process [15, 16].
The hetero-DA reaction with azadienes, a well known synthetic method for obtain-
ing nitrogen heterocycles, suffers from some difficulties, because of the low reactiv-
ity of the diene. For example, azadiene 2 did not react with DMAD under the action
of conventional heating [22]. Sequential exposure to MW irradiation (30 W) for
10 min on a graphite support (Tmax = 171 8C) led to the adduct 7 with 60 % conver-
sion (50 % in isolated product) [26, 27]. An equivalent yield was obtained by ultra-
sonic irradiation of the neat reaction mixture at 50 8C for 50 h [29].
The DA reaction of tetrazines such as 3 was also studied by use of the GS/
MW process [26, 27]. The expected adduct, however, decomposed by nitrogen
elimination followed by dehydrogenation, giving a pyridazine or a dihydropyrida-
zine [23±25]. With 2,3-dimethylbutadiene and cyclopentadiene as dienophiles,
SMWI gave dihydropyridazines 8 and 9, as with classical heating [23] (Tab. 7.1,
entries 6 and 7).
Under classical conditions, the reaction between 3 and styrene required 50 h of
heating at 110 8C, and gave the dihydropyridazine adduct 10a [24]. After SMWI
with 30 W incident power for 5 min (Tmax = 154 8C), the adduct 10a was not de-
tected whereas the totally dehydrogenated product, pyridazine 10b, was isolated in
almost quantitative yield (Tab. 7.1, entry 8). Ethyl vinyl ether and 3 gave the same
product, pyridazine 11, under both classical heating [25] and MW irradiation con-
ditions (Tab. 7.1, entry 9). In this instance the DA adduct lost nitrogen and etha-
nol.
The synthesis of these adducts was realized in very short times compared with the
same reactions under the action of classical heating. The efficiency of the MW pro-
7.2 Graphite as a Sensitizer 223
cess is all the more noteworthy because the three dienophiles dimethylbutadiene, cy-
clopentadiene and ethyl vinyl ether are volatile. Although an excess of these reagents
with respect to 3 was used, the adsorption power of graphite was responsible of their
retention, because the temperature of the reaction mixture exceeded their boiling
points of approximately 120±130 8C.
This retention of reagents by the graphite support has been shown in a series of
experiments involving volatile dienes such as 2,3-dimethylbutadiene (12) and iso-
prene (13) (Tab. 7.2) [30, 31]. The reaction of 12 with diethyl mesoxalate gave 14 in
75 % yield after SMW irradiation at low power (30 W) for 20 × 1 min only (Tab. 7.2,
entry 1). When conventional heating was used a satisfactory yield was obtained after
4 h at 135 8C in a sealed tube [32]. Ethyl glyoxylate is a weaker carbonyl dienophile
224 7 Microwave-assisted Reactions on Graphite
Tab. 7.2 Diels±Alder reactions with 2,3-dimethylbutadiene (12) and isoprene (13) by use of the
GS/MW process [27, 30, 31].
than diethyl mesoxalate, and a catalyst (ZnCl2) was added to afford the expected DA
adducts with dienes 12 and 13 in good yields (Tab. 7.2, entries 2 and 7). For reaction
with 13 the catalyst SnCl4 [34] was substituted by ZnCl2 to prevent the formation of
the ene reaction product (Tab. 7.2, entry 7). Although adduct 16 was previously pre-
pared from 15 [32], its direct DA synthesis from 12 and glyoxylic acid could be per-
formed under the action of MW and without a catalyst (Tab. 7.2, entry 3).
Reactions performed with methyl vinyl ketone and metacrolein as ethylenic dieno-
philes also showed the clear advantage of SMWI conditions over conventional heat-
ing (Tab. 7.2, entries 4±6) [31]. In the reaction of isoprene with methyl vinyl ketone
(Tab. 7.2, entry 6) selectivity for the para adduct (54 %) was much better than when
conventional heating was used (26 %), probably owing to the shortening of reaction
time.
Another example of the retention of volatile DA reagents is that of cyclopentadiene
in a tandem retro-DA/DA ªprimeº reaction [15, 16, 38]. This reaction type is the ther-
mal decomposition of a DA adduct (A) and the generation of a diene (generally the
initial diene) which is trapped in situ by a dienophile leading to a new adduct (B)
[39]. Cyclopentadiene (22) (b.p. 42 8C) is generated by thermolysis of its dimer at ap-
proximately 160 8C [40]. An equimolar mixture of commercial crude dicyclopenta-
diene (21) and dimethyl maleate was irradiated in accordance with the GS/MW pro-
cess, in an open reactor, under 60 W incident power, for 4 min (8 × 30 s). The ex-
pected adduct 23 was isolated in 40 % yield (Scheme 7.1). The isomeric composition
of 23 (endo±endo/exo±exo = 65/35) was identical with that obtained under classical
conditions from 22 and methyl maleate [41]. The overall yield of this tandem reac-
tion can be increased from pure dimer 21 (61%) and the same tandem reaction has
also been reported using ethyl maleate as dienophile [31].
7.2 Graphite as a Sensitizer 225
Scheme 7.1
The main advantages of the GS/MW process are the rapid increase of tempera-
ture, the retention of organic molecules, and the possibility of performing some re-
actions under one-pot conditions.
7.2.2
Ene Reactions
The ene reaction (or Alder reaction) is a cycloaddition which requires an activation
energy higher than that of the Diels±Alder reaction [41]. Without a catalyst it usually
occurs under pressure and/or at high temperature. The reaction with an alkene
(ene) is much easier if the latter is more substituted (high HOMO) and the enophile
is more electron-poor (low LUMO).
The alkene 1-decene (24) was poorly reactive in the carbonyl±ene reaction with
ethyl mesoxalate and required a temperature up to 170 8C for a very long time (5 h)
[42]. When performed in a homogeneous liquid medium at the same temperature
but under the action of MW irradiation the reaction gave a similar result. Reaction
time was appreciably shortened by use of GS/MW coupling [30]. Thus, irradiation at
60 W for only 10 min led to the ene adduct 25 in 50 % yield (Scheme 7.2). Under
these conditions a maximum temperature of 230 8C was measured. To obtain the
same yield with conventional heating at 170 8C reaction for 1 h is required. The
stereoselectivity of the reaction was not related to the mode of heating. A higher con-
Scheme 7.2
226 7 Microwave-assisted Reactions on Graphite
version of 24 was obtained by increasing the irradiation (incident power >60 W), but
the occurrence of side products made the isolation of 25 more difficult.
(±)-b-Pinene (26), a more reactive ene than 24, reacted with ethyl mesoxalate un-
der CCl4 reflux in 90 % yield after 5 h conventional or MW heating [42]. The reaction
supported on graphite occurred in only 2 × 1 min of MW irradiation with an incident
power of 60 W. The adduct 27 was obtained in 67 % isolated yield (Scheme 7.3) [30].
This yield was obtained after 2 h of reaction under CCl4 reflux.
The intramolecular cyclization of (+)-citronellal (28) leads to a mixture of isomeric
pulegols (29) and, particularly, to the (±)-isopulegol (29a) which is an intermediate in
the synthesis of (±)-menthol [43]. The reaction can be performed by heating in the
absence [44] or presence [43, 45] of a catalyst, including zeolite [46] and montmorillo-
nite [42] under the action of MW. The cyclization of 28, when performed by the GS/
MW process ± 4 min (8 × 30 s) of irradiation at 30 W incident power ± resulted in
88 % conversion (80 % isolated yield) to the pulegols (29) (Scheme 7.4).
Scheme 7.3
Scheme 7.4
7.2 Graphite as a Sensitizer 227
7.2.3
Oxidation of Propan-2-ol
Reaction of an alcohol over basic catalysts favors dehydrogenation to give the corre-
sponding carbonylated derivative. This reaction has been studied for propan-2-ol
over a series of alkaline carbon catalysts, under the action of conventional heating
and MW irradiation (Scheme 7.5) [17].
Scheme 7.5
7.2.4
Thermolysis of Esters
The thermolysis of esters is a much used reaction in organic [47] and organometallic
[48] syntheses, generally for the creation of a carbon±carbon double bond. The me-
228 7 Microwave-assisted Reactions on Graphite
Scheme 7.6
chanism is often like that of the retro-ene reaction, requiring high temperatures.
Among the esters, those of carbamic acid are more easily decomposed than those of
carboxylic acids.
The high temperatures reached in the GS/MW process have been used to achieve
the thermal decomposition of O-alkylcarbamates (Scheme 7.6) [15, 16]. The latter are
prepared from corresponding alcohols and phenyl isocyanate in the presence of stan-
nous octanoate [49].
With 1,1-dimethylbenzyl phenylcarbamate (30), a tertiary carbamate:
This example shows that the GS/MW process can be used to accomplish thermal
decomposition which cannot be performed efficiently by use of MW irradiation
alone, because of weak MW absorption by the starting compound (30) of probable
low dielectric loss.
The decomposition of a secondary carbamate, 1-methylbenzyl phenylcarbamate
(31), was more difficult, and only 60 % of styrene (33) was obtained under the same
conditions (Tmax = 340 8C). Attempts to decompose a primary carbamate, 1-octyl phe-
nylcarbamate, failed because its sublimation preceded its decomposition [15, 16]. All
these reactions have been performed in an open reactor by the above procedure
(Sect. 7.2.1).
Some esters not having an aliphatic hydrocarbon chain are liable to thermal rear-
rangement. This is observed for O-arylthiocarbamates, for which rearrangement into
S-arylthiocarbamates has been studied by Villemin et al. on different supports and
under the action of MW irradiation (Scheme 7.7) [18].
No rearrangement was observed for the pure compound 34a, adsorbed or not on
KF-Al2O3, probably owing to its low dielectric loss. By using supports known to con-
vert MW energy into thermal energy, the authors observed a conversion rate of 15 to
90 % for 34. The best yields (30 (35e) to 90 % (35c)) were obtained on graphite pow-
der.
7.2 Graphite as a Sensitizer 229
Scheme 7.7
7.2.5
Thermal Reactions in Heterocyclic Syntheses
Scheme 7.8
cently been studied under the action of MW irradiation by several authors [52], in-
cluding the GS/MW process for 4-substituted 7-aminocoumarins [53]. Synthesis of
aminocoumarin 38 by the Pechmann reaction involves heating a mixture of m-ami-
nophenol and dimethyloxalacetate at 130 8C (Scheme 7.9). Under such conditions,
however, the yield of the reaction is variable, and usually low (36 %). Use of graphite
as a support led to 38 in slightly better yield (44 %).
Scheme 7.9
7.2 Graphite as a Sensitizer 231
Addition of a solid acid catalyst such as Montmorillonite K10 increased the yield
significantly under the action of either thermal heating (64 %) or MW irradiation
(66 %) [52 c]. Under the latter conditions the reaction time was reduced. Comparable
results were obtained for the synthesis of aminocoumarins 39±42 (Tab. 7.3) [53].
Tab. 7.3 Preparation of 7-aminocoumarin derivatives 38±42 using GS,K10/D and GS,K10/MW
processes [53].
1 38 30 66 66 64
2 39 8 61 45 54
3 40 5 65 30 66
4 41 8 75 56 68
5 42 12 62 390 62
a
The irradiation was programmed to furnish a constant temperature of 130 8C
b
Reactions were performed at 130 8C (oil bath)
We believe that shortening of the reaction times for all GS,K10/MW experiments,
compared with those using GS,K10/D, are the result of ªhot spotsº in the graphite
powder (Sect. 7.4.2).
7.2.6
Decomplexation of Metal Complexes
Decomplexation of some metal complexes calls for drastic conditions. This is true
for (Z-arene)(Z-cyclopentadienyl)iron(II) hexafluorophosphates, [FeAr(Cp)][PF6] [54,
55]. Although their chemical decomplexation is known [55 a], the most widely used
method is pyrolytic sublimation at high temperatures (>200 8C) [55 b]. To evaluate
MW irradiation as the method of decomplexation of such iron complexes, Roberts et
al. performed the reaction in the presence of graphite [54]. They discovered that the
yield of the free ligand from the [Fe(Z-N-phenylcarbazole)(Z-Cp)][PF6] complex (43)
depended on the kind (flakes or powder) and amount of graphite, and on the irradia-
232 7 Microwave-assisted Reactions on Graphite
7.2.7
Redistribution Reactions between Tetraalkyl- or Tetraarylgermanes and Germanium
Tetrahalides
Synthesis of alkyl (or aryl) halogermanes (46) from a germanium tetrahalide (44) oc-
curs in two steps (Scheme 7.10) [56]. The most difficult to realize, and the least selec-
tive, is the second, i. e. the redistribution between alkyl (or aryl) and halide substitu-
ents of R4Ge and GeX4 compounds. Depending on the ratio of these two compounds
the reaction gives alkyl (or aryl) halogermanes, 46a, 46b, or 46c. It requires a catalyst,
the most frequently used being the corresponding aluminum halide, and its amount
must be relatively high (approx. 20 mol %) [56]. The experimental conditions are,
moreover, rather drastic ± heating in a sealed tube between 120 8C (arylated series)
and 200 8C (alkylated series) for several hours.
These redistribution reactions are possible at atmospheric pressure under the ac-
tion of MW; irradiation is performed for a few minutes in the presence of the same
catalysts [57]. These reactions with the less volatile germanium tetrabromide (44b)
(b.p. 184 8C) have also been performed by use of the GS/MW process, without added
catalyst (Tab. 7.4, entries 1 and 3) [15, 16]. In this instance, despite the use of weaker
incident power, the temperature reached 420 8C, very much higher than that ob-
tained under the action of MW irradiation of a reaction mixture containing AlBr3
(200 8C to 250 8C) (Tab. 7.4, entries 2 and 4). The presence of this catalyst consider-
ably favors redistribution towards the dibrominated products (46b) (84 % for
R = nBu, 85 % for R = Ph) relative to the monobrominated compounds (46a), which
are the major products of the GS/MW process (78 % and 43 % respectively). The tri-
brominated products (46c), the most difficult to prepare, have been obtained with a
rather high selectivity (73 to 80 %) by use of the catalytic process under the action of
MW [57]. In this reaction, therefore, the GS/MW process seems less effective than
the MW-assisted and AlX3-catalyzed process.
Scheme 7.10
7.2 Graphite as a Sensitizer 233
Tab. 7.4 Redistribution reactions between germanium tetrabromide (44b) and tetrabutyl- or tetra-
phenylgermanes (45) under the action of MW irradiation [15, 16].
1a n
Bu 5.20 5.75 90 W, 3 min e 17 83 (78/3/19) 80
2b n
Bu 5.20 6.10 1.0 300 W, 8 min f 100 (8/84/8) 87
3a Ph 5.75 5.20 90 W, 3 min e 9 91 (43/32/25) 85
4b Ph 5.70 5.20 1.0 210 W, 8 min f 3 97 (9/85/6) 96
a
Reaction performed by the GS/MW process with 6 g graphite as support
b
Ref. [57]
c
Products were analyzed by GC and 1H NMR after alkylation [57]
d
Recovered Ge products
e
Continuous irradiation
f
Sequential irradiation, 2 min × 4
7.2.8
Pyrolysis of Urea
The reaction usually used to produce cyanuric acid (48) is the thermolysis of
urea (47) between 180 8C and 300 8C (Scheme 7.11) [58]. The reaction occurs
with formation of ammonia, which itself can react with 48 to give secondary pro-
ducts. It is, therefore, necessary to eliminate NH3 and to operate with an open
reactor.
This reaction has been studied with classical and MW heating under homoge-
neous and heterogeneous conditions [59]. Tab. 7.5 summarizes the results. When
the reaction was conducted in the homogeneous phase at 200 8C (Tab. 7.5, entries
1±4), identical reaction rates and similar yields and selectivities were obtained for
both heating modes. Kinetic data for the first-order equation were similar:
Ea (MW) = 159 ± 3 kJ mol±1, Ea (D) = 160 ± 3 kJ mol±1 ; ln A (MW) = 35 ± 1, ln A
(D) = 34 ± 1. In contrast, in the presence of graphite (47/graphite = 4/1, w/w), im-
proved yield and selectivity were obtained under the action of MW irradiation com-
pared with conventional heating (Tab. 7.5, entries 5±8) at the same bulk tempera-
ture. The authors ascribed this phenomenon to localized superheating (ªhot spotsº)
on the graphite surface (Sect. 7.4.2).
Scheme 7.11
234 7 Microwave-assisted Reactions on Graphite
Tab. 7.5 MW-assisted thermolysis of urea (47) under solvent-free conditions (Scheme 7.11) [59].
Entry Conditions a,b Tmax (pC) Time (min) Yield (%) Selectivity (%) Reaction rate (103 s±1)
Homogeneous phase c
1 MW 200 1 5.2 20.2 8.8
2 D 200 1 4.5 30.3 8.7
3 MW 200 30 68.4 73.6 8.8
4 D 200 30 67.9 72.2 8.7
Heterogeneous phase d
5 MW 200 1 9.9 56.3 21.9
6 D 200 1 4.6 33.6 8.7
7 MW 300 3 61.2 93.5 12 028
8 D 300 3 15.2 45.6 7156
a
Microwave (MW) or conventional (D) heating
b
The incident MW power was adjusted to furnish the maximum temperature (Tmax)
c
From 20 g urea (mp 133±135 8C)
d
From 20 g urea + 5 g graphite.
7.2.9
Esterification of Stearic Acid by n-Butanol
Esterification of stearic acid (49) by butanols has been studied under homogeneous
and heterogeneous conditions [60]. The yield of butyl stearates [Me(CH2)14COOR
(50), R = nBu (a), sBu (b), tBu (c)] depended on the catalysts, on the isomeric form of
the butanol, and on the mode of heating (conventional heating and MW irradiation).
The esterification yield was similar under homogeneous conditions, irrespective of
the mode of heating. In contrast, under heterogeneous conditions (e. g. with iron(III)
sulfate or potassium fluoride as catalyst), after 2 h at the same bulk temperature
(140 8C), the MW-irradiated reaction resulted in a higher (1.2 to 1.4-fold) yield of 50a
than did conventionally heating. This difference was more evident when graphite
was added to the reaction mixture ± similar yields (75±95 % 50a) were obtained after
only 5 min MW irradiation. Despite the small amount of added graphite (about 1/
10, w/w, relative to the reagents), the reaction mixture rapidly reached a much higher
temperature (250 to 300 8C) than in the absence of graphite.
7.3
Graphite as Sensitizer and Catalyst
As a support for chemical reactions graphite has often been ªmodifiedº by addition of a
variety of substances which can be intercalated between the carbon layers (GIC ± gra-
phite intercalation compounds) or dispersed on the graphite surface, depending on the
preparation conditions [13]. The resulting compounds, especially the GIC, have been
used as reagents and as catalysts in numerous reactions, particularly in organic transfor-
mations [13, 61]. Depending on the intercalating guest the carbon lattice can behave as
7.3 Graphite as Sensitizer and Catalyst 235
an electron acceptor (e. g. with metals ± C8K) or as an electron donor (e. g. with halogens
± C8Br). The electron power does not, however, seem to give Lewis-type catalytic activity
to the graphite itself. As long ago as 1994 [15, 16] we reported that reactions known to re-
quire a Lewis acid catalyst can be conducted in the presence of unmodified graphite;
this has been confirmed more recently by other authors [62±65]. We showed that the cat-
alytic ability of graphite is a result of metal impurities, not the carbon lattice [66]. Most
reported graphite-catalyzed reactions have been performed under the action of conven-
tional heating, in the presence of a solvent and a small amount of graphite [62±64].
Use of graphite-supported methodology has been reported for three types of reac-
tion ± the Friedel±Crafts acylation [15, 16, 27, 66], the acylative cleavage of ethers [15,
16], and the ketodecarboxylation of carboxylic diacids [67, 68], either with conven-
tional heating (GS/D) or MW irradiation (GS/MW coupling); these are discussed be-
low. First, however, we describe the analysis of two commercial graphites of different
purity which are used for these experiments.
7.3.1
Analysis of Two Synthetic Commercial Graphites
. graphite A, Aldrich 28,286±3; purity 99.1%; particle size 1±2 mm; and
. graphite B, Fluka 50870; purity 99.9 %; similar granulometry.
7.3.2
Acylation of Aromatic Compounds
with the ketone produced instead of with the acylating agent [70] (Scheme 7.12).
Rarely does a metal chloride complex preferentially with the acylating agent, except
bismuth(III) chloride with acid chlorides [71, 72].
A stoichiometric amount of promoter, at least, is required for the reaction to pro-
ceed, leading to an environmentally hostile process with gaseous effluents and
mineral wastes. With some metal salts, however, an increase in reaction temperature
sets them free from their complex with the ketone, and a true catalytic reaction be-
comes possible [73]; this is observed for iron(III) chloride [74] and some metal tri-
flates [72, 75], including their use under the action of MW heating [76].
In 1994 our preliminary results revealed, surprisingly, that FC acylation could be
realized in the presence of graphite A, under solvent-free conditions, under the ac-
7.3 Graphite as Sensitizer and Catalyst 237
Scheme 7.12
tion of either classical heating or MW irradiation [15, 16]. More recently the same re-
action has been reported in the presence of a small amount of the same graphite
and using a solvent [62]. We have explained this ªcatalytic effectº of graphite [66].
The procedure entailed MW irradiation, at atmospheric pressure, of graphite pow-
der A impregnated with reagents. The first experiments were performed with an ac-
tivated aromatic, anisole (51), using a variety of acylating reagents (Tab. 7.6). With vo-
latile acid chlorides such as acetyl or isobutyryl the reaction occurred in convenient
yields (entries 1 and 2), and boiling was delayed as a result of graphite adsorption.
With nonvolatile benzoyl chloride, the conversion became quantitative (entry 4).
Good yields were also obtained for a long-chain derivative, which is more difficult to
desorb from graphite, and for a furan derivative, known to resinify in the presence
of a Lewis acid (entries 3 and 5).
Tab. 7.6. Graphite-supported acylation of anisole (51) by use of a variety of acylating reagents
(RCOX) under the action of MW irradiationa [27, 66].
a
Graphite A, 5 g; 51, 10 to 20 mmol
b
51/RCOX (mol) = 2/1 (entries 1, 2), 4/1 (entries 3±7)
c
Applied incident power; sequential irradiation time; period between two irradiations 2 min, except
entry 4, 1 min 40 s
d
Products: MeO(C6H4)COR, R = Me(58), iPr(59), Und(60), Ph(61), Fu(62)
e
Conversion determined by GC; yield of isolated product (in brackets) relative to the minor reagent
f
Und = undecyl; Fu = 2-furyl
g
p/o 6 95/5
h
p/o = 82/18
i
FeCl3/(RCO)2O = 1/10 (mol)
j
Continuous MW irradiation with degressive power
238 7 Microwave-assisted Reactions on Graphite
The method was also applied to the benzoylation of other aromatic compounds
(Tab. 7.7). The benzoylation of benzene itself, volatile and less reactive, seemed more
difficult to perform (Tab. 7.7, entry 4). Silyl-substituted aromatics reacted by ipso Si-
substitution [77], and were less volatile. With trimethylsilylbenzene, benzoylation oc-
curred with an overall yield higher than for benzene, but the competitive H-substitu-
tion was also observed (entry 5).
If graphite A has catalytic activity in these reactions, the amount of graphite could
be reduced, and 0.5 g (instead of 5 g) of graphite A was, indeed, sufficient to promote
these reactions (Tab. 7.8) [66]. In the process in which a small amount of graphite
was used:
. the temperature gradient was lower than for the GS/MW process; and
. vaporization of the reactants was not delayed.
Consequently:
. sequential MW irradiation was preferable to continuous (compare Tab. 7.8, entries
3 and 5); and
. the MW power must be reduced and the reaction time increased (compare
Tab. 7.6, entry 4, and Tab. 7.8, entry 5).
Tab. 7.8 Acylation of aromatic compounds in the presence of a small amount of graphite and under the action of
MW irradiation [27, 66].
Entry ArH a RCOX b Graphite c Conditions d, (Tmax) Product a Conversion and yield (%) e
able (naphthalene), the yield obtained under the action of MW was higher. SMWI
enables control and limitation of these phenomena.
With regard to a mechanistic hypothesis, the catalytic effect of graphite itself be-
having as a Lewis acid was excluded ± the use of graphite B instead of graphite A re-
sulted in no reaction, or sometimes only a trace of the acylation product [27, 66]. The
presence of relatively large amounts of Fe3O4 in graphite A (Sect. 7.3.1) led us to be-
lieve this impurity was responsible for the catalytic activity observed. Several further
experiments supported this.
Graphites combined with a variety of metals (ªGraphimetsº) are known for their
catalytic properties [13]. The iron±graphite compound in which the presence of
Fe3O4 crystallites has been shown [78] proved very efficient for the acylation of ani-
sole (Tab. 7.8, entry 4). Because the iron content (5 %) was much higher than that of
graphite A, this graphimet could be a convenient material for GS/MW experiments,
but its cost, especially relative to that of graphite A, limits its use.
For the benzoylation of anisole (Tab. 7.6, entry 4), graphite A was replaced by gra-
phite B doped with Fe3O4 (28 mg for 5 g graphite, the same iron content as in gra-
phite A). Although unachievable with graphite B alone, benzoylation of anisole was
now possible, but with lower yield than for use of graphite A (50 % instead of 100 %).
This showed the activity of Fe3O4 became much stronger when subjected to the high
graphitization temperature.
The catalytic effect of graphite A thus depends on iron impurities, e. g. Fe3O4, and
probably also on iron sulfides or sulfates, because sulfur is also present in this gra-
phite, and all these iron compounds are known catalysts of FC acylation [69, 73, 74]. In
this respect, it seems that FeCl3 could be the true catalyst generated in situ by the reac-
tion of the different iron compounds with acid chloride and hydrogen chloride. In the
240 7 Microwave-assisted Reactions on Graphite
absence of a chlorinating agent, for example using an acid anhydride as the reagent
and an iron oxide (Fe2O3 or Fe3O4) as the catalyst, acylation does not occur. We have ef-
fectively shown that the GS/MW process using acid anhydrides as reagents is efficient
only after addition of a catalytic amount of FeCl3 (Tab. 7.6, entries 6 and 7).
Finally, a sample of graphite A was analyzed after acylation using an acid chloride
as reagent. No Fe3O4 crystallites were observed and an EDX spectrum revealed small
deposits containing iron, chlorine, and oxygen. Thus formation of FeCl3 from Fe3O4
crystallites is highly probable. Loupy et al. have shown that a-Fe2O3 can be generated
from Fe3O4 under the action of MW at high temperature [79]; the formation of FeCl3
would be a result of chlorination of Fe3O4 and/or Fe2O3. Because Fe3O4 interacts
strongly with MW [4], the presence of hot spots in the region of Fe3O4 crystallites
could also lead to increased catalytic activity.
Comparison with previous FC acylations, the above processes are clean, without
aqueous workup, and therefore without effluents (ªgreen chemistryº). The graphite
is, moreover, inexpensive and can be safety stored or discarded. Its activity is, how-
ever, limited to activated aromatic compounds.
The process which seems to have the most possibilities for a scale-up development
is that using a low amount of graphite, for which the desorption treatment can be to-
tally suppressed in a continuous flow system. We recently proposed the use of such
a process to perform FC acylations under the action of MW with FeCl3 as catalyst
[76 d]. The replacement of FeCl3 by a graphite bed is quite conceivable in the same
continuous flow apparatus.
7.3.3
Acylative Cleavage of Ethers
Scheme 7.13
Concerning the mechanism of such a reaction and the nature of the catalyst, we
do not think that the graphite itself is the catalyst. In fact, diethyl- and di-n-butyl
ethers are inert towards benzoyl chloride in the presence of graphite B. Moreover, it
is known that metal chlorides, especially FeCl3 [82], are catalysts for this reaction.
After the careful mechanistic study realized in the case of FC acylations (Sects. 7.3.1
and 7.3.2), we propose that the catalyst of the graphite-assisted acylative cleavage of
ethers is FeCl3 generated in situ from Fe3O4 (and/or Fe2O3) and the acid chloride.
Then, the C±O bond cleavage would involve the O-acylation of ether ([R2O±COR']+)
followed by the nucleophilic displacement (SN1 or SN2) of one of the two hydrocar-
bon groups (R) of ether by the anionic part of the reagent (Cl±), as with the FeCl3±
Ac2O system [82]. It is interesting to note here that phenolic ethers, such as anisole
or veratrole, preferentially give the acylation of aromatic nuclei towards the cleavage
of the ether group (Sect. 7.3.2).
7.3.4
Ketodecarboxylation of Carboxylic Diacids
avoid vaporization of the diacid. For adipic acid, the limiting temperature is about
290±295 8C [83, 86, 88]. Thanks to its properties of rapid conversion of MW energy
and retention of organic molecules, graphite could allow a high reaction temperature
to be reached rapidly, although it delays the vaporization of the diacid. Moreover,
since magnetite (Fe3O4) is a catalyst for the decarboxylation of acids [91], no added
catalyst would be necessary.
To determine approximately the optimum temperature of a graphite-supported cy-
clization of adipic acid, a series of experiments was performed with classical heating.
Using the two graphites A and B (Sect. 7.3.1), no significant vaporization of adipic
acid (70) was observed up to 450 8C at atmospheric pressure. Graphite A proved to be
the more efficient, giving 90 % yield of cyclopentanone (74) after 30 min of heating
(Tab. 7.9, entry 1). Graphite B gave a lower yield under the same conditions (entry 2).
1 70 A 74, 90 (85)
2 70 B 74, 22
3 71 A 75, 60
4 71 A 75, 80 (74)
5 72 A 76, 80 (72)
6 73 A 77, 17
a
Temperature of the electrical oven: 450 8C; reaction time: 30 min (entries 1±4), 60 min (entries 5, 6);
pressure: atm. p. (entries 1±3), 300 mm Hg (entries 4±6)
b
Optimized diacid/graphite ratio: 15 mmol/5 g
c
Conversion determined by GC and isolated yield (in brackets) from four experiments
The retentive power of graphite towards adipic acid and the catalytic effect of the
magnetite, especially present in A, are obvious. TEM examinations of a graphite A
sample before and after reaction showed that crystallites of Fe3O4 appeared to be
smaller after the reaction. However, the same graphite sample was reused for three
successive reactions without significant loss in yield. When applied to the synthesis
of other cyclic ketones (Scheme 7.14), less volatile than 74, it was observed that pres-
sure had an effect on the recovery of product (Tab. 7.9, entries 3 and 4). A slightly re-
duced pressure (300 mm Hg) was necessary to obtain 3-methylcyclopentanone (75)
or cyclohexanone (76) in convenient yield (Tab. 7.9, entries 4 and 5). For the cycliza-
tion of suberic acid (73), a less favorable structure, the yield in cycloheptanone (77)
remained low (Tab. 7.9, entry 6).
Some MW-promoted decarboxylations have been reported in the literature [92],
even the decarboxylation of magnesium, calcium and barium salts of alkanoic acids
[92 a]. We have shown for FC acylations and acylative cleavage of ethers (Sects. 7.3.2
and 7.3.3) that the graphite-supported process takes advantage of MW, since graphite
and magnetite are among the solids having the most efficient MW absorbing power
[4], and then providing elevated temperature in core. Consequently, the cyclization of
70 was realized under GS/MW conditions. A SMWI mode was optimized and
7.3 Graphite as Sensitizer and Catalyst 243
Scheme 7.14
coupled with a limitation of the reaction temperature to 450 8C, using the two gra-
phites A and B (Tab. 7.10, entries 1±4). Graphite A showed again its superiority, giv-
ing under these optimized conditions (entry 2) a 90 % yield in cyclopentanone (74)
after only 6 × 2 min of irradiation. Under the same conditions, graphite B gave only
a 33 % yield (entry 4).
Tab. 7.10 Graphite-supported ketodecarboxylation of adipic acid (70) under MW irradiationa [67, 68]
1 A; none 90 W; 2 min × 2 60
2 A; none 90 W; 2 min × 2 + 75 W; 2 min × 4 90
3 B; none 90 W; 2 min × 2 19
4 B; none 90 W; 2 min × 2 + 75 W; 2 min × 4 33
5 B; Fe3O4 (28 mg) 90 W; 2 min × 2 51
6 B; Fe2O3 (29 mg) 90 W; 2 min × 2 51
7 B; FeO (26 mg) 90 W; 2 min × 2 35
8 B; Al2O3 (28 mg) 90 W; 2 min × 2 15
9 B; Bi2O3 (28 mg) 90 W; 2 min × 2 16
10 B; KF (21 mg) 90 W; 2 min × 2 14
11 B; Na2CO3 (80 mg) 90 W; 2 min × 2 29
12 B; Cs2CO3 (244 mg) 90 W; 2 min × 2 26
13 Fe3O4 (3.47 g) without graphite 90 W; 2 min × 2 10
a
Mass of 70: 2.19 g (15 mmol); mass of graphite: 5 g
b
Sequential MW irradiation controlled to a maximum temperature of 450 8C
c
Applied incident power and irradiation time; interval between two irradiations: 2 min
d
Yield of cyclopentanone (74) from GC analysis
To compare their activities further, various catalysts were added to graphite B, and
the results were analyzed with respect to the reference experiment (Tab. 7.10, en-
try 3) which gave a low yield (19 %). When doped with Fe3O4, graphite B gave a 51%
conversion of 70 (entry 5), almost as much as graphite A alone (entry 1). The two
244 7 Microwave-assisted Reactions on Graphite
other iron oxides, Fe2O3 and FeO, seemed to be active but other catalysts, Al2O3,
Bi2O3 and KF were inactive.
Because Fe3O4 itself is strongly absorbent of MW [4] and is a good catalyst for dec-
arboxylations [91], is the graphite necessary? When Fe3O4 was used in the absence of
graphite, the yield in ketone 74 decreased dramatically (10 %) (Tab. 7.10, entry 13).
Adipic acid was recovered almost completely on the walls of the reactor and on the
cold finger. Consequently, the presence of graphite as a support, able to adsorb and
retain the adipic acid, and then allow the cyclization before its vaporization, is essen-
tial.
Comparison of reaction times revealed a shortening under MW irradiation
(Tab. 7.10, entry 2: overall reaction time = 22 min) with respect to conventional heat-
ing (Tab. 7.9, entry 1, 30 min), for the same maximum temperature. This can be ex-
plained by a higher temperature gradient and the presence of ªhot spotsº at the gra-
phite surface under MW.
A reaction mechanism with Fe3O4 as catalyst has been proposed [68], in agree-
ment with previous work concerning decarboxylation of acids in the presence of a
metal oxide [83]. After the transient formation of iron(II) and iron(III) carboxylates
from the diacid and Fe3O4 (with elimination of water), the thermal decarboxylation
of these salts should give the cyclic ketone and regeneration of the catalyst.
This ªdryº, solvent-free GS/MW process, which allows high temperatures rapidly,
is very useful for the ketodecarboxylation of diacids. This has several advantages
such as:
7.4
Notes
7.4.1
MW Apparatus, Typical Procedures, and Safety Measures
Graphite reflects MW like a metal and its heating depends strongly on particle size.
With large particles (flakes), electric discharges were observed under MW irradia-
tion, whereas with fine ones (powder) a rapid increase of temperature was obtained.
The temperature of the reaction mixture must be controlled to avoid melting point
of the reactor, if it is Pyrex glass. The use of a quartz reactor is highly preferable.
A MW applicator, such as the Synthewave 402 from Prolabo or nowadays Discover
from CEM, equipped with:
7.4 Notes 245
. a monomode cavity;
. open vessels allowing reactions at normal or reduced pressure;
. a rotating reactor;
. a MW power modulator;
. an IR pyrometer allowing a continuous record of reaction mixture temperature;
. a computer to monitor the different irradiation modes and the main experimental
parameters (incident MW power, temperature and time),
7.4.2
Temperature Measurement
7.4.3
The Retention Mechanism of Reactants on Graphite
7.4.4
Graphite or Amorphous Carbon for C/MW Coupling?
7.5
Conclusion
Its inert behavior towards numerous chemical compounds and its adsorbent prop-
erties (responsible for the retention of volatile or sublimable organic compounds),
make graphite the choice support for thermal reactions. Among its impurities, mag-
netite was revealed to be an active catalyst, and some reactions can be performed
without any added catalyst. Two processes are then possible, the graphite-supported
reaction (ªdryº process), and the reaction in the presence of a small amount of gra-
phite (solid±liquid medium).
In a broad perspective, large scale development seems possible for some reactions,
either by a continuous-flow process through a MW-irradiated graphite plate (solid±li-
quid (or gas) reactions), or by a continuous supply of solid starting compound on a
graphite bed (ªdryº reactions). In this respect, continuous processes have recently
given excellent results in large scale development of MW-assisted reactions in homo-
geneous and heterogeneous media [76 d, 99].
Acknowledgments
References
1 C. R. Strauss, Microwave-assisted Or- science, New York, 1978, p. 556, and re-
ganic Chemistry in Pressurized Reactors, ferences cited therein.
this book, Chapt. 2. 12 a) D. P. E. Smith, J. K. H. Hærber,
2 a) A. Loupy, A. Haudrechy, Effets de G. Binnig, H. Nejoh, Nature 1990,
milieu en synth se organique. Masson, 344, 641; b) J. P. Rabe, S. Buchholz,
Paris, 1996, p. 277; b) A. Loupy, Science 1991, 253, 424; c) J. P. Rabe,
A. Haudrechy in: Mthodes et tech- Ultramicroscopy 1992, 42, and refer-
niques de la chimie organique. D. Astruc, ences cited therein. For analysis of vola-
(ed.), Presses universitaires, Grenoble, tile organic compounds by trapping in
France, 1999, Chapt. 6, p. 239; a carbonaceous adsorbent and by ther-
c) A.Loupy, A. Petit, J. Hamelin, mal desorption using MW, see
F. Texier-Boulet, P. Jacquault, d) M. Almarcha, J. Rovira, Tec. Lab.
D. Math, Synthesis 1998, 1213; 1991, 13, 322. For adsorption of two or-
d) R. S. Varma, Microwave and Reactions ganic solvents, methyl isobutyl ketone
using Supported Reagents, this book, and methyl isobutyl carbinol, on to gra-
Chapt. 6, and references cited therein. phite as a function of evaporation tem-
3 D. S. Cameron, S. J. Cooper, I. L. Dodg- perature, see e) D. S. Martin,
son, B. Harrison, J. W. Jenkins, Catal. P. Weightman, Surf. Sci. 1999, 441,
Today 1990, 7, 113. 549.
4 a) J. W. Walkiewicz, G. Kazonich, 13 a) M. A. M. Boersma, Catal. Rev. Sci.
S. L. McGill, Miner. Metal. Proc. 1988, Eng. 1974, 10, 243; b) H. B. Kagan,
5, 39; b) D. M. P. Mingos, D. R. Ba- Chemtech 1976, 510; c) H. B. Kagan,
ghurst, Chem. Soc. Rev. 1991, 20, 1; Pure Appl. Chem. 1976, 46, 177;
c) D. M. P. Mingos, Chem. Ind. 1994, d) H. Selig, L. B. Ebert, Adv. Inorg.
596. Chem. Radiochem. 1980, 23, 281;
5 K. J. Rao, B. Vaidhyanathan, M. Gan- e) R. Setton in: Preparative Chemistry
guli, P. A. Ramakrishnan, Chem. Ma- Using Support Reagents. P. Laszlo, (ed.),
ter. 1999, 11, 882, and references cited Academic Press, London, 1987,
therein. Chapt. 15, 255; f) R. Czuk, B. I. Glån-
6 J. K. S. Wan, T. A. Koch, Res. Chem. In- zer, A. Fçrstner, Adv. Organometal.
termed. 1994, 20, 29. Chem. 1988, 28, 85; g) P. Cintas, Acti-
7 J. K. S. Wan, Res. Chem. Intermed. 1993, vated Metals in Organic Synthesis.
19, 147. C. W. Rees, (ed.), CRC Press, Boca Ra-
8 M. Y. Tse, M. C. Depew, J. K. S. Wan, ton, 1993, p. 70; h) A. Fçrstner, Angew.
Res. Chem. Intermed. 1990, 13, 221. Chem. Int. Ed. Engl. 1993, 32, 164, and
9 a) D. A. Lewis, Mater. Res. Soc. Symp. references cited therein.
Proc. 1992, 269 (Microwave Processing 14 D. Avnir, D. Farin, P. Pfeifer,
of Materials III), 21; b) M. C. Hawley, J. Chem. Phys. 1983, 79, 3566.
J. Wei,V. Adegbite, Mater. Res. Soc. 15 R. Laurent, Th se, Universit Paul±Sa-
Symp. Proc. 1994, 347 (Microwave Pro- batier, Toulouse, 1994.
cessing of Materials IV), 669, and refer- 16 M. Audhuy-Peaudecerf, J. Berlan,
ences cited therein. J. Dubac, A. Laporterie, R. Laurent,
10 M. S. Ioffe, E. A. Grigoryan, Neftekhi- S. Lefeuvre, French Patent, 1994,
miya 1993, 33, 557; Chem. Abstr. 1994, no. 94.09073 (FR. Appl. 20 July 1994).
120, 109907y; b) D. D. Tanner, 17 G. Bond, R. B. Moyes, I. Theaker,
Q. Ding, P. Kandanarachchi, D. A. Whan, Top. Catal. 1994, 1, 177.
J. A. Franz, Prepr. Symp.-Am. Chem. 18 a) D. Villemin, M. Hachemi, M. La-
Soc., Div. Fuel Chem. 1999, 44, 133; laoui, Synth. Commun. 1996, 26, 2461;
Chem. Abstr. 1999, 130, 324766p. b) A. Ben Alloum, Th se, Universit de
11 Kirk±Othmer, Encyclopedia of Chemical Caen, 1991.
Technology, 3rd edn.,Vol. 4. Wiley±Inter- 19 A. de la Hoz, A. Diaz-Ortiz, F. Langa,
References 249
8
Microwaves in Heterocyclic Chemistry
Jack Hamelin, Jean-Pierre Bazureau and Franoise Texier-Boullet
8.1
Introduction
This chapter will deal with applications of microwave irradiation in the synthesis of
heterocycles by a variety of means, excluding cycloadditions, which will be described
in the next chapter. We have chosen to report first reactions in solution in organic
solvents, then heterogeneous reactions without solvent under a variety of conditions,
and finally to deal with emerging techniques which employ ionic liquids.
8.2
Microwave-assisted Reactions in Organic Solvents
8.2.1
Heck, Suzuki, and Stille reactions
The palladium-catalyzed Heck [1], Suzuki [2] and Stille [3] reactions are robust and
general methods for C±C bond-formation and have therefore emerged as important
reactions in the synthesis of natural heterocyclic compounds.
Hallberg and coworkers [4 a] have used a single-mode microwave oven [4 b] at
50 W to improve radical-mediated reduction and cyclization of the heterohalide 1
with HSn(CH2CH2C10F21)3 2 in benzotrifluoride (BTF). The authors claim the pre-
paration of 3 in high yield (93 %) after 5 min, as illustrated in Scheme 8.1.
The efficiency of fluorous Stille coupling reactions [5 a] is enhanced by use of mi-
crowave irradiation (Scheme 8.2). The reaction proceeds in 79 % yield after 2 min
with DMF as the microwave-active solvent.
I
AIBN, BTF
HSn(CH2CH2C10F21)3
N 5 min./50W N
Cbz Cbz
1 2 3 (93%)
Scheme 8.1
The same authors performed a microwave assisted Stille reaction on the Rink
amide (RAM) Tentagel polymer-tethered 4-iodobenzoic acid [5 b]. Successful palla-
dium-catalyzed coupling of heteroaryl boronic acid with anchored 4-iodobenzoic acid
enabled both >99 % conversion of the starting material within 3.8 min (45 W) and a
minimal decomposition of the solid support. The coupling reactions were realized
in a mixture of polar solvents (H2O±EtOH±DME, 2.5 : 1.5:6).
A recent report [6] has discussed the effect of monomode microwave irradiation in
the palladium-catalyzed phenylation of 5-iodouracil 4 with the nontoxic sodium tet-
raphenylborate 5 as phenyl reagent (Scheme 8.3). The authors showed that the use
of monomethylformamide (MMF) as solvent increases the yield of 6 (70 %), because
MMF has a high boiling point (180 8C) and is more polar (e = 182.5) than other
amides used in microwave-activated reactions.
A regiocontrolled Heck vinylation [7] of commercially available 2-hydroxyethyl vi-
nyl ether 7 in dry DMSO has been reported (Scheme 8.4). Flash heating by micro-
O O
I Pd(OAc)2
HN HN
Ph4BNa
O N MMF O N
H H
8 min./10W
4 5 6 (70%)
Scheme 8.3
OTf O O
Pd(OAc)2
O
OH
MeO DPPP, Et3N
MeO
DMSO
7 8 9 89 % (5 min./5W)
69 % (20 h./40°C)
Scheme 8.4
8.2 Microwave-assisted Reactions in Organic Solvents 255
8.2.2
Aziridine Synthesis
O Br CuBr 2
Me S N N Ts
O Na MeCN
81 % (MW, 12 min.)
Scheme 8.5
8.2.3
b-Lactam Chemistry
Bose and coworkers [10] have described hydrogenation using ammonium formate as
hydrogen donor and Pd/C as catalyst for selective transformations (Tab. 8.1) of b-lac-
tams 10, as shown in Scheme 8.6.
Y Ph HCO2NH 4 , 10 % Pd/C, MW Z Ph
H
N N
O R HO O R
OH
Ph Ph Ph
O Et3N
N N N
Cl 118°C, MW O Me O Me
Me
12 13 10 min. , 90% 14 cis (20) 14 trans (80)
Scheme 8.7
OTBDMS
H H
OTBDMS S NH-Boc
H H N
S conditions, O
NH-Boc see Table 8.2 CO2PNB
O MeP(OEt)3 18
N
O CO
PNB OTBDMS
16 O O H H S
17 NH-Boc
N O
O
CO2PNB 19
Scheme 8.8
The authors report reaction times similar to those achieved with a preheated oil
bath at 130 8C on a small scale; on a larger scale, however, microwave-assisted reac-
tions seem to proceed more rapidly.
The same authors also studied recently the preparation of substituted vinyl b-lac-
tams 14, with efficient stereocontrol [11], by use of limited amounts of solvent (chlor-
obenzene) (Scheme 8.7). Microwave oven-induced reaction enhancement (MORE)
chemistry techniques have been used to reduce pollution at the source and to in-
crease atom economy.
A comparative study [12] of the reactivity of the oxalimide 16 in a variety of sol-
vents (xylene, chlorobenzene, toluene) and of methylphosphinite 17 was performed
with the focused microwave Synthewave 402 reactor (Merck Eurolab, div. Prolabo,
France), using different conditions of power and exposure time (Scheme 8.8). In all
experiments yields were better than those of previous procedures with classical heat-
ing (Tab. 8.2), and the authors wrote ªit is clear that microwave technique is applic-
able to highly functionalized compounds containing stereogenic centers without ap-
preciable modification of these centersº.
Solvent Time (min) Power (W) Final temp. (pC) Yield of 18 (%)
8.2.4
1,2,4-Triazole, Pyrazole Synthesis
HO NH 2
OH N N N N
N
N CN H2N-NH 2, HCl N N N N
H2N-NH 2, H2O N N
H H
20 21 22
MW oil bath
95% in 5 min. 85% in 45 min.
Scheme 8.9 a
H2N-NH 2, H2O N N
68%, 10 min. in DMF
77%, 10 min. in EtOH
N CN S, N2 N N N N
solvent H H
125W
20 21
Scheme 8.9 b
R O
H MW R = phenoxy, octyl
N Ar = 2-hydroxyphenyl, 2-hydroxynaphthyl
R N Ar N H
HCO 2H Ar 3-nitrophenyl, phenyl, 4-chlorophenyl
O N
4-7 min. 4-methoxyphenyl
H
75-86%
Scheme 8.10
258 8 Microwaves in Heterocyclic Chemistry
at 100±120 8C whereas the same reaction is complete in 4±7 min with improved yield
when performed under the action of microwave irradiation.
8.2.5
Multistep Synthesis of Polyheterocyclic Systems
OEt
NC NO 2
i S S NC NO 2 ii NO 2 iii
N N
H2N 78% N MW MW
Cl NC N
22 23 24
25 26 27
CN
OEt S O S
vi N vii H N
N N
MW MW
NC N N
28 29
Scheme 8.11 Reagents and conditions (for reflux; (iii) SnCl2.2H20, EtOH, 70 8C; (iv) Br2,
time and yields of steps ii, iii, vi, and vii see AcOH, rt, 24 h; (v) 4,5-dichloro-1,2,3-dithiazo-
Tab. 8.3). (i) 4,5-dichloro-1,2,3-dithiazolium lium chloride, rt, 4 h; (vi) CuCN, pyridine, reflux;
chloride, pyridine, rt, 10 h; (ii) NaH, EtOH, (vii) HCl, reflux.
8.2 Microwave-assisted Reactions in Organic Solvents 259
Tab. 8.3 Comparison of conventional heating and microwave irradiation for steps (ii), (iii), (vi),
and (vii) of Scheme 8.11.
ii 23 24 640 37 80 61
iii 24 25 60 72 10 94
vi 27 28 90 50 20 53
vii 28 29 60 49 10 50
was easily accomplished at 60 8C (in 55 min), under the action of focused microwave
irradiation, by reaction of the starting 2-(2-aminophenyl)benzimidazole 30 with CS2
in the presence of KOH and a protic solvent (MeOH). Under the same conditions
conventional heating led to the same yield but with a reaction time of 24 h.
Alkylation of 31 with dibromomethane and 1,2-dibromoethane was performed un-
der solvent-PTC conditions with good yields and short irradiation times (15 min)
[16]. The synthesis of original benzimidazo-[1,2-c]-quinazoline dimers 32(a,b) was
successfully achieved by use of potassium carbonate in the microwave active DMF
solvent (Scheme 8.13).
N N ii N
i
N MWI N 95% N
H H2N 95% N N
HS MeS
31 32
30
N
N
iii, MWI 33a : R = H, (95%), 1.5 h. under MWI
N
N 33b : R = Me (75%), 2 h. under MWI
R NH2 O
CO2H
33
R
Scheme 8.12 Reagents and conditions: (i) CS2, MeOH, reflux,
MW, 55 min; (ii) MeI, NaH, DMF, 25 8C, 5 min; (iii) graphite, MW.
N
N
N Br-(CH2)n-Br S
32a : n = 1 (86%)
n 32b : n = 2 (81%)
N K2CO3 S
N TBAB, DMF N
HS 60°C, MW, 15 min. N
31
N
Scheme 8.13
260 8 Microwaves in Heterocyclic Chemistry
H NH2 N2
N O H i ii
N N N
O R X X
X
O R OR
O
33 34 35 36
Me
N3 N
iii iv X = H, Cl, Br
N R = Me, allyl
X X N
OR O
R
37 38
H
NH 2 NH 2
KOH, H 2O COCl 2 N O
CO2Me CO2K O
S MW, 30 min. S S
O
39 40 41
O
CO2Me CO2K
KOH, H2O COCl2 O
NH 2 MW, 15 min. S NH 2
S S N O
H
42 43 44
Scheme 8.15
Me
H N
O 1
R N Cl R1 N N
R1 N i ii
2
R R2
R2 N HN N Me
N N
O H H
45 46 47
Scheme 8.16 Reagents and reaction conditions: (i) POCl3, pyridine,
reflux, MW (700 W), 1 h 45 min; (ii) DMF, 240 8C, 6 bar; MW, 1 h.
8.2.6
Claisen Rearrangement
R3
R3
NMF, H2O
HO O O
R1 O O O MW, 10 min.
R2 R1
R2
48 R3
BF3/Et2O, NMF
R1, R2 = H, Me O O
MW, 10 min. O
R3 = H, Me, Ph
R2 R1 49
Scheme 8.17
formamide (NMF) [22]; the short reaction time (10 min) associated with the micro-
wave-assisted procedure is the best means of preparation of these compounds, as
shown in Scheme 8.17.
8.2.7
Hantzsch Cyclocondensations
Cl
O O O NH3, H2O
RO2C CO2R 55%, 4 min., MWI
EtOH, reflux, MWI 44%, 24 h, oil bath
Cl OR
Me N Me
H
50 51 52
Scheme 8.18
8.2 Microwave-assisted Reactions in Organic Solvents 263
R
Aq. hydrotope,
O CO2R1 MW
R R1O2C CO2R1
CO2R1 MeO2C Me
Me N Me
NH 2 H
50 53 52
Scheme 8.19
microwave cavity in four cycles of 6 min each; a 2-min gap between each cycle was
imposed to avoid excessive heating.
It should be noted that the Hantzsch-1,4-DHP synthesis can be conducted with a
soluble polymer. Successful derivatization of poly(styrene±co-allyl alcohol) under the
action of microwave irradiation with a variety of ethyl oxopropanoates and ethyl
3-aminobut-2-enoates were reported by Vanden Eynde's group [27].
Khmelnitsky and coworkers have also examined microwave-assisted parallel
Hantzsch pyridine synthesis [28]. They have demonstrated the benefits of microwave
irradiation in a 96-well plate reactor for high throughput, automated production of a
pyridine combinatorial library (Scheme 8.20).
In each reaction ethyl acetoacetate 51 a was used as one of the components of the
Hantzsch synthesis, whereas the second 1,3-dicarbonyl compound 51 (or 53) and the
aldehydes 50 were used in all possible combinations (one unique combination per
O O O O MW
R2 O
O RO R1 NH 4NO 3 / bentonite
DMF
51a 51 50
O R2 O O R2 O O R2 O
O O RO OR RO OR
N R1 N R1 R1 N
α
52α β
52β 52γγ
O
O O MW
R3 R2 O
O NH 4NO 3 / bentonite
O
DMF
51a 53 50
O O R2 O
R2 O O R2 O
O O O R3 R3
R3
N N
N
α
52α α
54α β
54β
Scheme 8.20
264 8 Microwaves in Heterocyclic Chemistry
well). Each well plate containing ammonium nitrate (on bentonite) was impregnated
with N,N-dimethylformamide as energy-transfer medium. The 96-well reactor
placed in a microwave oven (1300 W) was irradiated for 5 min at 70 % power level.
8.2.8
1,3,4-Oxadiazole Synthesis
O O THF
R1 O R2
R1 R2 O N NEt 3
PEG S
N N 2-4 min.,100W N N
H H O O O
55 56 (75-98%)
R1 = furyl, 3-pyridyl, 4-pyridyl, PhSO 2 CH 2
R2 = Ph, Me, NHPh
Scheme 8.21
8.2.9
Biginelli Cyclocondensation
MW R2
O H (30% of 360W) O H
N NH 2 N
R2 O R
O O X EtOH, HCl cat. Me N X
8-11 min.
R
51a 50 57 58 (65-80%)
,
R2 R F
F3C F
Scheme 8.22
8.2.10
Condensation Reactions with Creatinine and Thiohydantoin
Me
N NH
NH O Me
O N N
H 60 N H
R
H3BO3, DMSO O O
O MW (270W) 61
O SH
R AcOK, (AcO) 2O O
N
O MW (270W)
N H
59 R
H O
O
R = H, Me, Cl, Br, AcO, NO 2 N 63
S
O N
H 62
Scheme 8.23
266 8 Microwaves in Heterocyclic Chemistry
the presence of potassium acetate and acetic anhydride as solvent, using both meth-
ods. Here, the effect of microwave irradiation was shortening of the reaction times
and a small increase in the yields.
Tab. 8.4 Condensation of 3-formyl chromone 59 with creatinine 60 and thiohydantoin 62.
61a H 60 71 3 75
61b Cl 60 72 2 76
63a H 60 72 8 79
63b Br 60 88 9 96
8.2.11
Deprotection of N-Cbz and N-Bn derivatives; Deuterium-labeled Aromatic
Dehalogenation
Ph Ph
O O
i
CbzHN OMe H2N OMe
MW (600W) O
O 3 cycles OH
OH
64 65 (95%)
OH OH
Ph Ph
i
O NBz 2 O NH2
MW* (30W)
OCH2Ph 75°C, 20 min. OCH2Ph
66 67 (70%)
H O H O
N MW 20% of 750W N
N N
20 sec.
DCO2K, H2O
R Pd(AcO) 2 cat. D
68 (R = Br, I) DMSO 69 (95%)
Scheme 8.25
This procedure was also compatible with the benzyl ester, as demonstrated in
Scheme 8.24 by selective deprotection of compound 66 (67 was isolated in 70 % yield
by chromatography after 20 min irradiation at 75 8C (30 W) in the Synthewave 402
reactor).
It is interesting to remark that classical hydrogenation is a method used to prepare
deuterium-labeled compounds by aromatic dehalogenation, but the usual reaction
conditions suffer from several limitations. A microwave-enhanced catalytic dehalo-
genation procedure for rapid and specific deuterium labeling of N-4-picolyl-4-
halogenobenzamide 68, by use of deuterated formate, [41] was recently reported
(Scheme 8.25).
8.3
Solvent-free Synthesis
Although the initial reason for the development of solvent-free conditions for micro-
wave irradiation was safety, it soon became apparent that use of these conditions had
many other benefits ± simplicity, efficiency, easy work-up, very often higher yields,
and enhanced reaction rates. The absence of solvent is, furthermore, time- and
moneysaving and often enables elimination of waste treatment.
Solvent-free synthesis can be realized under a variety of conditions; for each we
give selected results and, when available, results from comparison with the same sol-
vent-free conditions but with classical heating.
8.3.1
Solvent-free Synthesis under Acidic Conditions
This section covers synthesis catalyzed either by organic acids or by inorganic acidic
supports.
Scheme 8.26
H O
Silica gel N N
+ H
"dry media" H
N N
N-H
method are its rapidity and more straightforward workup; yields suffer from the
same limitations as under classical conditions. It has been shown later that the yield
increases to 31% when the irradiation was performed in the presence of toluene [43].
H2C OH H2C O
CH R
HC OH HC O
O R-CHO O
HO O HO O
HO HO
Example : R = CH3-(CH2)11
DMF, H2SO4, CuSO4 24 h 40°C 25%
KSF "dry media" 10 min. 60W 66%
K10 "dry media" 10 min. 60W 89% Scheme 8.27
R R X H
+ RCH=C(X)Br
Bentonite
R(Br)HC CH(Br)X 1
H N H + H N X
R NH2
R1 R1 + R-CH=N-R1
X = electron withdrawing group Scheme 8.28
8.3 Solvent-free Synthesis 269
The reaction was studied under the action of focused microwaves and conven-
tional heating and it is noteworthy that elimination is more efficient than Michael
addition (leading to aziridines) under irradiation, the conditions being the same.
OSO3H OH
O NH2OSO3H N H2 O N
+ H2SO4
H2O
H2SO4
H
H2SO4/H2O N
HSO4-, +NH 3(CH2)5CO2H O
Scheme 8.29 NaOH
Cyclopentanone Valerolactam 15 60
Cyclohexanone Caprolactam 10 86
Cycloheptanone 2-Azacyclooctanone 20 72
Cyclooctanone 2-Azacyclononanone 15 65
Cycloundecanone 2-Azacyclododecanone 15 72
Cyclododecanone 2-Azacyclotridecanone 20 82
a
Irradiation at 30 W, the temperatures reached by the reaction mixture are in the range 100±120 8C, de-
pending on the nature of the ketone.
In a multimode oven the power is too high and cannot be modulated, so only de-
composition products are obtained.
NH 2 Scheme 8.30
OEt KSF N
+ H C OEt
NH 2 OEt N
H
Monomode MW, dry media 60W 5 min. 74%
∆, toluene 110°C 12 h 79%
R1 NH 2 R1 NHCOCF3 R1 N
N-trifluoroacetylation Acid
CF3
R2 NH2 R2 NH2 R2 N
H
R1 NHCOCF3
tions only traces of heterocycles are observed. The difference is discussed in terms of
the polarity of the transition state compared with that of the initial state (cf. Chapt. 3
and Ref. [49 b]).
R
R-CHO NH 2
Clay 1 N
R
+ N X N
MW N X
R1-NC Y H Y
X=Y=C
X=C;Y=N
X= N ; Y = C Scheme 8.32
8.3 Solvent-free Synthesis 271
Scheme 8.33 O
H H
CO2CH3
S NHBoc a, b S NH
S
N N
N R N R
H
H O H O
a) adsorption on flash column silica gel (250-400 mesh)
b) irradiation 5 minutes at 500 W
H Ar
R1O2C CO2R1
O 1+ 2
+ 2 H C C CO2R R NH 2
Ar H N
Scheme 8.34 R2
thetic Organic Chemistry (ECSOC-5), a special session of which was devoted to mi-
crowave-assisted synthesis.
Among various solid supports (K10, acidic alumina, zeolite HY, silica gel) silica
gel proved to be the best; after irradiation for 4 min the reactions led to yields
ranging from 62 to 94 %.
R3
R2 R2
InCl3 / SiO 2
R + R1 R3 R
MW
NH 2
O 5 to 10 minutes N R1
55 to 87% yield
Scheme 8.35
H
O NH 2 N O
H+
OEt +
EtO MW
O NH 2 N O
Scheme 8.36 H
272 8 Microwaves in Heterocyclic Chemistry
O R1 R2
4 SiO 2
R1 R3 + R NH 2 +R5 NO 2
MW R5 N R3
R2
R4
R5 R1
R5 NO 2
H Al 2O3
R1 + R2NH 2 +
O R4 MW R4 N H
R2
O R2
R2 NO 2
Al 2O3
+ R1NH 2 + ( )n
( )n
R3 MW R3 N
R1 Scheme 8.37
A wide range of starting compounds was used and yields of isolated products ran-
ged from 60 to 86 % for irradiation times of 5 to 15 min. This green procedure avoid-
ing solvents is a better and more practical alternative to existing methods. Different
heterocycles [57 a], e. g. furan, pyrrole, N-benzylpyrrole, indole, and pyrazole react
with methyl a-acetamidoacrylate to give a-amino acid precursors under irradiation
with silica-supported Lewis acids as catalysts. In homogeneous catalysis, long reac-
tion times were required. The reaction of vinylpyrazoles with imines has also been
realized [57 b].
A variety of bisindolylnitroethanes [58] has been obtained in high yields (70±86 %)
by reacting 3-(2-nitrovinyl)indole with indole and 1- or 2-alkylindoles on silica gel
(TLC-grade) under irradiation (Scheme 8.38).
Silica gel or zeolite HY have also been used successfully for the synthesis of imida-
zoles, isoxazoles, and pyrazoles [59], under irradiation, without solvent (Scheme 8.39).
NO 2 NO 2
X
Silica gel (TLC) (3-Indolyl) X
+ R'
N N MW ; 7-10 min. R'
or rt ; 8-14 h N
H R 2 eq.
1 eq. R
a: R = R' =X = H d: R = Et ; R' = X = H
MW: 70-86%
b: R = CH3 ; R' = X =H e: R = i-Pr ; R' = X = H
rt: 69-84%
c: R = CH3 ; R' = H, X = Br f: R = X = H ; R' = CH3
Scheme 8.38
Ph O Ph
Ar O Zeolite HY N
+ + RNH 2 + NH4OAc
H or silica gel
Ph O Ph N Ar
MW
R
Scheme 8.39 without RNH 2 : R = H
CO2Me
O O
+ MeO
OMe
H2N OH H2N O O
O
Conventional heating 45 min. (54%)
Scheme 8.41 MW 8 min. (61%)
hyde) with ethyl and methyl malonate under solvent-free conditions and with MW ir-
radiation (Scheme 8.40).
Irradiation for 2±15 min, rather than many hours under reflux in a variety of sol-
vents in the presence of base (pyridine or piperidine), gives high yields.
Some 7-aminocoumarins-4-carboxylates [60 a] were synthesized by the Pechmann
reaction, by microwave irradiation of the reactants on solid supports (graphite±K10)
(Scheme 8.41). Synthesis of unsubstituted coumarins (C-4 position) has been also re-
ported [60 b].
R1 R2 Scheme 8.42
R1 R2 HO pTSA
+ O O + H2O
O MW
HO
Example: R1 = Ph; R2 = H
Focused microwave oven, EG 1.5 equiv. 15 min 80 8C 80 % [61]
Domestic oven, EG: 10 to 15 equiv. 2 min 650 W 81% [62]
Example: R1 = Ph; R2 = CH3
Focused microwave oven, EG 3 equiv. 30 min 120 8C 90 % [61]
Domestic oven, EG: 10 to 15 equiv. 2 min 650 W 71% [62]
Unfortunately, the comparison is not reliable because the temperature was not
measured in the domestic oven. Dioxolane formation by acid-catalyzed exchange be-
tween 2,2-dimethyl-1,3-dioxolane (DMD) and a ketone in a inert solvent, or simply
in excess DMD, requires 4 to 7 h under classical conditions [63]. This reaction
is readily achieved under microwave irradiation in high yields in 4 to 30 min
(Scheme 8.43).
R1 R2 Me Me R1 R2
MW
+ O O O O + Me2CO
O catalyst
DMD 60 to 100% Scheme 8.43
R1 R2 DMD Catalyst Power (W) Time (min) T max (pC) Yield (%)
It has been shown [61] that yields are higher under the action of irradiation. This
dioxolane exchange was subsequently scaled up to 250 g in the Synthewave 1000
with the same yields [64] and this type of carbonyl protection was extended to dithio-
lanes and oxathiolanes [65].
8.3.2
Reactions under Basic Conditions
Me OSiMe 3 Ph
K10, MW NHPh
PhCH N Ph +
H OMe 5 min., 65% CO2Me
Me
KF,
18-C-6
MW, 7 min., 93%
Ph Ph
N
Scheme 8.45 Me O
H
O Al 2O3 N
+ HN NH 2
2 MW, 15 min. N
CO2Et
86%
Scheme 8.46 O
276 8 Microwaves in Heterocyclic Chemistry
CO 2Et R
R KF/LiCl/Al 2O3
+ CO 2Et CO 2Et
H MW, 150 W O
O NHCOMe NHCOMe
O Scheme 8.47
N Basic clay N
+ CO2Et
N N
H
The reaction proceeds with high yields and selectivity under the action of irradia-
tion and is more efficient than with conventional heating.
Quite recently [71] an expeditious solvent-free synthesis of imidazoline derivatives,
using basic or neutral alumina under microwave irradiation, was reported. The reac-
tion time was reduced from hours to minutes with improved yield compared with
conventional heating.
Scheme 8.49 R1
1 2
R R X R2
+ X-CH2-CO2R'
O N-NH-Ph N
O N
Ph
- H2 O - R'OH
R2
1
R N-NH-Ph
—
OR'
X
O
tained by use of classical heating under the same conditions. With the dry media
procedure it was possible to isolate the intermediate alkene, which was not obtained
in the previously reported procedure. When the active methylene compound is a
keto ester the reaction follows a different pathway [73].
R2 CHO MW R2 R3 - CH2=C=O R2 R3
3
+ R CH2CN
Al 2O3
R1 NHAc R1 N NH R1 N NH 2
3 H
R = CN or CO2Et O C
Scheme 8.50 H2
R
O
EtO2C CO2Et
NH3, H2O
RCHO + O
N
O
H
R = XPh, iPr, n-heptyl, cyclohexyl
Scheme 8.51 yields from 51 to 92% in 10 minutes
R S N Scheme 8.52
S N O S N ( )n
R O ( )n
N N N
Method A or B MW, 70°C
O O catalyst O
catalyst MW, ∆ R1
O 2
catalyst R
R1 N R1
S N 2
S
R
NH 5% N N R2
OH - H2O
O O
O O
KMnO 4/Al 2O3
Ph
N Ph N +
O
15 min. O Me
H Me H
Domestic oven 255 W 73%
Monomode reactor 300 W 140°C 83%
Classical heating 140°C <2% Scheme 8.53
O KF/Al 2O3 O
+ CH3CH2COCl
MW O
OH O CCH2CH3 Scheme 8.54
8.3 Solvent-free Synthesis 279
R1 R3 R4
O O N N +
R2 O O
R3
N R1 = Ph, R2 = H
O N R1 = R2 = Me
R4 O Me
Ph
MW N Me
O R1 20°C
Piperidine Piperidine O N
N N
O R
4
4-20°C R1 = Ph R3 4 R2 O
4 O R O R
R R2 = H 3 OH R 4
OH R1 = R2 = Me R
3 R = H or Me
R 3
O N R
N H
Ph
Scheme 8.55
8.3.3
Enzymatic Catalysis in ªDry Mediaº
Enzymes are usually used in aqueous or organic media and the temperature is lim-
ited to 40 8C to preserve enzyme activity; as a consequence the reactions need very
long times. With enzymes immobilized on solid supports [80] it is possible to ope-
rate at higher temperatures.
Systems such as Pseudomonas lipase dispersed inside Hyflo Supercell (a diatoma-
ceous silica of pH 8.5±9) and SP 435 Novozym (Candida antarctica lipase grafted
on an acrylic resin) are thermally stable and have optimum activity in the range 80±
100 8C. They can therefore be used with conventional or microwave heating if the
temperature is strictly controlled.
OH OCOR
Novozym 435
HO O + R CO2H HO O + H 2O
HO HO
Scheme 8.56 OH OCH3 OH OCH3
280 8 Microwaves in Heterocyclic Chemistry
8.3.3.2 Transglycosylations
The transesterification of phenyl-b-d-glucoside catalyzed by Sulfolobus solfataricus
[81 b] was quantitative within 2 h at 110 8C under the action of irradiation whereas
the yield was only 60 % after 40 h under classical conditions (Scheme 8.57).
OH
OH Enzyme
O +
HO OPh HO
HO Me (support)
OH
OH OH
Me + HO O
HO O O
O OH HO OH
HO
OH OH
Me Scheme 8.57
8.3.4
Solvent-free Solid±Liquid Phase-transfer Catalysis (PTC)
This method is described in Chapt. 5; it is specific for anionic reactions because it in-
volves ªanionic activationº.
Ph O KOtBu (2eq.) Ph O
NBu 4Br (10%)
O Br O
MW monomode reactor 75W 5 min. 75°C 87%
Ultrasound (cleaning bath) 5 min. 75°C 55%
Conventional heating (oil bath) 5 min. 75°C 36%
KOH
+ R-X Aliquat
HOH 2C CH2OH ROH 2C CH2OR
O (10%) O
Reaction times were improved by microwave irradiation, and the same conditions
were extrapolated to the synthesis of a series of new furanic diethers by alkylation of
furfuryl alcohol by dihalides [83].
HO RO
O Base, PTC O
+ R-X
O MW O
OH OR
Scheme 8.60 Yields > 90%
CHO
K2CO3/TBAB
+ ClCH2CO2R3 CO2R3
R1 MW, 8-10 min. R1 O
OH
R2 R2
R1 = H, Et3N R3 = Et, allyl, cyclohexyl Yields : 65 to 96%
Scheme 8.61 R2 = H, OMe
282 8 Microwaves in Heterocyclic Chemistry
The mixture was stirred and then irradiated for 8 to 10 min. After extraction with
CH2Cl2 and evaporation, the products were purified by column chromatography. In
an oil bath, the full conversion of an aldehyde needs usually 3 h.
8.3.5
Solvent-free Reactions without Support or Catalyst
CH3 R CH3
N MW H N
R-CHO + NH NH
O N O N
H H Scheme 8.62
At 160±170 8C under the action of focused irradiation (40±60 W), condensation oc-
curs within 45 s to 4 min, without solvent, without base, and without catalyst. Unfor-
tunately, no comparisons were made with conventional heating and classical me-
thods.
H X
N R1 R2N=C=X N NHR 2
N EWG N EWG
H H
R1 = H ; EWG = CO2Et, CN R1 = H ; EWG = CN ; X = O, S Scheme 8.63
8.3 Solvent-free Synthesis 283
Synthesis of 2,3-dihydroimidazo-[1,2-c]-pyrimidines
Focused irradiation of N-acylimidates mixed with imidazolidine ketene aminals
provides a new means of access to 2,3-dihydroimidazo-[1,2-c]-pyrimidines [90]
(Scheme 8.64).
H X
R2 H R1 N
N - EtOH
N +
1
EtO R N X
- H2 O N N
O
H R 2
X = CN
Scheme 8.64 X = CO2Et
Ar
Ar NH 2 EtO2C + EtO 2C H
H N
+ +
H O H2N Z MW or ∆
Me O Me N Z
Scheme 8.65 Z = O, S H
Ar H
EtO 2C
N O MW (400 W) Ar H OH
+
Ph N NH 2 20 min. EtO 2C
O N O
H
O H Ph N N H
H
60 to 80%
Ar H Ar H O
EtO 2C EtO CO2Et EtO 2C CO2Et
N MW (400 W) N
+
10 min.
Ph N NH 2 H CO2Et Ph N N H
H H
Yields > 95%
Scheme 8.66
O O
O + NH2R N R + H2O
O O
R = PhCH2, -(CH2)6OH, -(CH 2)7CH3, -CH2CO2H Scheme 8.67
O O
X
N N
H2N NH 2
O O
O MW O
N O
OHOH O H
X = O 63%, 10 min.
X = S 85%, 15 min. Scheme 8.68
More recently this procedure has been used for a new synthesis of thalidomide,
now a ªrehabilitated drug [94]º (Scheme 8.68).
Pyrazoles
Condensation of aromatic acyl compounds with N,N-dimethylformamide diethyl
acetal in a pressure tube under the action of microwave irradiation affords easy ac-
cess to 1-aryl-3-dimethylaminoprop-2-enones in almost quantitative yield after
6 min. These intermediates can then be reacted with hydrazine hydrate under con-
ventional reflux in ethanol to form the corresponding 3-substituted pyrazoles [95]
(Scheme 8.69).
b-Hydrazinoacrylates derived from benzimidazole, benzoxazole and benzothiazole
are readily prepared in good yields by transamination of the corresponding 3-di-
methylamino acrylates with a variety of hydrazines [96] (Scheme 8.70).
8.3 Solvent-free Synthesis 285
Scheme 8.69 H
OCH 3 N N
O MW N2 H4, H 2O
+ H N(CH3)2 N
R 360 W O Reflux, EtOH
OCH 3
R R
R = C 6H5, C5H 4N, C6H4OH, naphthyl
R1
N R2
N NMe 2 R1 N N
+ H2N N H
X CO2Et R2 -Me 2NH
71-95% X CO2Et
X = NH, O, S
N NMe 2 R1 = H N H
N
X -EtOH X N
CO2Et
74-96% R2
X = NH, O O
R1 Me Me C6H5 CO2Me
H2N N H2N N H2N N H2N N H2N N
Scheme 8.70 R2 H Me H H
R R Scheme 8.71
N4 MW N N N
1
2
N + R-X
No solvent
N + N + N X
N N N N
H R R
N1 N4 N1, N4
N
N
ClH2C N
R-X = C Cl H2C
O C Cl
Cl O
MW 20 minutes 140°C N1 92% Cl
∆ oil bath 140°C 36% N1 / 27% N4 / 37% N1, N4
NHR 2
OH
MW, 600 W
+ R2NH 2
14-20 min.
H3C N O
H3C N O
R1
Yields 60 to 91% Scheme 8.72
(20 mmol) methylacetoacetate (20 mmol) and arylaldehydes (20 mmol) in a domestic
oven. Yields from 59 to 77 % are reported for 10 min reaction.
A variety of conditions (solution, dry media, solvent-free) has been used for micro-
wave-assisted synthesis of Hantzsch 1,4-DHP; only procedures involving solvent-free
conditions under the action of irradiation led to the aromatized pyridine derivatives.
2-Pyridones were studied for N- and C-alkylation reactions by de la Hoz et al.
[100]; as already mentioned for 1,2,4 triazoles, the selectivity of the alkylation is
highly dependent on the activation technique (microwave or conventional heating).
The microwave-assisted nucleophilic substitution of 4-hydroxy-6-methyl-2(1H)-
pyridones [101] has also been studied (Scheme 8.72).
It is noteworthy that the reaction works with araliphatic amino compounds. This
is, nevertheless, an improvement on classical methods.
OH
CO2Et R2
+ R 2 MW, 500 W
8.4
Room-temperature Ionic Liquids (RTIL) ± Synthesis and Applications in Organic
Synthesis under the Action of Microwaves
RTIL are a new class of solvents composed entirely of ions, most often alkylpyridi-
nium or dialkylimidazolium salts and a Lewis acid [103]. These solvents have a num-
ber of interesting properties, especially their lack of vapor pressure, a wide accessible
temperature range, lack of flammability, and ease of reuse [104]. RTIL have emerged
as a set of new green solvents [105], mainly as a replacement for conventional volatile
organic solvents. The use of a large excess of conventional volatile solvents to per-
form a chemical reaction is of ecological and economic concern. Much attention has
been focused on organic reactions catalyzed by RTIL, and several organic reactions
of high importance catalyzed by ionic liquids have been reported.
8.4.1
Synthesis of 1,3-Dialkylimidazoliums as RTIL
240W R
N N R X N N
70-100°C ,X
Scheme 8.74 70 71 (X = Cl, Br, I) 72
288 8 Microwaves in Heterocyclic Chemistry
Tab. 8.5 Comparison of conventional and microwave heating for preparation of ionic liquids.
1-Chlorobutane 30 + 15 + 15 + 15 + 15 72a 76 50
1-Bromobutane 30 + 15 + 15 + 15 72b 86 76
1-Bromohexane 30 + 15 + 15 + 15 72c 89 78
a
Use of MW, power 240 W
b
Oil bath at 80 8C
friendly method uses only stoichiometric amounts of reactants and yields were en-
hanced.
This method has some major drawbacks, however:
. the hygroscopic nature of salts might not enable large-scale preparation by the pre-
vious method, because the reaction is performed in an open vessel;
. heating irritant volatile alkyl halides in an open container in a microwave oven
can be hazardous;
. the waste contains corrosive and hygroscopic methyl imidazole.
Khadilkar and Rebeiro have investigated a new method [107] that overcomes all
these problems and is far safer. The authors used closed pressure reactor [108], with
no apparent loss of yield. The microwave reactor used for these reactions has a possi-
bility for recording temperature and pressure during irradiation. For example, 1-bu-
tyl-3-methylimidazolium chloride was isolated in 91% yield in 24 min [109] at 150 8C
and 57 psig was the maximum pressure reached.
8.4.2
N-alkylation in RTIL
Indole 73 can be selectively alkylated [110] (Scheme 8.75) by benzyl bromide to give
74 in high yield (>90 %) when the reactants were mixed with 1-butyl-3-methylimi-
dazolium hexafluorophosphate ([bmim][PF6]) as polar solvent and the reaction was
performed with brief (1 min) microwave irradiation at 180 8C. Product 74 was fully
extracted with diethyl ether and [bmim][PF6] was reused in another cycle of syn-
thesis.
Ba(OH) 2
Br
N
N [bmim][PF 6]
H MW180°C, 60 s
73 74 (>90%) Scheme 8.75
8.5 Conclusions 289
8.4.3
Knoevenagel Reactions on a Grafted Ionic Liquid Phase
Previous work [111] by our group has demonstrated that RTIL-catalyzed 1,3-dipolar
cycloaddition under the action of microwave irradiation leads to dramatically shorter
reaction times with better yields of isolated products. We have recently investigated
the reactivity of the formyl group covalently grafted on the ionic liquid phase 75
in the Knoevenagel reaction with malonic derivatives 76 [112], as shown in
Scheme 8.76.
CN 76
O CO2Me (CN) O
N N O N N O
O O CN
, BF4 O piperidine (2%) 113 , BF4
75 MW, 80°C 77 (98%) CO2Me (CN)
15 min.
OH
MeONa, MeOH N N
MeO2C O
25°C, 18 h. CN , BF4
78 (97%) CO2Me (CN) [hydemim][BF4 ]
Scheme 8.76
The bound products 77 prepared in high yields (98 %) with reduced reaction times
(15 min) were subjected to a very efficient cleavage from the IL-phase. The expected
products 78 were extracted with dichloromethane in 97 % yield and the insoluble [hy-
demim][BF4] was reused in another cycle of synthesis.
8.5
Conclusion
1 For reviews on the Heck reaction see: 14 Kidwai, M.; Bhushan, K.R.; Misra, P.
a) de Mejeire, A.; Meyer, F.E. Angew. Indian J. Chem. 2000, 39B, 458.
Chem., Int. Ed. Engl. 1994, 33, 2379; 15 Besson, T.; Guillard, J.; Rees, C.W.
b) Jefferey, T. in: Advances in Metal Or- Tetrahedron Lett. 2000, 41, 1027.
ganic Chemistry,Vol. 5. Liedeskind, L.S. 16 Soukri, M.; Guillaumet, G.; Bes-
(ed.), JAI Press, Greenwich, 1996, son, T.; Aziane, D.; Aadil M.; Essassi,
p. 153. El M.; Akssira, M. Tetrahedron Lett.
2 For a review on the Suzuki reaction see: 2000, 41, 5857.
Miyaura, N.; Suzuki, A. Chem. Rev. 17 Santagada,V.; Persutti, E.; Fio-
1995, 95, 2457. rino, F.; Vivenzio, B.; Caliendo, G.
3 Stille, J.K.; Angew. Chem., Int. Ed. Tetrahedron Lett. 2001, 42, 2397.
Engl. 1986, 25, 508. 18 Fabis, F.; Jolivet-Fouchet, S.;
4 a) Olofsson, K.; Kim, S.Y.; Larhed, M.; Robba, M.; Landelle, M.; Rault S.
Curran, D.P.; Hallberg, A. J. Org. Tetrahedron 1998, 54, 10789.
Chem. 1997,62, 5583; b) Namboo- 19 Rault, S.; Derobert, M. Eur. Pat. Appl.
diri,V.V.; Varma, R.S. Green Chem. 059 5084 A1, 1993.
2001, 3, 146. 20 a) Rault, S.; Cugnan de Svri-
5 a) Larhed, M.; Hoshino, M.; Hadi- court, M.; Robba, M. Rec. Trav. Chim.
da, S.; Curran, D.P.; Hallberg, A. 1982, 101, 205; b) Robba, M.;
J. Org. Chem. 1997, 62, 5583; b) Lar- Lecomte, J.M.; Cugnan de Svri-
hed, M.; Lindeberg, G.; Hallberg, A. court, M. Bull. Soc. Chim. Fr. 1974,
Tetrahedron Lett. 1996, 37, 8219. 2864.
6 Villemin, D.; Gomez-Escolinilla, 21 Hinschberger, A.; Gillard, A.C.;
M.J.; Saint-Clair, J.F. Tetrahedron Lett. Bureau, I.; Rault, S. Tetrahedron 2000,
2001, 42, 635. 56, 1361.
7 Vallin, K.S.A.; Larhed, M.; Johans- 22 a) Saidi, M.R.; Rajabi, F. in: Fifth Int.
son, K.; Hallberg, A. J. Org. Chem. Electronic Conf. Synthetic Organic Chem-
2000, 65, 4537. istry (ECSOC-5), 1±30 September, 2001,
8 Chanda, B.M.; Vyas, R.; Bedekar, A.V. http://www.mdpi.org/ecsoc-5/e0002/
J. Org. Chem. 2001, 66, 30. e0002.htm; b) Moghaddam, F.M.;
9 Saoudi, A.; Hamelin, J.; Benhaoua, H. Sharifi, A.; Saidi, M.R. J. Chem. Res.
J. Chem. Res. (S) 1996, 492. (S) 1996, 338; c) Saidi, M.R.; Big-
10 Bose, A.K.; Banik, B.K.; Barakat, K.J.; deli, M.R. J. Chem. Res. (S) 1998, 800.
Manhas, M.S. Synlett 1993, 575. 23 AlajarÌn, R.; Vaquero, J.J.; GarcÌa-
11 a) Bose, A.K.; Barik, B.K.; Mathur, C.; NavÌo, J.L.; Alvarez-Builla, J. Synlett
Wagle, D.R.; Manhas, M. Tetrahedron 1992, 297.
2000, 56, 5603; b) Bose, A.K.; 24 Khadilkar, B.M.; Madyar,V.R. in: Fifth
Barik, B.K.; Mathur, C.; Wagle, D.R.; Int. Electronic Conf. Synthetic Organic
Manhas, M. Tetrahedron 2000, 56, 5587. Chemistry (ECSOC-5), 1±30 September,
12 Mauro, P.; Eileen, C.; Paola,V.; Gab- 2001, http://www.mdpi.org/ecsoc-5/
riele, B. in: Fifth Int. Electronic Conf. e0021/e0021.htm.
Synthetic Organic Chemistry (ECSOC-5), 25 Hydrotopy is the phenomenon by which
1±30 September, 2001; http://www. otherwise water-insoluble compounds
mdpi.org/ecsoc-5/e0010/e0010.htm. can be solubilized in the aqueous solu-
13 a) Bentiss, F.; Lagrene, M.; Barbry, D. tion of some compounds, e. g. arene sul-
Tetrahedron Lett. 2000, 41, 1539; fonates. These compounds, known as
b) Barbry, D.; Gassama, F. in: Fifth Int. hydrotopes, are readily water-soluble.
Electronic Conf. Synthetic Organic Chemis- The aqueous concentrated solutions (20
try (ECSOC-5), 1±30 September, 2001; to 50 %) of the hydrotopes can signifi-
http://www.mdpi.org/ecsoc-5/e0040/ cantly increase the water solubility of
e0040.htm. many water-insoluble compounds.
References and Notes 291
26 The authors used for the first time aro- see Sect. 4.2.1, Protecting Groups, in this
matic hydrotope solution system such publication.
as 50 % sodium p-toluene sulfonate 39 See Sect. 8.1.3, b-Lactam Chemistry, in
aqueous solution (NaPTSA), 40 % so- this chapter.
dium cumene sulfonate aqueous solu- 40 Daga, M.C.; Taddei, M.; Varchi, G.
tion (NaCuS), and 20 % sodium p-xy- Tetrahedron Lett. 2001, 42, 5191.
lene sulfonate (NaXS) aqueous solution 41 Jones, J.R.; Lockley, W.J.S.; Lu, S.Y.;
to perform Hantzsch ester synthesis. Thompson, S.P. Tetrahedron Lett. 2001,
27 Vanden-Eynde, J.J.; Rutot, D. Tetra- 42, 331.
hedron 1999, 55, 2687. 42 Petit, A.; Loupy, A.; Maillard, P; Mo-
28 Cotterill, I.C.; Usyatinsky, A.Y.; Ar- menteau, M. Synth. Commun. 1992, 22,
nold, J.M.; Clark, D.S.; Dordik, J.S.; 1137.
Michels, P.C.; Khmelnitsky,Y.L. Tetra- 43 Lu, M.W.; Hu,W.X.; Yun, L.H. Proc.
hedron Lett. 1998, 39, 1117. First National Meeting on Microwaves,
29 Brain, C.T.; Paul, J.M.; Loang,Y.; Beijing, China, 1996, p. 54.
Oakley, P.J. Tetrahedron Lett. 1999, 40, 44 Csiba, M.; Clophax, J.; Loupy, A.;
3275. Malthte, J.; Gro, S.D. Tetrahedron
30 a) Kappe, C.O. Eur. J. Med. Chem. 2000, Lett. 1993, 34, 1787.
35, 1043±1052; b) Kappe, C.O. Molecules 45 Saoudi, A.; Hamelin, J.; Benhaoua, H.
1998, 3, 1. J. Chem. Res. (S) 1996, 492.
31 Atwal, K.; Rovnyak, G.C.; Schwartz, J.; 46 Laurent, A.; Jacquault, P.; Di Mar-
Moreland, S.; Hedberg, A.; Gougou- tino, J.L.; Hamelin, J. J. Chem. Soc.
tas, J.Z.; Malley, M.F.; Floyd, D.M. Chem. Commun. 1995, 1101.
J. Med. Chem. 1990, 33, 1510. 47 Olah, G.A.; Keumi, T.; Fung, A.P.
32 Dandia, A.; Saha, M.; Taneja, H. Synthesis 1979, 537.
J. Fluorine Chem. 1998, 90, 17. 48 Villemin, D.; Hammadi, M.; Mar-
33 a) Ionic liquids were recently used as cat- tin, B. Synth. Commun. 1996, 26, 2895.
alysts for the solvent-free Biginelli con- 49 a) Bougrin, K.; Loupy, A.; Petit, A.;
densation reaction: Peng, J.; Deng,Y. Daou, B.; Soufiaoui, M. Tetrahedron
Tetrahedron Lett. 2001, 42, 5917; b) For a 2001, 57, 63; b) Perreux, L.; Loupy, A.
microwave-mediated regioselective Tetrahedron 2001, 57, 9199.
synthesis of novel pyrimido-[1,2-a]-pyri- 50 Varma, R.S.; Kumar, D. Tetrahedron
midines under solvent-free conditions, Lett. 1999, 40, 7665.
see: Vanden Eynde, J.J.; Hecq, N.; Ka- 51 Paudrey, S.K.; Awasthi, K.K.;
teavu, O.; Kappe, C.O. Tetrahedron 2001, Saxena, A.K. Tetrahedron 2001, 57,
57, 178. 4437.
34 Varma, R.S.; Kappe, C.O. in: Third Int. 52 Balalaie, S.; Kowsari, E. in: Fifth Int.
Electronic Conf. Synthetic Organic Chem- Electronic Conf. Synthetic Organic Chem-
istry (ECSOC-3), 1±30 September, 1999, istry (ECSOC.5),1±30 September, 2001.
http://www.mdpi.org/ecsoc-3/b0008/ 53 Ranu, R.C.; Hajra, A.; Jana, U. Tetra-
b0008.htm. hedron Lett. 2000, 41, 531.
35 Lacova, M.; Gasparova, R.; Loos, D.; 54 Vzquez, E.; de la Hoz, A.; Elan-
Liptay, T.; Pronayova, N. Molecules der, N.; Moreno, A.; Stone-Elan-
2000, 5, 167. der, S. Heterocycles 2001, 55, 109.
36 a) Nohara, A.; Sugihara, H.; 55 Paul, S.; Gupta, M.; Gupta, R.;
Ukawa, K. Japan Kokai Tokyo Koho Loupy, A. Tetrahedron Lett. 2001, 42,
1978, 111070; Chem. Abstr. 1979, 90, 3827.
5482z; b) Klutchko, S.; Kaminski, D.; 56 Hajra, A.; Ranu, B.C. Tetrahedron Lett.
Von Strandtmann, M. US Patent, 2001, 57, 4767; a) de la Hoz, A.; DÌaz-
1977, 4008252; Chem. Abstr. 1977, 87, Ortiz, A.; GÕmez, M.V.; Mayoral, J.A.;
5808a. Moreno, A.; Snchez-MigallÕn, A.M.;
37 Villemin, D.; Martin, B. Synth. Com- Vzquez, E. Tetrahedron Lett. 2001, 57,
mun. 1995, 25, 3135. 5421; b) Carillo, J.R.; Diaz-Ortiz, A.;
38 Caddick, S. Tetrahedron 1995, 51, 10403; Gomez-Escalonilla, M.J.; de la
292 8 Microwaves in Heterocyclic Chemistry
Hoz, A.; Moreno, A.; Prieto, P.Tetra- reau, J.P.; Texier-Boullet, F.; Hame-
hedron 1999, 55, 9623. lin, J. Tetrahedron Lett. 1998, 39, 541.
58 Chakrabarty, M.; Basak, R.; Ghosh, N. 78 Mariani, E.; Genta, M.T.; Bar-
Tetrahedron Lett. 2001, 42, 3913. gagna, A.; Neuhoff, C.; Loupy, A. in:
59 Besson, T.; Thiry,V.; Fr re, S. in: Application of the Microwave Technology
Fifth Int. Electronic Conf. Synthetic Or- to Synthesis and Materials Processing.
ganic Chemistry (ECSOC.5), 1±30 Sep- Acierno, D.; Leonelli, C.; Pella-
tember, 2001; a) Besson, T.; Thiry,V.; cini, G.C. (eds.), Mucchi Editore 2000,
Fr re, S. Tetrahedron Lett. 2001, 42, p. 157.
2791; b) de la Hoz, A.; Moreno, A.; 79 Jolivet-Fouchet, S.; Hamelin, J.;
Vasquez, E. Synlett 1999, 608. Texier-Boullet, F.; Toupet, L.; Jac-
61 Perio, B.; Dozias, M.J.; Jacquault, P.; quault, P. Tetrahedron 1998, 54, 4561.
Hamelin, J. Tetrahedron Lett. 1997, 38, 80 Guib-Jampel, E.; Rousseau, G. Tetra-
7867. hedron Lett. 1987, 28, 3563.
62 Matloubi-Moghaddam, F.; Sharifi, A. 81 a) Gelo-Pujic, M.; Guib-Jampel, E.;
Synth. Commun. 1995, 25, 2457. Loupy, A.; Galema, S.A.; Math, D.
63 Rotstein, D.M.; Kertesz, D.J.; Wal- J. Chem. Soc., Perkin Trans I 1996, 2777;
ker, K.A.; Swinney, D.C. J. Med. Chem. b) Gelo-Pujic, M.; Guib-Jampel, E.;
1992, 35, 2818. Loupy, A.; Galema, S.A.; Math, D.
64 Perio, B.; Dozias, M.J.; Hamelin, J. J. Chem. Soc., Perkin Trans I 1997, 1001.
Org. Proc Res. Dev. 1998, 2, 428. 82 DÌaz-Ortiz, A.; Prieto, P.; Loupy, A.;
65 Perio, B.; Hamelin, J. Green Chem. AbenhaÒm, D. Tetrahedron Lett. 1996,
2000, 2, 252. 37, 1695.
66 Limousin, C.; Clophax, J.; Petit, A.; 83 Majdoub, M.; Loupy, A.; Petit, A.;
Loupy, A.; Lukacs, G.J. Carbohydr. Res. Roudesli, S. Tetrahedron 1996, 52, 617.
1997, 16, 327. 84 Chatti, S.; Bortolussi, M.; Loupy, A.
67 Texier-Boullet, F.; Latouche, R.; Ha- Tetrahedron Lett. 2000, 41, 3367.
melin, J. Tetrahedron Lett. 1993, 34, 85 Chatti, S.; Bortolussi, M.; Loupy, A.
2123. Tetrahedron. 2000, 56, 5877.
68 Pilard, J.F; Klein, B.; Texier-Boul- 86 Bogdal, D.; Warzala, M. Tetrahedron
let, F.; Hamelin, J. Synlett 1992, 219. Lett. 2000, 56, 8769.
69 AbenhaÒm, D.; Loupy, A.; Mahieu, C.; 87 Villa, C.; Genta, M.T.; Bargagna, A.;
Smria, D. Synth. Commun. 1994, 24, Mariani, E.; Loupy, A. Green Chem.
1809. 2001, 3, 196.
70 MartÌn-Aranda, R.M.; Vicente-Ro- 88 Villemin, D.; Martin, B. Synth. Com-
drÌguez, M.A.; LÕpez-Pestaµa, J.M.; mun. 1995, 25, 3135.
Banares-Muµoz, M.A. J. Mol. Catal. A: 89 Cado, F.; Di Martino, J.L.; Jac-
Chem. 1997, 124, 115. quault, P.; Bazureau, J.P.; Hamelin, J.
71 Kidwai, M.; Sapra, P.; Bushan, K.R.; Bull. Soc. Chim. Fr. 1996, 133, 587.
Misra, P. Synthesis 2001, 1509. 90 Rahmouni, M.; Derdour, A.; Bazu-
72 Jolivet, S.; Texier-Boullet, F.; Hame- reau, J.P.; Hamelin, J. Synth. Com-
lin, J.; Jacquault, P. Heteroatom mun. 1996, 26, 453.
Chem. 1995, 6, 469. 91 Kappe, C.O.; Stadler, A. J. Chem. Soc.,
73 Jolivet, S.; Toupet, L.; Texier-Boul- Perkin Trans. 2000, 2, 1363.
let, F.; Hamelin, J. Tetrahedron 1996, 92 Vanden Eynde, J.J.; Hecq, N.; Ka-
52, 5819. taeva, O.; Kappe, C.O. Tetrahedron
74 Sharma, U.; Ahmed, S.; Boruah, R.C. 2001, 57, 1785.
Tetrahedron Lett. 2000, 41, 3493. 93 Vidal, T.; Petit, A.; Loupy, A;
75 Ohberg, L.; Westman, J. Synlett 2001, Gedye, R.N. Tetrahedron 2000, 56, 5473.
8, 1296. 94 Seijas, A.; Vzquez-Tabo, M.P.; Gon-
76 Cherouvrier, J.R.; Boissel, J.; Car- zlez-Baude, C.; MartÌnez Mar-
reaux, F.; Bazureau, J.P. Green Chem. tÌnez, M.; Pacios-Lopez, B. Synthesis
2001, 3, 165. 2001, 999.
77 Benhaliliba, H.; Derdour, A.; Bazu- 95 Pleier, K.; Glas, H.; Grosche, M.;
References and Notes 293
Sirsch, P.; Thiel,W.R. Synthesis 2001, 105 a) Freemantle, M. Chem. Eng. News,
55. 1 January, 2001, p. 21; b) Free-
96 Meddad, N.; Rhamouni, M.; Der- mantle, M. Chem. Eng. News, 15 May,
dour, A.; Bazureau, J.P.; Hamelin, J. 2000, p. 37; c) Carmichael, H. Chem.
Synthesis 2001, 581. Ber. 2000, 36.
97 a) Cleophax, J.; Liagre, M.; Loupy, A.; 106 Varma, R.S.; Namboodiri,V.V. J. Chem.
Petit, A. Org. Proc Res. Dev. 2000, 6, Soc., Chem. Commun. 2001, 643.
498; b) AbenhaÒm, D.; Diaz-Barra, E.; 107 Khadilkar, B.; Rebeiro, G.L. in: Fifth
de la Hoz, A.; Loupy, A.; Sanchez-Mi- Int. Electronic Conf. Synthetic Organic
gallon, A. Heterocycles 1994, 38, 793. Chemistry (ECSOC-5), 1±30 September,
98 Vanden Eynde, J.J. Mayence, A. Fifth 2001, http://www.mdpi.org/ecsoc-5/
Int. Electronic Conf. Synthetic Organic e0020/e0020.htm.
Chemistry (ECSOC.5), 1±30 September, 108 Reactions were realized in the CEM mi-
2001. crowave digester MARS 5; http://
99 Zhang,Y.W.; Shen, Z.X.; Pan, B.; www.cem.com/products/MARS5.html.
Lu, X.H.; Chen, M.H. Synth. Commun. 109 For the total reaction time (24 min.),
1995, 85, 857. 2 min were used to reach the given
100 Almena, I.; DÌaz-Ortiz, A.; DÌez- temperature (1508C) and then 22 min
Barra, E.; de la Hoz, A.; Loupy, A. to continue irradiation at the given tem-
Chem. Lett. 1996, 333. perature.
101 Stoyanov, E.V.; Heber, D. Synlett. 110 Westman, J. Patent WO 00/72956 A1,
1999, 11, 1747. 2000.
102 Lange, J.H.M.; Verveer, P.C.; Osna- 111 Fraga-Dubreuil, J.; Bazureau, J.P.
brug, S.J.M.; Visser, G.M. Tetrahedron Tetrahedron Lett. 2000, 41, 7351.
Lett. 2001, 42, 1367. 112 Fraga-Dubreuil, J.; Bazureau, J.P.
103 Seddon, K.R. J. Chem. Technol. Biotech- Tetrahedron Lett. 2001, 42, 6097.
nol. 1997, 68, 351. 113 Abdallah-El Ayoubi, S.; Texier-Boul-
104 Welton, T. Chem. Rev. 1999, 99, 2071. let, F.; Hamelin, J. Synthesis 1994, 258.
295
9
Microwaves in Cycloadditions
Antonio de la Hoz, Angel DÌaz-Ortiz, and Fernando Langa
9.1
Introduction
9.2
Reactions in Solution
9.2.1
Reactions under Pressure
9.2.2
Reactions under Reflux
Reactions involving a solvent under reflux require the use of modified commercial
microwave ovens. In these modified systems the oven is perforated on the top to ac-
commodate a reflux condenser and a 10 cm pipe is used to avoid microwave leakage;
the turnable dish is replaced by a magnetic stirrer or by monomode reactors espe-
cially designed for chemical synthesis [15].
The advantage of reactions involving solvent is that the reaction temperature is
controlled by the reflux temperature of the solvent [16]. It should, however, be re-
MeO2 C
MeO2 C CO 2Me
2 CO 2Me
1 3
MICROWAVE CONTROL
10 min 4h
p-xylene, 87% p-xylene, 67%
Scheme 9.1 325º<361 ºC 138 ºC
9.3 Solvent-free Conditions 297
membered that overheating by between 13 and 26 8C above the normal boiling point
of polar solvents can occur as a result of the ªinverted heat transferº effect, because
boiling nuclei are formed at the surface of the liquid [17]. This superheating effect
was shown to disappear when the experiments were performed with efficient stir-
ring and use of low microwave power [18].
For instance, cycloadditions of [60]fullerene (4) under the action of microwave ir-
radiation usually require the use of this technique, because reactions are performed
on a very small scale and C60, in common with many dienophiles, does not absorb
microwaves efficiently [19].
9.2.3
Microwave Organic Reaction Enhancement (MORE)
9.3
Solvent-free Conditions
Solvent-free conditions are especially suitable for microwave activation. Several ad-
vantages of this approach are evident [22]:
. in the absence of solvent the radiation is absorbed directly by the reagents, so the
effect of microwaves is more marked.
. solid supports can be used efficiently; many mineral oxides are poor conductors of
heat yet absorb microwave radiation very efficiently [23].
298 9 Microwaves in Cycloadditions
. the use of large volumes of solvent can be avoided, thus reducing solvent emis-
sions and precluding the need for redistillation;
. work-up procedures are considerably simplified, because the pure product can of-
ten be obtained directly from the crude reaction mixture by simple extraction, dis-
tillation, or sublimation;
. recyclable solid supports can be used efficiently in place of mineral acids and oxi-
dants;
. scale-up is facilitated by the absence of solvent; and
. under these conditions, reaction times are very short, which leads to savings in
time, money, and energy.
9.3.1
Reactions using Mineral Supports
Alumina, silica, clays, and zeolites are increasingly used as acidic or basic supports
[26]. Cycloaddition reactions often require Lewis-acid catalysts if good yields are to be
obtained. Clay and doped silica gel catalysts have emerged as useful alternatives to
the use of Lewis acids. Cycloaddition of furan (5) under solvent-free conditions, cata-
lyzed by K10 montmorillonite, results in a decrease in the reaction time; the endo±exo
relationship is no different that obtained by use of classical heating (Scheme 9.2) [27].
The use of silica-supported Lewis acids as catalysts for the Diels±Alder reactions of
2,5-dimethylfuran leads to fairly good yields of adducts [28]. Solid supports such as
O O O
O
K 1O
+ NPh O + NPh
O
NPh O
O O
5 6 7 8
Scheme 9.2
9.3 Solvent-free Conditions 299
silica gel [29] and alumina [30] have also been used to generate 1,3-dipoles and, in
this way, heterocyclic compounds can be obtained in a one-pot procedure. In these
reactions changes in the regioselectivity of the reaction have not been observed un-
der the action of microwave irradiation.
9.3.2
Reactions without Support
An alternative to the use of supported reagents is the use of uncatalyzed ªneat reac-
tionsº. Under these conditions the radiation is absorbed directly by the reagents and
this results in spectacular acceleration, higher yields and purity of the reaction pro-
ducts, a simple workup procedure, and sometimes changes in the selectivity of the
cycloaddition.
Diels±Alder reactions [31] and 1,3-dipolar cycloadditions [32, 33] have been per-
formed by use of this methodology. For example, DÌaz-Ortiz described the hetero-
Diels±Alder and 1,3-dipolar cycloaddition reactions of ketene acetals. The reactions
were improved and products were isolated directly from the crude reaction mixture
without polymerization of the ketene acetals [34].
When the enantiomerically pure ketene acetal 16 was used, cycloaddition with
N,a-diphenylnitrone (11) or chalcone (12) was achieved within 3 min in 98 and 96 %
yield, respectively. Under the same reaction conditions the use of classical heating in
the absence of solvent at 120±124 8C for 3 min caused the yields to decrease to 3±
4 %. These results suggest that the excellent yields achieved by use of microwave irra-
diation are perhaps not entirely a result of the rapid heating of the reaction mass
(Scheme 9.3).
9.3.3
Reactions with a Heat Captor
Garrigues described the use of graphite as a heat captor under the action of micro-
wave irradiation. Graphite is a chemically inert support that couples strongly with
microwaves by a conduction process and is able to transmit intense thermal energy
to the supported reagents [35].
Methods have been described that involve microwave-assisted graphite-supported
dry media for the cycloaddition of anthracene, 1-azadienes and 1,2,4,5-tetrazines
with several C±C dienophiles and carbonyl compounds in hetero-Diels±Alder reac-
tions [35]. This technique leads to a shortening of reaction times, a situation that en-
ables work to be undertaken at ambient pressure in an open reactor to avoid the for-
mation of unwanted compounds by thermal decomposition of reagents or products.
Similarly, the use of a higher input power in retro-Diels±Alder reactions of anthra-
cene derivatives has been reported to afford complete reaction in 3±5 min [36]. This
method is an alternative to the use of flash thermolysis. The use of graphite is a pre-
requisite for obtaining high temperatures in a short time.
This methodology has been extended by the same author to include carbonyl-ene
reactions [37]. The cyclization of (+)-citronellal (19) to pulegols 20 and 21 is faster
300 9 Microwaves in Cycloadditions
O O
O Ar N
Ar C N O O 11 Ph O
O 10 Ar
O O
MW O MW H N
Ar N
9
13 Ph
R1 14
MW
12
O R2
R1 H
O
R2 O
O
15
H Ph Ph H
O Ph O Ph
N N
Ph O Ph O
Ph O O
MW 17a (38%) 17b (36%)
O O + Ph N
3 min H Ph
O Ph H
Ph 16 11 O
N O
Ph O N
O Ph O
Ph O
17c (18%) 17d (8%) Ph
H Ph Ph H
O O
Ph O Ph Ph
Ph Ph O
O O
MW 18a (43%) 18b (37%)
O O +
3 min H Ph Ph H
Ph Ph O
16 12
O O
Ph O Ph O
O O
18c (14%) Ph 18d (6%) Ph
Scheme 9.3
9.4 Specific Effects in Cycloaddition Reactions 301
MW
+
CHO Graphite
OH OH
80%
Scheme 9.4 19 20 21
when performed under the action of microwave irradiation with a graphite support
and the stereoselectivity was different from that obtained by use of classical heating,
with the amount of (+)-neoisopulegol (21) being greater in the former (Scheme 9.4).
9.4
Specific Effects in Cycloaddition Reactions
The existence of results that cannot be explained solely by the effect of rapid heating
has led authors to postulate the existence of a so-called ªmicrowave effectº. Hence,
acceleration or changes in reactivity and selectivity could be explained by a specific
radiation effect and not merely by a thermal effect. Several reviews have collected
synthetic results that have been attributed to the microwave effect [5 b, 38, 39].
Several authors have proposed that changes in thermodynamic behavior under
the action of microwave irradiation are the cause of the ªmicrowave effectº. When a
compound absorbs microwaves, however, dielectric heating causes an increase in
the temperature of the system. When the internal energy of the system is raised it is
distributed among translational, rotational, or vibrational energies, irrespective of
the mode of heating. Consequently, it was concluded there should be no kinetic dif-
ferences between reactions heated by microwaves or by classical heating if the tem-
perature is known and the solution is thermally homogenous [8 c, 40].
The existence of the so-called ªmicrowave effectº has not been proved. It does,
however, seem to have been demonstrated that overheating of polar liquids [17] oc-
curs and that ªhot spotsº are present in heterogeneous systems, especially at the in-
terface [38]. Similarly, microwave irradiation results in an increase in the molecular
mobility in solids [5 b].
The utility of microwaves in improving numerous processes or in modifying
chemo, regio-, or stereoselectivity is evident. These changes often seem to arise from
the rate of heating which results from use of microwaves, a situation that is not pos-
sible with classical heating. Suard considered two important differences between
conventional heating and microwave radiation [41 a]. First, under the action of micro-
wave irradiation the initial slope of the sample temperature is different from zero
and, second, in contrast with the situation observed with conventional heating, con-
duction flow is the major flow of the system. As a consequence, a rapid heating rate
must be responsible for the observed effects [41 b].
In cycloaddition reactions few examples have been described in which changes in
selectivity have been observed on use of microwave irradiation. In concerted pro-
302 9 Microwaves in Cycloadditions
cesses the regio- and stereoselectivity of the reaction is governed by frontier orbital
interactions, and changes in selectivity are not expected unless a change in the reac-
tion mechanism occurs.
These effects have been justified in competitive reactions by two compatible hy-
potheses:
1. by considering that under the action of microwave irradiation the more polar
route will be favored [42]; and
2. by considering that the relative ratio of isomers is related to the hardness and, un-
der the action of microwave irradiation, formation of the intermediate of greatest
hardness should be favored [43].
9.5
[4+2] Cycloadditions
9.5.1
Diels±Alder Reactions
The usefulness of the Diels±Alder reaction in organic synthesis comes from its ver-
satility and its high regio- and stereoselectivity. A wide variety of dienes and dieno-
philes can be used with a varied range of substitution patterns. Numerous types of
ring structure can be designed by use of this approach. Not all the atoms of the diene
or dienophile need be carbon atoms, meaning that both carbocyclic and heterocyclic
rings can be constructed.
The Diels±Alder reaction is important because it is frequently used at an early stage
of a synthesis to establish a structural core which can then be elaborated to give the
more complex target structure. Classical Diels±Alder reactions have been greatly im-
proved by the use of microwaves ± yields have been increased and reaction times re-
duced. For instance, dramatic improvements have been seen in reactions involving
trans,trans-1,4-diphenylbutadiene (22) [12, 31], and in the reaction of cyclopentadiene
(23) under pressure and solvent-free conditions [44] or, in the latter circumstances, by
using graphite [45] as a support for the tandem retro Diels±Alder/Diels±Alder reaction
starting from dicyclopentadiene. Similarly, anthracene (1) has been reacted under pres-
sure under both solvent-free conditions [12, 31] and in solution in 1,2-xylene [14]. Fi-
nally, cyclohexadiene (24) and a-terpinene (25) have been reacted with diethyl acetylene-
dicarboxylate (26a) [46]. Some Diels±Alder adducts undergo a retro Diels±Alder reaction
to provide, by loss of ethylene and aromatization, substituted benzenes (Scheme 9.5).
Giguere performed tandem ene-intramolecular Diels±Alder reactions between 1,4-
cyclohexadiene (29) and dimethyl acetylenedicarboxylate (26b) in sealed tubes in a
commercial microwave oven (Scheme 9.6) [47].
Berlan found that cycloaddition reactions performed under reflux in xylene or di-
butyl ether (Scheme 9.7) were always faster under microwave conditions than when
9.5 [4+2] Cycloadditions 303
R' R'
CO2Et R'
MW CO2 Et
CO2 Et
+
R CO2 Et
CO2Et CO 2Et
R R
24, 25 26a 27 28
Scheme 9.5
O O
MW, 95 ºC
+
Xylene or
Dibutyl ether
Scheme 9.7 31 32 33
using classical heating methods [48]. The observed acceleration is more significant
in apolar solvents, for which dielectric losses are weak. On this basis the authors pro-
pose that G{ is changed, possibly as a result of a change in the entropy of the system.
Subsequently, Strauss indicated that the kinetics of these and other reactions are si-
milar under the action of microwave irradiation and classical heating ± hence there
is no specific effect [14].
Similarly, in the cycloaddition of cyclopentadiene (23) with methyl acrylate (34),
described by Gedye, microwave radiation does not alter the endo/exo selectivity and
the observed changes can be explained on the basis of the reactions under micro-
wave conditions occurring at higher temperatures than those occurring under reflux
and under pressure (Scheme 9.8) [49].
Reaction of pyrones under classical conditions requires the use of high tempera-
tures to obtain low to moderate yields. The Diels±Alder reaction of pyrones has been
performed in a commercial microwave oven under solvent-free conditions on solid
supports such as silica gel, montmorillonite, fitrol clay and alumina. The reaction
time was dramatically reduced ± from 4 h to 4 min (Scheme 9.9) [50].
304 9 Microwaves in Cycloadditions
R1 O O
MW, SiO2 R1
+
R2 O O 4 min
O R2 O
36 37 38 (73%)
Scheme 9.9
O
MW, 5 min O
O
O OH 1) SiO2 /H 2O (1:1)
2) TBSCl, DMF TBSO
40 41 (64 %)
Scheme 9.10
Cycloaddition of furan (5) has again been performed successfully under pressure
and solvent-free conditions [12, 44]. Usually, however, the cycloaddition of furan and
heterocyclic compounds requires a Lewis-acid catalyst to give good yields.
Reactions of furan (5) under solvent-free conditions, catalyzed by Montmorillonite
K10, have been described by Cintas [27]. The reaction with methyl vinyl ketone (32)
produced Michael addition in positions 2 and 5, whereas reaction with symmetrically
substituted cyclic dienophiles produced a mixture of the endo and exo adducts with
the kinetically favored endo adduct predominating, except when maleic anhydride
(39) was used as the dienophile (Scheme 9.2).
Although the intramolecular Diels±Alder reaction of furan 40 does not occur with
classical heating (Scheme 9.10) [51], the reaction has been performed successfully in
64 % yield by using microwaves and absorbing the product on to silica gel±water.
Moreno described the cycloaddition of 2,5-dimethylfuran (42) catalyzed by silica-
supported Lewis acids under solvent-free conditions in closed Teflon vessels using a
commercial microwave oven (Scheme 9.11) [28, 52]. Under these conditions coordi-
nation of the silica-supported catalyst with the oxygen bridge favors ring opening,
thus leading to the aromatic compounds in one step. The use of Si(Ti) gave the best
results for aromatic compounds.
Most of these compounds gave very low yields when the reactions were performed
using classical heating in an oil bath under comparable reaction conditions and, in
9.5 [4+2] Cycloadditions 305
CH 3 O CH 3
CN CH 3 O
MW CH3 CN
M(Si)
O + +
M(Si) CN
H3 C CN H 3C
CH 3 CH 3
Scheme 9.11 42 43 44 45 46 (50-92%)
NMe NMe
NMe
MW
+
COMe MeO MeO
MeO OH OH
OH H COMe
H COMe
32
48 49 50
fact, mixtures of the cycloadducts and aromatic compounds were often obtained.
When the same reactions were performed in a monomode reactor the cycloadducts
could be isolated relatively easily.
Computational results showed that activation barriers for the aromatization pro-
cess are always lower for aluminum-catalyzed reactions and, similarly, lower for tita-
nium-catalyzed than for zinc-catalyzed reactions [52].
Surprisingly, the aromatic product is not obtained in the reaction with fumaroni-
trile (47) ± the Diels±Alder cycloadduct is the only product. The activation barriers
calculated for this aromatization were the highest reported in this work [52].
In an attempt to prepare new diprenorphine analogs, Linders, in one of the first
examples of microwave-induced organic reactions, reported the reaction between
methyl vinyl ketone and 6-demethoxy-b-dihydrothebaine (48) [53]. The Diels±Alder
reaction, when performed under classical conditions, led to extensive polymerization
of the dienophile. A dramatic improvement was achieved when the cycloaddition
was conducted in a modified microwave oven at the reflux temperature of methyl vi-
nyl ketone. By use of these conditions adducts 49 and 50 were obtained in a 3 : 2 ra-
tio, according to HPLC (Scheme 9.12).
The same authors showed that in the cycloaddition of northebaine derivatives the
reaction time with microwave heating was considerably shorter than when conven-
tional heating was used [54]. When, however, cycloadditions were performed in ben-
zene, a poor absorber of microwave irradiation, microwave heating did not result in
a significant increase in the rate of the reaction.
Fallis, in the synthesis of longifolene (53), a bridged sesquiterpene, performed the
intramolecular cycloaddition of compound 51 as a key reaction in the construction
of the bridged system (Scheme 9.13) [55].
The cycloaddition afforded 10 % of cycloadduct 52 after 24 h under reflux in to-
luene in a modified household oven, whereas decomposition predominated at higher
temperatures. When the triene was heated in a sealed glass vessel in a modified mi-
crowave oven the adduct was obtained in 92 % yield.
306 9 Microwaves in Cycloadditions
OR MW
CO2 Me
Toluene
51
52 53
t
R = SiMe2 Bu
52 Conventional heating, 10%
Microwave irradiation, 92%
Scheme 9.13
O
O MW A
H B
MOMO Toluene, reflux, 10 h H C
MOMO
54 55 (40%)
Scheme 9.14
The same author used this strategy to synthesize tricyclic taxoid skeletons 55, an
intramolecular Diels±Alder approach that proceeds in the direction from left to right
(ring A to BC) (Scheme 9.14) [56]. Once again, microwave irradiation was necessary
to perform the required cycloaddition in good yield.
The cycloaddition of ketone 54 could be effected in a sealed glass tube in a modi-
fied microwave oven to afford the tricyclic system stereoselectively. This major ad-
duct arose via the preferred transition state, in which the nonbonded interactions
were minimized, because of the alignment of the dienophile beneath the triene unit
furthest from the MOM substituent. This pattern of p-facial selectivity implies that,
with the ªnaturalº C2 stereoselectivity, the preferred geometry should provide the re-
lative stereochemistry required for taxol itself.
The synthesis of the decalin unit of compactin (59), a potent competitive inhibitor
of 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase, which acts as an
effective hypocholesterolemic agent, was planned to incorporate an intramolecular
Diels±Alder reaction (Scheme 9.15) [57].
The reaction was performed by heating the reagents at 150 8C for 10 h in o-xylene.
The desired exo adduct was obtained with high stereoselectivity and subsequent for-
mation of the diastereoisomer 58 was considered to occur by isomerization of the
exo adduct. The reaction was dramatically accelerated by use of microwave irradia-
tion in a domestic microwave oven, and the carboxylic acid was obtained after
45 min with a small increase in stereoselectivity (Scheme 9.15) [57].
The Diels±Alder reaction between azonianthracene (61) and bis-substituted olefins
to give 6,11-ethanobenzo-[b]-quinolizinium salts 62 was greatly accelerated by the ac-
tion of microwave irradiation and, consequently, the reaction might be applicable to
9.5 [4+2] Cycloadditions 307
O R1 O OH
O O
2 O O
O R O
R1 O
R2
MW R2
R1 +
O
Toluene H
56 58
57
57 R1 = CO2 Et, R 2 = H; Conventional heating: exo/endo: 91/9
Microwave irradiation: exo/endo: 94/6 Compactin 59
58 R1 = CO2 H, R2 = H; Conventional heating: exo/endo: 82/18
Microwave irradiation: exo/endo: 85/15
Scheme 9.15
R1
S
N MW, 500 W, 3 min
+ N S R2
CH 2CH 2F
S S TFE, 10% AcOH
Br Br
60 61 62
a R1 = H; R2 = CH 2CH 2F
Scheme 9.16 b R1 = CH 2CH2 F; R2 = H
labeling with 18F, which has a short half-life (Scheme 9.16) [58]. The most efficient
reaction occurred in 10 % AcOH in TFE in a commercial microwave oven and the re-
action was approximately 200 times faster than that in CH3CN. In the synthesis of
62a, the radiochemical yield was 20 % in a 1 : 1 ratio after a synthesis time of 77 min,
including HPLC purification.
Deshayes described the cycloaddition of 1-amino-1,3-butadienes 63 with ethyl ac-
rylate (34) under the action of microwave irradiation in a monomode reactor with
temperature control [59]. Irradiation for 30 min at 70 8C in the absence of solvent af-
forded a 60 : 40 ratio of an inseparable mixture of the endo and exo isomers in 90 %
yield. Classical heating under the same conditions did not affect the selectivity but
the yield was lower (Scheme 9.17).
Lacova described the cycloaddition of vinylchromones with a variety of dienophiles,
by use of both classical heating and microwave irradiation. Only reaction with maleic
MW +
+ CO 2Et CO 2Et
CO2 Et 30 min, 70ºC
NR2 NR2 NR2
Scheme 9.17 63 34 64 (60%) 65 (40%)
308 9 Microwaves in Cycloadditions
anhydride (39) was successful under classical conditions [60]. The use of microwave
irradiation for the Diels±Alder experiments was unsuccessful in this instance. Silva
and de la Hoz have, however, performed the cycloaddition of styrylchromones 66
with N-methyl- and N-phenylmaleimide (39) and (6), respectively. The reaction per-
formed with conventional heating resulted in cycloaddition then aromatization,
whereas the same reaction under the action of microwaves gave the cycloadduct 68,
which could be isolated without migration of the double bond (Scheme 9.18) [61].
Diels±Alder cycloadditions of vinylpyrazoles under classical conditions require
highly reactive dienophiles and extreme conditions, i. e. high pressures and tempera-
tures (8±10 atm and 120±140 8C) for long reaction times (several days), and usually
afford moderate yields only [62]. The main obstacle to these reactions is extensive
polymerization of the reagents. 1-Phenyl-4-vinylpyrazole (69) reacted with dimethyl
acetylenedicarboxylate (26b) within 6 min under the action of microwave irradiation
to afford the adducts 70 and 71 in 72 % overall yield (Scheme 9.19) [63]. The cycload-
Ph
O
N
O
O O
Cl CH
O Cl
66 O
67
+
MW Ph
O
O N
270 W, 30 min O O
N Ph
O Cl
6 O
Scheme 9.18 68
CO2 CH 3
H3 CO2 C
DMAD CO2 CH 3
N N CO2 CH 3
MW N + N
N CO 2CH3 N
Ph Ph CO 2CH 3
Ph
69 70 (10%) 71 (62%)
H3 CO 2C
H 3CO 2C
CO 2CH3
N
N
CO 2CH 3
Ph
Scheme 9.19 72
9.5 [4+2] Cycloadditions 309
dition was performed in a Teflon vessel in a domestic oven. The clean nature of the
reaction enabled characterization of intermediate 72, although this compound is not
detected when classical heating is used, because it rearranges spontaneously in a
few minutes at the reaction temperature to give the aromatic compound.
The reaction proceeds even with a very poor dienophile such as ethyl phenylpro-
piolate (73), which does not react under classical conditions.
C60 (4) reacts as an electron-deficient polyolefin and, consequently, one of the
main ways of functionalizing fullerenes involves cycloaddition reactions [64] in
which C60 (4) is a reactive 2 component. In this context, [1+2], [2+2], [3+2], and [4+2]
cycloadditions have been performed (Scheme 9.20) and the conditions for cycload-
duct formation found to depend strongly on the gap between the controlling orbitals.
For this reason it is frequently necessary to use conditions involving several hours
(or days) under reflux in high boiling solvents. It thus seemed interesting to investi-
gate the potential of microwave irradiation in the preparation of fullerene derivatives
when this type of reaction is involved. The characteristics of [60]fullerene (4), i. e. the
absence of a dipole moment and the need to work on a very small scale, imply the
use of a solvent in these reactions.
[4+2] Cycloadditions selectively afford the adducts on the 6,6-ring junctions [65], and
the products occasionally undergo a facile retro-Diels±Alder reaction as a consequence
of the low thermodynamic stability of the adduct. Very stable Diels±Alder cycloadducts
have, however, been prepared by use of different substituted o-quinodimethanes, prob-
ably because of stabilization by aromatization of the resulting adducts [66].
ER2
R
R
R
R
[1+2] [2+2]
[4+2] 4 [3+2]
R a
b
R c
Scheme 9.20
310 9 Microwaves in Cycloadditions
MW C60
O
S 4
O
74
MW, 15 min
C60 +
Toluene
4 1
R2 NH2 O R2 N NaI
Br Br
+
Br Br
R1 NH2 O R1 N 18-Crown-6
77 78 79a-c
R2 N C60, ODCB N R2
R1 N MW
N R1
CH
R1 = R2 = a: CN, b: (-CH=CH) 2, c:
80a-c
CH
Scheme 9.23
radiation in a focused microwave reactor. Once again, the use of microwave irradia-
tion led to higher yields than classical heating for 80a and 80b (up to 4.5-fold) and re-
action times were significantly reduced ± from 24 h to 0.5 h. For 80c, large amounts
of polyadducts were detected when microwave irradiation was used; this reduced the
yield.
Interestingly, in syntheses of thiophene cycloadducts 81 use of microwave irradia-
tion led to lower yields [73] than conventional heating (23 % compared with 43 %),
although reaction times were significantly reduced [74].
9.5.2
Retro Diels±Alder Reactions
Retro Diels±Alder reactions often require severe conditions, high temperatures, and,
sometimes, flash vacuum thermolysis. Microwave irradiation has been used as an al-
ternative to these harsh conditions [12], even using graphite as a support [45]. Simi-
larly, the use of higher input power has enabled retro-Diels±Alder reactions of an-
thracene derivatives to occur in 3±5 min [41]. This method is an alternative to the
use of flash thermolysis. The use of graphite is a prerequisite for obtaining high tem-
peratures in a short time.
Bortolussi described the synthesis of unsaturated amino alcohols 83 by thermoly-
sis of furan adducts 82; use of microwaves resulted in a significant reduction in reac-
tion time (Scheme 9.24) [75].
R
O R' N CH2 Ph H
MW
R' CH2 OH
H +
10-15 min R N CH 2Ph O
CH2 OH
Scheme 9.24 82 83 (98%) 5
312 9 Microwaves in Cycloadditions
9.5.3
Hetero Diels±Alder Reactions
9.5.3.1 Heterodienes
The introduction of a nitrogen atom into a conjugated system will confer electrophi-
lic character on the system. This has led to the development of the inverse-electron-
demand Diels±Alder reaction [1 b, 10]. Substitution of the azadiene with electron-
withdrawing groups accentuates the electron-deficient nature of the azadiene and
enables the use of electron-rich dienophiles. The magnitude of the LUMOdiene±
HOMOdienophile energy separation has been related to the rate of inverse-demand cy-
cloadditions. Alternatively, addition of a strong electron-donating substituent to the
azadiene enables the use of conventional electron-deficient dienophiles in a normal
demand Diels±Alder reaction.
The Diels±Alder reaction of 1-azadienes suffers from low conversion, because of
the low reactivity of these systems. In an attempt to overcome this problem, consider-
able effort has been focused on both inter- and intramolecular versions of the reaction.
Garrigues described the reaction of the 1-azadiene 84 with dimethyl acetylenedi-
carboxylate (26b) on a graphite support [35]. After sequential microwave irradiation
for 10 min in a monomode reactor pyridine 85 was obtained in 60 % yield. This aza-
diene does not react when conventional heating is used (Scheme 9.25).
Similarly, Motorina described the intramolecular hetero-Diels±Alder reaction of N-
alkyl-2-cyano-1-azadienes 86 (Scheme 9.26) [76]; although the reaction had already
been performed by heating the reactants in benzene overnight in a sealed tube at
110 8C, it was found that reaction time could be reduced to 14 min in a microwave oven
at 650 W.
COOMe
Me MW 10 min. Me COOMe
C
+
C Graphite N COOMe
N
COOMe
NMe2
84 26b 85 (60%)
Scheme 9.25
Me Me
NC N ∆ or MW NC N
H
86 87
Scheme 9.26
9.5 [4+2] Cycloadditions 313
NH
CO 2Et
N
H
90 NH
R R
R CO2 Et
NH N
CO2 Et 89 R
N
H R
H
88
91 (67-98%)
NH
CO 2Et
N
R
Scheme 9.27
O
H O
H
HO
OH
O N
N
O
O 93 O
N OH
39 N
N O O
N
O MW, 466 W, 9 min O H
OH H O
OH O 95
92 OH
N
N
Scheme 9.28 94 O
314 9 Microwaves in Cycloadditions
Ph Ph
NO 2
NO 2 + N
Ph 97 N N N
N N
MW, 5 min Et Et
130 ºC
N NMe2 99 (64%) 100 (8%)
N
N
Et MW, 5 min
96 130 ºC
S
S NO 2 NO2
98
N N
N
Et
Scheme 9.29 101 (84%)
O O O
H 3C O 60 W, 3 min H3 C
N + N Ar
Ar MW
H3 CS N N N(CH3 )2 H3 CS N N
N 33-71%
(CH2 )n
106
OH MW C60
OH O 4
O
110
ines. This methodology is a dramatic improvement on the classical method and en-
ables such heterocyclic derivatives to be obtained in good yield (Scheme 9.31).
The reaction of C60 with o-quinonemethide, prepared from o-hydroxybenzyl alco-
hol (110) (Scheme 9.32), was performed in a modified domestic oven at 800 W and
gave 111 in 27 % yield after only 4 min [72]. Although Eguchi et al. [83] reported the
same reaction with a slightly better yield (31%) by thermolysis in a sealed vessel, the
microwave approach to this adduct has the advantage of simplicity, and avoids the
risks associated with high-pressure conditions.
9.5.3.2 Heterodienophiles
Examples of the use of heterodienophiles under the action of microwave irradiation
are not common. Soufiaoui [84] and Garrigues [37] used carbonyl compounds as die-
nophiles. The first example employed solvent-free conditions; the second is an exam-
ple of the use of graphite as a susceptor. Cycloaddition of a carbonyl compound pro-
vided a 5,6-dihydro-2H-pyran derivative. These types of reaction proceed poorly with
aliphatic and aromatic aldehydes and ketones unless highly reactive dienes and/or
Lewis acid catalysts are used. Reaction of 2,3-dimethyl-1,3-butadiene (31) with ethyl
glyoxylate (112) occurred in 75 % yield in 20 min under the action of microwave irra-
diation. When conventional heating is used it is necessary to heat the mixture at
150 8C for 4 h in a sealed tube to obtain a satisfactory yield (Scheme 9.33).
Jimnez et al. studied the asymmetric Diels±Alder reactions of 1-aryl-1,2-diaza-1,3-
butadienes 114, heterodienes derived from sugars, with diethyl azodicarboxylate
(115), a heterodienophile [85]. The reactions were performed without solvent in a fo-
cused microwave reactor for periods of a few hours. The reaction is stereoselective
316 9 Microwaves in Cycloadditions
H3C H 3C
O MW, 30 W O
+ CO2 Et
H3C H CO2 Et 20 x 1 min H 3C H
31 112 113
Scheme 9.33
Ar Ar
N CO 2Et N CO 2Et
N N [R(1)ReR(4)Re] N N
+
N N
EtO2 C AcO CO 2Et
AcO H HH
R*
R*
114 115 116
Scheme 9.34
and the major diastereoisomers have the (6R)-configuration if the chiral descriptor
of the heterodiene at C-1 is R. The relative induction arises from the preferential re-
action of DEAD with the Re face of the azoalkane (Scheme 9.34). This results in a
like (lk) combination and the stereochemical course can be designated self-consis-
tently as [R(1)ReR(4)Re].
9.5.4
Miscellaneous
Hong described the [6+4] cycloaddition of fulvenes 117 with a-pyrones 118 to give
azulene-indoles 119, which have potential antineoplastic activity [86 a]. Reactions
were performed in DMF solution by use of a monomode reactor (30 W for 60 min).
It is worthy of note that reactions performed using conventional heating require
long reaction times (5 days) and give low yields of the final product, along with re-
covered starting materials and decomposition products (Scheme 9.35).
The same authors showed that microwave irradiation can alter the reaction path-
way. Occasionally with conventional heating the Diels±Alder adducts are favored
whereas the tandem [6+4]±[4+2] cycloaddition products were obtained exclusively un-
der the action of microwaves [86 b].
Me2 N H O
O MW, DMF
30 W, 60 min N
N H
H
117 118 119 (65%)
Scheme 9.35
9.6 1,3-Dipolar Cycloadditions 317
9.6
1,3-Dipolar Cycloadditions
1,3-Dipolar cycloadditions are one of the best and most widely used procedures to
prepare five-membered heterocycles. The process involves a concerted reaction be-
tween a 1,3-dipole and a multiple-bonded compound.
The 1,3-dipole is often very unstable, its formation requires high temperatures
and the subsequent cycloadditions require often long reaction times. Both of these
conditions result in a decrease in yields and purity of products. The rapid heating in-
duced by microwave irradiation avoids the harsh reaction conditions associated with
classical heating and facilitates 1,3-dipolar cycloadditions that are very difficult (or
impossible) to achieve with classical energy sources.
In the nineteen-nineties, numerous examples of 1,3-dipolar cycloadditions under
microwave heating were reported with a wide variety of dipoles. In many cases the
product yields and/or reaction times were improved and the method was used to pre-
pare valuable compounds that could not be obtained by classical heating and, in
some instances, the regioselectivity of the reaction was modified. Most of these pro-
cesses were performed in the absence of solvent.
In the following section the main 1,3-dipolar cycloadditions under microwaves
will be reviewed according to the nature of the dipole.
9.6.1
Azomethine Ylides
Me H
C N CH
EtO CO 2Me R N A OH
R
120 121 MW / 45-120 W Me N
9-18 min
+
N
HO A O
Me H
NH EtOH, MeOH
A 122 (88-95%)
EtO N CO2 Et
H R O
H
120'
121'
A = Alkyl chain
Scheme 9.36
318 9 Microwaves in Cycloadditions
Comparison of this reaction with that performed using classical heating shows an
acceleration of the 1,3-dipolar cycloaddition upon irradiation and yield enhance-
ments for the synthesis of imidazolones 122. These advantages occur by virtue of di-
minished formation of the byproduct 125 (see Scheme 9.37).
The same authors also investigated the reaction of the imidate 120 and aldehydes
123 with dry acetic acid as a catalyst in the absence of solvent. The mixture was imme-
diately submitted to focused microwave irradiation to afford compounds 125 together
with alcohols, ethyl acetate and a small amount of imine 124 (Scheme 9.37) [88].
Likewise, imidate 126 undergoes highly regioselective cycloadditions with a wide
range of aldehydes ± as dipolarophiles ± in short reaction times to afford ethyl 4-cy-
ano-2-oxazoline-4-carboxylates 127 in good yields (81±98 %) and moderate diastereo-
meric ratios (Scheme 9.38). The use of microwave radiation gave higher yields than
classical heating but the moderate diastereomeric composition was the same [89]. In
a similar way, addition of imidate 126 to the dipolarophile N-benzylidenemethyl-
Me - MeCO2 Et
C N CH 2CO2 Me + R CHO R N CO2 Me
EtO
120 123 124
Me H
C N CH
EtO CO2 Me R
120 Me N
124 / MW
N
MeO2 C O
Me H
EtOH, MeOH 125
EtO N CO2Et
H
120'
Scheme 9.37
Me N CN
Me CO2Et
C N CH O CO2 Et
EtO CN R
EtOH
126 H 127 cis
R CHO Me N CO 2Et
+
MW EtO CN
O
Me CO2Et R Me N CO2Et
EtO N CN O CN
H
R
126' 127 trans
Scheme 9.38
9.6 1,3-Dipolar Cycloadditions 319
Ph C C CO2Et
73 CO 2Et
N N N Ph
MW / 120 W +
C
NC CN 25 min NC NC
Ph CO2Et
130 131 132
celite 88 12
bentonite 98 2
Scheme 9.39
amine (128) under focused microwaves gave the expected product ethyl 4-cyano-
2-methyl-5-phenyl-2-imidazoline-4-carboxylate (129) in good yield.
Solid supports such as silica gel and alumina have also been used in the genera-
tion of 1,3-dipoles to obtain heterocyclic compounds. However, changes in regios-
electivity were not observed under microwave irradiation conditions. The regioselec-
tivity can be modified by changing the polarity of the solid support [90]. De la Hoz re-
ported the reaction of pyridinium dicyanomethylidene ylide (130) with ethyl phenyl-
propiolate (73), and found that the selectivity changes from 88 : 12 to 98 : 2 when the
support is changed from celite, to neutral bentonite (Scheme 9.39).
4,4-Bipyridinium ylides 133, generated in situ from 4,4-bipyridinium diquaternary
salts 133, undergo 1,3-dipolar cycloaddition with activated alkynes under micro-
waves, on KF±alumina in the absence of solvent, to give 7,7-bis-indolizines 134 in
81±93 % yield (Scheme 9.40) [91]. The same reactions, when performed using ben-
zene as a solvent under classical heating, yielded 7,7-bis-indolizine derivatives in
yields of only 50±60 % [92].
Microwave irradiation in solvent-free conditions induces the cleavage of the 2,3-
bond of 2-aroyl-aziridines 135 to give an azomethine ylide intermediate, which sub-
sequently undergoes cycloadditions to a multiple bond and leads to oxazolidine, imi-
dazoline, naphthooxazole and pyrroline derivatives 136 in good yields (Scheme 9.41)
[32 b]. Reactions were performed at atmospheric pressure in an Erlenmeyer flask
placed in a commercial domestic oven. The reactions were complete in 10±15 min
while the conventional method requires 18±20 h.
Highly stereoselective intramolecular cycloadditions of unsaturated N-substituted
azomethine ylides have been conducted under microwave irradiation. Oritani re-
ported that a mixture of the aldehyde 137 and N-methyl- or N-benzylglycine ethyl es-
ter (138) on the surface of silica gel, irradiated under microwaves for 15 min, gener-
ated azomethine ylides 139 that subsequently underwent in situ intramolecular cy-
cloadditions to afford the corresponding tricyclic compounds 140 in 79 and 81%
yield, respectively (Scheme 9.42) [93].
A general method for the functionalization of C60 (4) is the 1,3-dipolar cycloaddi-
tion of azomethine ylides. This process was first described by Prato [94] and leads to
fulleropyrrolidines. Several fulleropyrrolidines (143a±c) have been prepared under
microwave irradiation by Langa et al. [72]. These authors observed that microwave ir-
radiation again competes favorably with thermal heating and, in this way, 143 a was
320 9 Microwaves in Cycloadditions
R' Br Br R'
R C C N N C C R
H H2
O 2 O
133
MW KF / alumina
R' R'
R C C N N C C R
O H H O
133'
R'' C C COOR'''
73
R' O O R'
C N N C
R R
R'' R''
COOR''' R'''OOC
H R R
CHO RNHCH2 COOEt N H H
138 CH 2COOEt COOEt
O N O N O
MW / 15 min N H
O O O
137 139a R = Me 140a (79%)
139b R = Bn 140b (81%)
Scheme 9.42
CHO
CO2 H R C60
+
N NH
NH2
R
tively (Scheme 9.44) [43]. Under conventional heating the 7±21 isomer (145c) was
formed in only a low proportion and the 1±2 isomer (145a) was found to predomi-
nate. Under microwave irradiation, and on using o-dichlorobenzene (ODCB), a polar
a solvent that absorbs microwaves efficiently, significant changes were observed. In
contrast to classical conditions, 145c was not formed under microwave irradiation re-
gardless of the irradiation power, and isomer 145b was found to be predominant at
higher power (Scheme 9.44 and Fig. 9.1).
A computational study on the mode of cycloaddition showed that the reaction is
stepwise, the first step consisting of a nucleophilic attack on the azomethine ylide.
The most negative charge of the fullerene moiety in the transition states leading to
145 a and 145b are located at the carbon adjacent to the carbon±carbon bond being
formed. In the transition state that affords 145c, however, the negative charge is delo-
calized all around the C70 subunit. The relative ratio of isomers 145a±c is related to
the greatest hardness, and the formation of the product along the harder transition
state should be favored under microwave irradiation. It is noteworthy that purely
Bond Length
b
a 1-2 a: 1.375 Å
c 5-6 b: 1.387 Å
d 7-21 c: 1.414 Å
20-21 d: 1.465 Å
144
CH 3NHCH 2COOH /HCHO
C70 N CH3
C70
∆ or Microwaves
144
a) b) c)
Fig. 9.1 1H NMR spectra (methyl groups) of cal heating in ODCB; (c) microwave irradiation
the adducts 145a, 145b, and 145c from left to at 180 W in ODCB.
right. (a) classical heating in toluene; (b) classi-
R3
O
O O N O
O
R1 MW
OCH 3 + + N R3
R2 H DCE MeO2C
NH2 R2
O 7 min R1 N
H
138 123 6
146
Scheme 9.45
9.6 1,3-Dipolar Cycloadditions 323
O O
R3
O + N R1 + COOH
N N
R2 O H
147 6 148
MW
O
N N
R3 O R3 H
stereoselective
O O
cycloreversion
N N
R2 R2
149 150
1,3-dipolar
cycloaddition
HH O
N H N R1
R3
O O
N
R2
Scheme 9.46 151
phile, to produce the adduct stereoselectively [97]. It is noteworthy that the adduct
has five chiral centers, but the synthesis affords only one diastereoisomer owing to
the dipole configuration and exo-transition state structure.
9.6.2
Nitrones
Ph
F3 C
Ph O Ph MW / 30 W
N + O Ph
Me F3C 3 min N
Me
152 153
154
Scheme 9.47
Me
pF3 C-C6H 4 N
pF3C-C 6H4 O 135 ºC O
EtO2 C CN + N
3 min N
Me CO 2Et
155 152
microwaves 50% 156
Scheme 9.48 classical heating 17%
CH 2
Ph MW
Ph O
N R + N + N Ph
Ph O O N O N O N
O R Ph R Ph
157 11
Scheme 9.49 158 159
9.6 1,3-Dipolar Cycloadditions 325
OH O H CO 2Et O H CO2 Et
O O
O N Bn + O N Bn
O O
H H
160 HO HO
MW 162 (64%) 163 (23%)
+
dioxane
H CO2 Et 10 min O H O H CO2 Et
N O
Bn O O N Bn + O N Bn
O
161 H CO2 Et H
HO HO
Scheme 9.50 164 (10%) 165 (3%)
small change in the stereoselectivity (in favor of the exo-162 isomer). Moreover, the
use of microwaves also led to the formation of a small amount of the unexpected syn
adduct 165 [101].
When the aldehyde 137 and N-methylhydroxylamine (166) were mixed with silica
gel and irradiated with microwaves for 15 min the adduct obtained was 168 (82 %
yield). A mixture of 137 and N-methylhydroxylamine (166) in alcohol, heated at
120 8C in a sealed tube, afforded the products 168 and 169 in an 8 : 1 ratio
(Scheme 9.51) [93]. The authors proposed that the selectivity observed in the micro-
wave-induced reaction results from a faster transformation leading to the kinetically
controlled product.
In a recent study, classical heating, microwave radiation and gamma radiation
have been compared as energy sources to perform 1,3-dipolar reactions between un-
saturated oximes and conventional dipolarophiles. On using gamma radiation the
reactions were clean and yields obtained were similar to those for the thermal reac-
tions. However, microwave radiation reactions were extremely clean, occurred more
rapidly and gave higher yields [102].
Me Me
H CH3 H
H N N
CHO MeNHOH·HCl C N O
166 O O
O + O
O N O N N N
H H
O O O
O
137 167 168 169
Microwaves / silica gel - 15 min 100 0
Scheme 9.51 NaHCO 3 / EtOH 90% reflux 88.9 11.1
326 9 Microwaves in Cycloadditions
9.6.3
Nitrile Oxides
Microwave irradiation has been extensively employed to generate nitrile oxides and
to promote 1,3-dipolar cycloadditions of the previously prepared dipole.
The first application of microwave irradiation in conjunction with dry media in
the generation of nitrile oxide intermediates was reported by Hamelin [29]. In this
example, methyl nitroacetate (170) was mixed with a dipolarophile in the presence
of catalytic amounts of toluene-p-sulfonic acid (PTSA) (10 % weight). Subsequent mi-
crowave irradiation led to the formation of the corresponding heterocyclic adducts
(Scheme 9.52). Reactions were performed in an open vessel from which water was
continuously removed [103]. Likewise, irradiation in a domestic oven of a mixture of
ethyl chloro(hydroxyimino)acetate (173) and a dipolarophile over alumina led to the
same results in only a few minutes (Scheme 9.52) [103].
In 1999 Hamelin and Benhaoua reported a comparative study between dry media
conditions and the procedure in homogeneous solution under microwaves and clas-
sical heating [104]. Reactions undertaken with aromatic oxime chlorides 174 over
alumina, without solvent and under microwaves, showed that alkynes and disubsti-
tuted alkenes give moderate yields of cycloadducts (40±60 %). The corresponding
procedure with trisubstituted alkenes did not give any products. These reactions
were also performed in the presence of N-methylmorpholine (NMM) in xylene at
130 8C during 30 min under focused microwaves. The same products were obtained
in these cases and the yields were better than for the reactions in dry media. The use
of classical heating resulted in worse yields in all cases (Scheme 9.53).
Soufiaoui reported that arylnitrile oxides 10 can be generated under microwaves
not only from aromatic oxime chlorides, such as 174, by the action of alumina
(Method A) but also from aryloximes, such as 176, by the addition of a chlorination
agent, N-chlorosuccinimide (NCS), supported on alumina (Method B) (Scheme 9.54)
[29 a]. Both methods afforded identical cycloadducts in similar yields ± when indene
was used as the dipolarophile the yield was 86 %. In the absence of alumina, meth-
od B fails (reagents are less reactive and decompose) and method A does not yield
any cycloadduct but a dimer of the dipole.
MeO2 C CO 2Me
PTSA / MW DMAD
MeO2 C CH 2NO2 MeO2 C C N O O
30 min MeO2C N
- H2 O
170 171 172 (91%)
MeO 2C CO 2Me
DMAD
Cl alumina / MW O
C N OH EtO2 C C N O 30 min EtO2 C N
EtO2 C - HCl
173 171 172 (56%)
Scheme 9.52
9.6 1,3-Dipolar Cycloadditions 327
E E
EHC CHE
Ph C N OH Ph C N O O
Cl Ph N
10
174 175
alumina / neat / MW / 30 min 60%
NMM / xylene / MW / 130 ºC, 30 min 65%
Scheme 9.53 classical heating 30%
Cl H
Cl C N OH Cl C N OH
174 176
Cl C N O
10
177
Cl
O N
+
N
O
Cl
Cl
C3 H7 -(E) (E)-C 3H 7
Ph NOH Ph
174 OH
(E)-C3H 7 OH + O
NEt3, benzene N O N Ph
OH
180 181 182
classical method:
57 : 43
50 h, r. t. (54% yield)
microwave irradiation:
46 : 54
Scheme 9.55 400 W, 2 min (53% yield)
O OCOAr
N
ArOC COAr
N O C Ar C
MW X Y ArOC
Ar C C C N O X
N N 12-18 min
O O O N Y
O
O 2N C N Me Me
Me
186 N
O
Me C N O
MW / 450 W / 8 min Me N
Me
NO 2
10
187 (68%)
Scheme 9.57 classical heating, 48% yield
9.6 1,3-Dipolar Cycloadditions 329
Ar C N O
10
H
H O
1 N
Ar
188
Ar C N O
10
Ar
N O N
H H Ar
O
H H
H + H
O H O
H
N N
Ar Ar
189 190
Products (yield, %)
Classical heating with solvent
188 (15.7) + 189 (traces) + 190 (4.6)
24 h in refluxing toluene
Microwave irradiation
188 (11) + 189 (10.2) + 190 (32.7)
650 W, 6 min
Scheme 9.58
case with classical heating and, moreover, the symmetrical bis-cycloadduct 189 was
also isolated. Reaction times were shortened from 24 h to 3±10 min and yields of cy-
cloadducts increased by a factor of 3±7 on changing from heating at reflux to micro-
wave irradiation. The highest yields of bis-cycloadducts suggested that these could
be because of the effect of the lowest temperature used in classical heating. An at-
tempt to perform the reaction under reflux in decalin failed, however, because the
1,3-dipole underwent degradation rather than cycloaddition.
Methylmagnesium bromide (191) exerts a great influence on the stereoselectivity
of the reactions between mesitonitrile oxide 10 and the Baylis±Hillman adducts 192.
In the absence of a Grignard reagent, a mixture of isomers is formed in which com-
pounds 194 are the main products. The presence of a Grignard reagent reverses the
stereoselectivity (Scheme 9.59). When FiÉera [107] performed these reactions under
microwave irradiation, the reaction times decreased from days to less than 5 min
without any loss of stereoselectivity for noncatalyzed cycloadditions, but with a small
change in the stereoselectivity in the chelated reactions.
Nitrile oxides have been used in conjunction with microwaves in the field of fuller-
ene chemistry. For example, 3-(N-phenylpyrazolyl)isoxazolo[60]fullerene dyads 198
330 9 Microwaves in Cycloadditions
Scheme 9.59
N OH Br
N OH
NBS / Benzene
N
N N
Ph N
Ph
195 196
Et 3N
N
N O
N
C60, MW N O
Ph N
10 min, 210 W N
Ph
a N e
N S
N Ph
R
O
N
b N f Me 2 N
Ph
c g Fe
Me
198
d N h
O
Scheme 9.61
9.6.4
Nitrile Imines
Ph N Ph
C N
Ph C N N Ph
Cl H
199 200
177
Ph Ph
N N N
+
N
Ph Ph
201 202
Scheme 9.62
N Ar'
N Ar Ph N
Ar' N Ph Et3 N / MW N
N Ar
O +
200 W / 5 min O
N Cl H N
H H
203 199 204 (85-95%)
Scheme 9.63
nitrile imines are generated and react with several dipolarophiles. These reactions
appear better in homogeneous solution in xylene in the presence of N-methylmor-
pholine (NMM) under focused microwaves than in dry media [104]. In any case,
both methods afford better yields of cycloadducts than the corresponding reactions
under classical heating.
The formation of several pyrazolylpyrazolino[60]fullerene adducts (208a±c) from
nitrile imines have been described (Scheme 9.64) [111]. The nitrile imines are gener-
ated in situ from the corresponding hydrazone 205 and NBS in the presence of Et3N
9.6 1,3-Dipolar Cycloadditions 333
Br
NH NBS / Benzene NH
N N N
N
N N
R R
Ph Ph
205 206
N 1. Et 3N
N Ph 2. C60, MW
N
+ -
N N N
N
N
Ph R
R
208a-c 207
a: R=H, b: R=OMe, c: R=NO2
Scheme 9.64
and reacted with C60 (4) under microwave irradiation. This approach has also been
used to obtain a 2-pyrazoline ring fused to C60 in a one-pot process from the hydra-
zone. The route is simpler than the previously described method, which involved cy-
cloaddition of C60 to nitrile imines prepared from the corresponding N-chlorobenzy-
lidene derivatives [112].
This route has been used to prepare fullerodendrimers in 31±34 % yield [113].
9.6.5
Azides and Azomethine Imines
Other 1,3-dipoles, such are azides and azomethine imines, have also been employed
in microwave-induced cycloadditions. The main results reported are reviewed in this
section.
1,2,3-Triazole derivatives are very interesting compounds that can be prepared by
1,3-dipolar cycloadditions between azides and alkynes. Loupy and Palacios reported
that electron-deficient acetylenes react with azidoethylphosphonate 209 to form the
regioisomeric substituted 1,2,3-triazoles 210 and 211 under microwaves in solvent-
free conditions (Scheme 9.65) [114]. This procedure avoids the harsh reaction condi-
tions associated with thermal cycloadditions (toluene under reflux) and the very long
reaction times.
334 9 Microwaves in Cycloadditions
(EtO)2OP N3
N N
209 (EtO)2OP N N (EtO)2OP N N
+
+ H CO2Et EtO2C H
H C C CO2Et
210 211
89
Scheme 9.65
Aryl and vinyl nitriles have been prepared very efficiently from the corresponding
bromides by palladium-catalyzed reactions under microwaves. This energy source
has been employed for the conversion of these nitriles into aryl and vinyl tetrazoles
by cycloaddition reactions with sodium azide (Scheme 9.66). The direct transforma-
tion of aryl halides to the aryl tetrazoles in a one pot procedure could be accom-
plished both in solution and on a solid support [115]. The reactions were complete in
a few minutes, a reaction time considerably shorter than those previously reported
for the thermal reactions. The cycloadditions were performed with sodium azide and
ammonium chloride in DMF and, although no explosion occurred in the develop-
ment of this work, the authors point out the necessity of taking adequate precautions
against this eventuality.
The thermal hydrazone±azomethine imine isomerization can be easily performed
under microwave irradiation in the absence of solvent. The subsequent 1,3-dipolar
cycloadditions with electron-deficient dipolarophiles occur in only a few minutes to
afford the corresponding cycloadducts. The use of pyrazolyl hydrazones 205 leads to
valuable compounds, such as bipyrazoles 213, in good yields and this provides a new
approach to the preparation of these heterocyclic derivatives [116] (Scheme 9.67). Re-
actions undertaken with classical heating under comparable reaction conditions
(time and temperature) lead to cycloadduct yields that are considerably lower and, in-
deed, several dipolarophiles do not react at all.
HN N
CN
NaN 3, NH 4Cl N
N
H3 CO DMF / MW
25 min / 20 W H3 CO
186
212 (96%)
Scheme 9.66 thermal heating, 24 h, 68% yield
9.7 [2+2] Cycloadditions 335
R R'
H R C C R'
N MW N N Ar
N N Ar N N Ar 26 N
N N H N
N
Ph Ph 10-45 min
Ph
205 205' 213 (30-84%)
Scheme 9.67
9.7
[2+2] Cycloadditions
[2+2] Cycloadditions give rise to four-membered rings. Thermal concerted [2+2] cy-
cloadditions have to be antarafacial on one component and the geometrical and orbi-
tal constraints thus imposed ensure that this process is encountered only in special
circumstances. Most thermal [2+2] cycloadditions of alkenes take place by a stepwise
pathway involving diradical or zwitterionic intermediates [1 a]. Considerably fewer
studies have been performed regarding the application of microwave irradiation in
[2+2] cycloadditions than for other kinds of cycloaddition (vide supra). Such reactions
have been commonly used to obtain b-lactam derivatives by cycloaddition of ketenes
with imines [18±20, 117, 118].
In 1991 Bose described the synthesis of a-vinyl b-lactams by reaction of a,b-unsa-
turated acyl chlorides with a Schiff base in chlorobenzene under microwave irradia-
tion (an example of the ªeco-friendlyº MORE chemistry, in which only a limited
amount of solvent is used) [20 b]. Under these conditions, a-vinyl b-lactam formation
can be achieved in 65±70 % in approximately 5 min (classical conditions require sev-
eral hours and lead only to modest yields).
This methodology has also been used by Bose, who described the synthesis of the
thienamycin side chain [119], the first step of which was a [2+2] cycloaddition under
microwave irradiation. Likewise, Khajavi described the reaction of trichloroacetic an-
hydride with imines [120]; with classical heating the reaction requires the use of
Fe2(CO)9 as a catalyst, whereas under microwave irradiation a catalyst is not required.
Bose has described reactions between acid chlorides 214 and Schiff bases 215
where the stereoselectivity depends on the order of addition of the reagents
(Scheme 9.68) [117]. When the condensation was conducted by a ªnormal additionº
sequence (i. e. acid chloride last), only the cis b-lactam (216b) was formed. However,
R2 ZO R2 ZO R2
ZO NMM
+ +
N Chlorobenzene N N
1 1
O Cl R O R O R1
214 215 MWI 216a 216b
if the ªinverse additionº technique (triethylamine last) was used, 30 % cis and 70 %
trans b-lactams were obtained under the same conditions. When the reaction was
conducted in a microwave oven using chlorobenzene, the ratio of trans and cis b-lac-
tams was 90 : 10 irrespective of the order of addition. Moreover, there was no isomeri-
zation to the thermodynamically more stable trans b-lactam (216a). These data can
be rationalized by assuming that at higher reaction temperatures the rate of forma-
tion of the trans b-lactam is faster than that of the cis b-lactam [20].
This effect has recently been explained by considering that under microwave irra-
diation the route involving direct reaction between the acyl chloride and the imine
competes efficiently with the ketene±imine reaction pathway, a situation highlighted
by theoretical calculations (Scheme 9.69) [42 a].
A rapid approach to a-amino b-lactams has been developed by Bose and uses the
tetrachlorophthalimido group as a masked amino substituent (Scheme 9.70) [118].
The trans b-lactam 218a could be obtained almost exclusively in 98 % yield after 3±
5 min under strong microwave irradiation. Under classical conditions the cis isomer
(218a : 218b = 10 : 90) is obtained in 52 % yield.
N-(4-hydroxycyclohexyl)-3-mercapto-3-cyano-4-arylazetidine-2-ones were synthe-
sized from N-(4-hydroxycyclohexyl)aryldiimine by reaction with ethyl a-mercapto-a-
cyanoacetate on basic alumina under microwaves. The reaction time was reduced
from hours to minutes in comparison to conventional heating and, moreover, the
yield was improved [121].
R1 H R2 R1 R2 R1 R2
R1
. N N
N
O O R3 O R3 O R3
R2
Base
Base
N
R1 R2
R2
R1 R2 Cl Base R 1 Cl
Cl R1
R1 N
N
Cl N O R3 O R3
O O R 3
Scheme 9.69
R2 TCP-N R2 TCP-N R2
TCP-N NMM
+ +
N Chlorobenzene N N
O Cl R1 O R1 O R1
217 MW 218a 218b
215
R1 = R2 = p-methoxyphenyl
Scheme 9.70
9.7 [2+2] Cycloadditions 337
R1 R2
R1 R3 Z Y
Y Z
R3 N R1
R2
N N O
O R O R R3 R2
Scheme 9.71 219 220 221
O
O BnOCH2 COCl / NEt 3 H H
BnO HO
H HO 223 H
H NH O
NH O MW, 3 min O
Ph O
Ph O
224 (72%)
Scheme 9.72 222
Fe COX + R1HC=N-R 2 Fe
R1
225 215 N
O R2
Scheme 9.73 X= OH, Cl 226
338 9 Microwaves in Cycloadditions
R1 H R1 R1
Cbz R2
Cbz R2 Cbz R2
N N2 + MWI or hυ N N
H N H H
O N +
N
R3
O R3 O R3
227 215 228a 228b
Scheme 9.74
utilizing photochemical reaction conditions but also under the action of microwave
irradiation. b-Lactams with a trans-substitution pattern on the ring were obtained ex-
clusively in 40±85 % yield.
9.8
Conclusions
Acknowledgments
References
10
Microwave Catalysis in Organic Synthesis
Milan Hjek
10.1
Introduction
The aim of this chapter is to draw the attention of experimental organic and catalytic
chemists to a new field of catalysis, especially to catalytic methods which use micro-
wave irradiation as a new means of activation of chemical reactions, called ªmicro-
wave catalysisº. It is intended to advise synthetic organic chemists about the choice of
catalytic steps which might be more efficient than conventional synthetic methods.
This chapter focuses exclusively on microwave heterogeneous catalysis. Microwave
homogeneous catalysis by transition metal complexes is treated in Chapt. 11, phase
transfer catalysis in Chapt. 5, catalytic reactions on graphite in Chapt. 7, photocataly-
tic reactions in Chapt. 14, and catalytic synthesis of labeled compounds in Chapt. 13.
The development of microwave heterogeneous catalysis is, however, impeded be-
cause most synthetic chemists are not well acquainted with factors affecting both
heterogeneous catalysis and microwaves. Although attempts have been made to ac-
celerate reaction rate or to improve yield and selectivity, they have been based on the
rule ªlet's try microwaves and see what happensº.
This chapter is written to help the synthetic chemist to understand the effect of
microwaves on heterogeneously catalyzed reactions. Not only are metal catalysts dis-
cussed, but also metal oxides, zeolites, clays, and similar materials that act either as
catalysts or as potential supports for the catalytically active species. The factors in-
volved in the preparation of catalysts under microwave conditions are summarized
and discussed in terms of their effect on catalyst activity and selectivity. Basic me-
chanistic understanding can frequently be used to modify reaction conditions to
achieve product formation in high yield. Such an approach cannot be readily applied
to heterogeneously catalyzed processes, because the mechanistic understanding
needed by the synthetic chemist is not yet commonly available for this type of reac-
tion. This chapter is intended to call the attention of synthetic chemists to microwave
catalysis, because it has some advantages over conventional heterogeneous catalysis
(there are also some problems). Hopefully, this may encourage synthetic and cataly-
tic chemists to use the process more frequently and thus extend the application of
microwaves to the preparation of a variety of materials.
10.1.1
Definitions
Microwave radiation can be used to prepare new catalysts, enhance the rates of che-
mical reactions, by microwave activation, and improve their selectivity, by selective
heating. The heating of the catalytic material generally depends on several factors in-
cluding the size and shape of the material and the exact location of the material in the
microwave field. Its location depends on the type of the microwave cavity used [2].
The aim of this chapter is to describe some advances in catalysis achieved by use
of microwave heating. The aspects of microwave catalytic reactions which differ
from traditional thermal methods are emphasized. The input of microwave energy
into a reaction mixture is quite different from conventional (thermal) heating and it
is the task of synthetic chemists to exploit this special situation as fully as possible.
10.2
Preparation of Heterogeneous Catalysts
The interaction of microwaves with solid materials has proven attractive for the pre-
paration and activation of heterogeneous catalysts. It has been suggested that micro-
wave irradiation modifies the catalytic properties of solid catalysts, resulting in in-
creasing rates of chemical reactions. It is evident that microwave irradiation creates
catalysts with different structures, activity, and/or selectivity. Current studies docu-
ment a growing interest in the preparation of microwave-assisted catalysts and in
the favorable influence of microwaves on catalytic reactions.
The preparation of catalysts usually involves the impregnation of a support with a
solution of active metal salts. The impregnated support is then dried, calcined to de-
compose the metal salt and then reduced (activated) to produce the catalyst in its ac-
tive form. Microwaves have been employed at all stages of catalyst preparation. Bene-
ficial effects of microwave heating, compared with conventional methods, have been
observed especially in the drying, calcination, and activation steps.
10.2.1
Drying and Calcination
It is well known that microwave drying of many solid materials is a very efficient
and widely used process even on an industrial scale [3]; it is also an attractive means
of drying of heterogeneous catalysts. Microwave drying of catalysts and supported
sorbents has several advantages:
Microwave heating has been reported to produce materials with particular physi-
cal and chemical properties [4]. Stable solid structures are formed at low reaction
348 10 Microwave Catalysis in Organic Synthesis
temperatures with unusually high surface areas, making them very useful as cata-
lysts or catalyst supports. Calcination of solid precursors in a microwave field has
significant advantages over conventional heating. The effective synthesis of the cata-
lysts and supporting adsorbents has been reported for the examples below.
Microwave drying of an alumina-supported nickel catalyst reduced the drying time
by a factor of 2±3 [4]. The results indicated that better dispersion of nickel was achieved
by microwave drying, presumably because of minimization of moisture gradients dur-
ing drying. Analysis of the results confirmed, moreover, that the microwave-dried
samples were significantly stronger than those dried conventionally. As a conse-
quence of minimal moisture gradients, the metal ions are not redistributed to the
same extent during the microwave-drying process as they are when the samples are
dried conventionally, because of the moisture leveling [2, 3] process. As a consequence
of minimal moisture gradients, metal ions are not distributed to the same extent dur-
ing MW drying process as under conventional drying. This causes the resulting dry
pellets to be stronger, as shown by analysis of crushing strengths [4]. Because the heat-
ing effect is approximately proportional to the moisture content, microwaves are ideal
for equalizing moisture within the product in which the moisture distribution is initi-
ally nonuniform. Microwave drying also proceeds at relatively low temperatures, and
no part of the product needs be hotter than the evaporating temperature.
The microwave technique has also been found to be a potential method for the
preparation of the catalysts containing highly dispersed metal compounds on high-
porosity materials. The process is based on thermal dispersion of active species, fa-
cilitated by microwave energy, into the internal pore surface of a microporous sup-
port. Dealuminated Y zeolite-supported CuO and CuCl sorbents were prepared by
this method and used for SO2 removal and industrial gas separation, respectively [5].
The results demonstrated the effective preparation of supported sorbents by micro-
wave heating. The method was simple, fast, and energy-efficient, because the synth-
esis of both sorbents required a much lower temperature and much less time com-
pared with conventional thermal dispersion.
The V2O5/SiO2 catalyst for o-xylene oxidation prepared by wet impregnation under
microwave irradiation had several advantages [6] compared with that prepared by the
conventional thermal method:
The more active cobalt catalyst for pyrolytic reactions was prepared by microwave
calcination of cobalt nitrate which was converted to cobalt oxide by rapid microwave
heating [7].
The high dispersion of inorganic salts (CuCl2, NiCl2, AuCl3, RuCl3, etc.) on the
surface of zeolites (NaZSM-5, NaY, NaBeta) and alumina with high loading of the ac-
tive components has recently been achieved by microwave techniques [8±10]. The
10.2 Preparation of Heterogeneous Catalysts 349
catalysts were very active in NOx decomposition, even at room temperature, CO and
NO were partially converted to CO2 and N2. It was concluded that microwave treat-
ment is a new route for dispersing a high loading of inorganic compounds onto the
surface of supports to form highly active catalysts.
The microwave technique has been also found to be the best method for preparing
strongly basic zeolites (ZSM-5, L, Beta, etc.) by direct dispersion of MgO and KF.
This novel procedure enabled the preparation of shape-selective, solid, strongly base
catalysts by a simple, cost-effective, and environmentally friendly process [11, 12].
New solid bases formed were efficient catalysts for dehydrogenation of 2-propanol
and isomerization of cis-2-butene.
In the microwave synthesis of zeolites, a mixture of a precursor and a zeolite sup-
port is heated in a microwave oven. The sample is then tested for its catalytic activity
and the results compared with the sample obtained by the conventional method. Mi-
crowave irradiation at the calcination stage led to samples with a more uniform parti-
cle-size distribution and microstructure and to a bimetallic catalyst with different
morphology. Microwave calcination of magnesia-, alumina-, and silica- supported Pd
and Pd±Fe catalysts resulted in their having enhanced catalytic activity in test reac-
tions ± hydrogenation of benzene and hydrodechlorination of chlorobenzene ± com-
pared with conventionally prepared catalysts [13±15]. The higher catalytic activity
was attributed to the prevention of formation of a Pd±Fe alloy of low activity, which
occurs at the high reduction temperature used in conventional heating.
The microwave technique for drying then calcination is an excellent way of obtain-
ing highly porous silica gel with a high surface area (as high as 635 m2 g±1) for use
as a catalyst and as a catalyst support [16].
The dispersion and solid-state ion exchange of ZnCl2 on to the surface of NaY zeo-
lite by use of microwave irradiation [17] and modification of the surface of active car-
bon as catalyst support by means of microwave induced treatment have also been re-
ported [18]. The ion-exchange reactions of both cationic (montmorillonites) and anio-
nic clays (layered double hydroxides) were greatly accelerated under conditions of
microwave heating compared with other techniques currently available [19.]
Microwave irradiation has also been applied to the preparation of Fe2O3/SO2± 4
superacid [20, 21] and high-surface aluminum pillared montmorillonites [20].
The most successful application of microwave energy in the preparation of hetero-
geneous solid catalysts has been the microwave synthesis and modification of zeo-
lites [21, 22]. For example, cracking catalysts in the form of uniformly sized Y zeolite
crystallites were prepared by microwave irradiation in 10 min, whereas 10±50 h were
required by conventional heating techniques. Similarly, ZSM-5 was synthesized in
30 min by use of this technique. The rapid internal heating induced by microwaves
not only led to a shorter synthesis time, and high crystallinity, but also enhanced
substitution and ion exchange [22].
Microwave processing of zeolites and their application in the catalysis of synthetic
organic reactions has recently been excellently reviewed by Cundy [23] and other
authors [24]. The microwave synthesis of zeolites and mesoporous materials was sur-
veyed, with emphasis on those aspects which differ from conventional thermal
methods. The observed rate enhancement of microwave-mediated organic synthesis
350 10 Microwave Catalysis in Organic Synthesis
achieved by use of these catalysts was caused by a variety of thermal effects, includ-
ing very high rates of temperature increase, bulk superheating, and differential heat-
ing. Examples of microwave activation of chemical reactions catalyzed by zeolites
will be presented in Sect. 10.3.
An efficient oxidation catalyst, OMS-1 (octahedral mol. sieve), was prepared by mi-
crowave heating of a family of layered and tunnel-structured manganese oxide mate-
rials. These materials are known to interact strongly with microwave radiation, and
thus pronounced effects on the microstructure were expected. Their catalytic activity
was tested in the oxidative dehydrogenation of ethylbenzene to styrene [25].
In the preparation of microporous manganese oxide materials different chemical
properties were observed for the microwave and thermal preparations. In the conver-
sion of ethylbenzene to styrene the activity and selectivity of the materials was differ-
ent [26].
An alternative approach for the preparation of supported metal catalysts is based
on the use of a microwave-generated plasma [27]. Several new materials prepared by
this method are unlikely to be obtained by other methods. It is accepted that use of a
microwave plasma results in a unique mechanism, because of the generation of a
nonthermodynamic equilibrium in discharges during catalytic reactions. This can
lead to significant changes in the activity and selectivity of the catalyst.
10.2.2
Catalyst Activation and Reactivation (Regeneration)
Microwave irradiation of catalysts before their use in chemical reactions has been
found to be a new promising tool for catalyst activation. Microwave irradiation has
been found to modify not only the size and distribution of metal particles but prob-
ably also their shape and, consequently, the nature of their active sites. These phe-
nomena might have a significant effect on the activity and selectivity of catalysts, as
found in the isomerization of 2-methylpentene on a Pt catalyst [2].
As an example, the microwave activation of the platinum catalyst under conditions
when it was highly sensitive to thermal treatment resulted in an increase of its cata-
lytic activity and selectivity (from 40 to 80 %) [28].
Durable changes of the catalytic properties of supported platinum induced by mi-
crowave irradiation have been also recorded [29]. A drastic reduction of the time of
activation (from 9 h to 10 min) was observed in the activation of NaY zeolite catalyst
by microwave dehydration in comparison with conventional thermal activation [30].
The very efficient activation and regeneration of zeolites by microwave heating can
be explained by the direct desorption of water molecules from zeolite by the electro-
magnetic field; this process is independent of the temperature of the solid [31]. Inter-
action between the adsorbed molecules and the microwave field does not result sim-
ply in heating of the system. Desorption is much faster than in the conventional
thermal process, because transport of water molecules from the inside of the zeolite
pores is much faster than the usual diffusion process.
Very little is known about the reactivation (regeneration) of used catalysts by mi-
crowave irradiation. Catalyst activity has been shown to decay with increasing carbon
10.3 Microwave Activation of Catalytic Reactions 351
deposition, and several patents disclose the decarbonization of cracking zeolite cata-
lysts by microwaves [32]. Because carbon is a very lossy material (it absorb micro-
waves very efficiently), any carbon deposited on the surface of the catalyst is strongly
heated. In the presence of air or hydrogen the carbon is removed in the form of car-
bon dioxide or methane, respectively. When carbon deposition reaches a certain level
it starts to absorb microwave energy strongly and is, therefore, subsequently re-
moved, leading to an increase in the activity of the catalyst.
Alumina spheres polluted by carbon residues have been also reactivated by use of
microwaves [33]. Their regeneration has been performed in a stream of air and in
the presence of silicon carbide as an auxiliary microwave absorber. Microwave heat
treatment led to full recovery of the catalyst in times varying from a half to a quarter
of the conventional treatments. Regeneration of a commercial Ni catalyst (Ni/Al2O3)
deactivated, presumably, by coke formation, by means of a flow of hydrogen or oxy-
gen and water vapor under the action of microwave irradiation was, however, unsuc-
cessful [34].
Microwaves are frequently used in the laboratory by synthetic organic chemists
for regeneration and activation of solids such as molecular sieves, silica gel or alu-
mina when fast and complete drying is required.
10.3
Microwave Activation of Catalytic Reactions
10.3.1
Reactions in the Liquid Phase
Cat.
CH3(CH2 )14COOH + CH3(CH2)2CH2OH CH3(CH2)14COOCH2(CH2)2CH3 + H2O
It was found that the reaction under microwave conditions was faster than that
heated conventionally. Yields were greatly improved ± from 50±82 % for conventional
heating to 71±95 % for microwave heating when heterogeneous catalysts were used.
The catalysts Fe2(SO4)3, TiBu4, KF, KSF, and PTSA were used in a continuous-flow
reactor. The results were further improved by use of a heat captor (graphite) and si-
multaneous use of ultrasound. Better results (yield and rate) under microwave condi-
tions were also achieved in other esterification reactions of stearic and acetic acid in
the presence of a heterogeneous catalyst (Fe2(SO4)3) adsorbed on montmorillonite
clay pellets 4±5 mm in diameter [36, 37]. The reaction rate increased by 50±150 %
compared with conventional heating. This increase was most probably because of
superheating of voluminous pellets (5 mm), the temperature of which was calcu-
lated to be 9±18 K above the bulk temperature. When the esterification was homoge-
neously catalyzed by sulfuric acid using the two modes of heating no differences be-
tween yield and reaction rates were observed.
The effect of the mode of heating was also studied in heterogeneously catalyzed
esterification of acetic acid by isopentyl alcohol in the presence of Amberlyst-15 ca-
tion exchange resin catalyst [38], Scheme 10.2.
Amberlyst
CH3COOH + (CH3)2CHCH2CH2OH CH3COOCH2CH2CH(CH3)2 + H2O
R CO O CH2 CH2 OH
B4C
R COO CH + 3 CH3OH 3 R COOCH3 + CH OH
R COO CH2 CH2 OH
Y Ph HCOONH4 Y Ph
10% Pd/C
I
5% Pd/MgO CN
+ CN
E/Z
Scheme 10.5 Arylation of acrylonitrile with iodobenzene.
The E/Z ratio of the isomers formed over a supported Pd catalyst differed from
that obtained with the homogeneous counterpart. The activity of the catalyst under
the action of microwave and conventional conditions was comparable, but micro-
wave irradiation improved yields and reduced reaction times.
Heck C±C coupling reactions were also facilitated by the presence of a palladium
catalyst when Pd was deposited on a tubular membrane of porous glass. Thus, the
coupling of iodobenzene with allyl alcohol affording 3-phenylpropionaldehyde in the
presence of this Pd catalyst had several advantages ± the ease of catalyst manufac-
ture, mechanical strength, thermal stability, and resistance to organic solvents [46].
The hydrolysis of sucrose catalyzed by the strongly acidic cation-exchange resin
Amberlite 200C in RH form was chosen as a model reaction to compare the use of
stirred tank and continuous-flow reactors [47±49], Scheme 10.6.
No rate enhancement was observed when the reaction was performed under mi-
crowave irradiation at the same temperature as in conventional heating [47]. Similar
reaction kinetics were found in both experiments, presumably because mass and
heat effects were eliminated by intense stirring [47]. The model developed enabled
accurate description of microwave heating in the continuous-flow reactor equipped
with specific regulation of microwave power [47, 48]. Calculated conversions and
yields of sucrose based on predicted temperature profiles agreed with experimental
data.
Several other miscellaneous heterogeneously catalyzed reactions have been per-
formed in the liquid phase. Hexane was successfully oxyfunctionalized with aqueous
hydrogen peroxide by use of the zeolite TS-1 catalyst [50] and microwave-promoted
acetalization of a number of aldehydes and ketones with ethylene glycol proceeded
readily (2 min) in the presence both of heterogeneous (acidic alumina) and homoge-
neous (PTSA, Lewis acids) catalysts [51], Scheme 10.7.
Amberlite
C12H22O 11 + H2O C6H12O 6 + C6H12O 6
sucrose glucose fructose
O Al2O 3
O O + H2O
R1 R2 HO OH R1 R2
Cat.
CH3O + COCl CH3O C
- HCl
O
Scheme 10.8 Benzoylation of anisole.
Yields were high (up to 97 %) and comparable with those of homogeneously cata-
lyzed reactions. A much higher catalyst-to-substrate ratio had to be used with hetero-
geneous alumina (10±15) than with the homogeneous catalysts (0.015±0.050), how-
ever. It was concluded that the microwave method led to considerable improvement
of acetalization reactions, compared with conventional methods.
Friedel±Crafts acylation of aromatic ethers has been performed in the presence of
a variety of metal chlorides and oxides (FeCl3, ZnCl2, AlCl3, Fe2O3, Fe3O4, etc.) but
without temperature control [52], Scheme 10.8.
The short reaction time (1 min, 160 8C) in the benzoylation of anisole was prob-
ably a result of large temperature gradients rather than a nonthermal microwave ef-
fect.
Some reactions have been found to proceed with better results in the absence of
solvent, probably because of the creation of temperature gradients which are elimi-
nated in the presence of a stirred solvent. This was observed for the Diels±Alder reac-
tion of a-amino acid precursors with cyclopentadiene catalyzed by heterogeneous cat-
alysts (SiO2±Al, SiO2±Ti), when the reaction was performed in toluene or in the ab-
sence of solvent [53]. Microwave activation increased the rate of reaction without re-
ducing the selectivity of the reaction.
The microwave activation of Michael additions in the preparation of N-substituted
imidazoles afforded excellent yields in very short reaction times under mild reaction
conditions, Scheme 10.9. Basic clays (Li+, Cs+) exchanged montmorillonites were
found to be very active and selective catalysts for the Michael addition of imidazole
and ethyl acrylate [54].
A total of 75 % conversion with 100 % selectivity was obtained in only 5 min in the
absence of solvents.
An efficient method for conversion of a variety of acids into their corresponding
amides in the presence of zeolite±HY under the action of microwave irradiation has
recently been described [55], Scheme 10.10.
N Clay N CO2Et
NH + CO2Et N
O O
Zeolite HY
+ H2N R 2 + H2O
R1 OH R1 NH R 2
Ra-Ni
NH2 + ROH NH R
OH OH OH
Zeolite Zeolite
+
H2O 2 H2 O2
HO OH
Scheme 10.12 Oxidation of benzene with hydrogen peroxide.
OH OH OH OH
KSF
+ + +
(K10)
10.3.2
Reactions in the Gas Phase
catalyst (Ni/ZrO2, Ni/La2O3) had higher activity and selectivity than the Co catalyst
(Co/ZrO2, Co/La2O3). It was proposed that nonuniform distribution of the micro-
wave energy on the catalyst surface created ªhot spotsº (Sect. 10.3.3) as active centers
for the catalytic reaction.
Microwaves have been used to generate plasma in methane at 5±50 Torr. The radi-
cals produced in such a system were then allowed to react over a nickel catalyst, af-
fording a mixture of ethane, ethene, and ethyne [74].
Methane has also been used as the reducing agent in the catalytic conversion of
NO to N2 over Co-ZSM-5 zeolites [75] in the presence of oxygen. The high NO con-
versions (>70 %) were achieved by microwave irradiation at 250±400 8C, whereas un-
der similar conditions thermal runs failed to convert either NO or methane in signif-
icant amounts. The high activity and selectivity of the reduction of NO by methane
achieved with microwave irradiation was probably because of the activation of
methane to form methyl radicals at relatively low reaction temperatures.
M / Al 2O3
CH4 + NH3 HCN + 3 H2
The reaction was performed over a series of Pt/Al2O3, Ru/Al2O3, and carbon-sup-
ported catalysts under the action of pulsed microwave radiation; conversions ex-
ceeded 90 % and acetonitrile was formed as the byproduct.
under the action of conventional and microwave heating [28, 79]. The isomer selec-
tivity of transformation of 2-methylpentane to 3-methylpentane, methylcyclopen-
tane, and hexane on this catalyst increased from 40 to 80 % when the catalyst was
pretreated with microwave energy [79]. The beneficial change in selectivity was
found to be permanent. Microwave irradiation has been found to be a new, original
means of activation of supported metal catalysts, in particular reforming catalysts
[28, 29]. Hydrogenolysis of methylcyclopentane was studied mechanistically, because
methylcyclopentane is one of the four molecules in equilibrium with the intermedi-
ate cyclopentane in the cyclic mechanism [2]. Four possible effects of the electromag-
netic field on catalytic reactions were identified.
10.3.3
Mechanistic Aspects
Although microwave activation of catalytic reactions has been the subject of many
studies (Sects. 10.3.1 and 10.3.2), the mechanism of these reactions is not yet fully
understood. In heterogeneous catalytic liquid/solid and gas/solid systems many re-
sults have revealed significant differences between the rates of conventionally and
microwave heated reactions. As a rule, at the same temperature microwave heated
reactions were faster than conventional and their rate enhancement was over one or-
10.3 Microwave Activation of Catalytic Reactions 363
der of magnitude. Such rate differences have not been reported for other catalytic re-
actions, so how should they be interpreted? The reactions can be divided into two ca-
tegories according to the rate enhancement effect observed under microwave condi-
tions compared to conventional heating ± those accelerated by application of micro-
waves and those which are not.
12. Cyclization of citronellal in the presence of KSF catalyst [93]. Reaction condi-
tions: reflux, CCl4 as solvent, single-mode reactor¹
13. CO oxidation over Pd/g-Al2O3 and Pt/g-Al2O3 [94, 95]. Reaction conditions: sin-
gle-mode packed bed microwave reactor operated on continuous basis at
915 MHz.
Several reasons have been proposed to account for the effect of microwave heating
on chemical reactions and catalytic systems. The results summarized in 1 to 7,
above, show that under specific conditions microwave irradiation favorably affects re-
action rates of both the liquid- and gas-phase processes. This phenomenon has been
explained in terms of microwave effects, i. e. effects which cannot be achieved by con-
ventional heating. These include superheating, selective heating, and formation of hot
spots (and possibly nonthermal effects).
These microwave effects can be regarded as thermal. The proposal of some authors
on the operation of nonthermal effects is still controversial (see Chapt. 3).
In the literature, thermal microwave effects are still the subject of some misunder-
standings. For that reason, let us discuss the matter in more detail.
Selective heating
Selective heating generally means that in a sample containing more than one com-
ponent, only that component which couples with microwaves is selectively heated.
The nonabsorbing components are not thus heated directly but only by heat transfer
from the heated component. For example:
The selective heating is more enhanced when the catalyst contains a low micro-
wave absorbing support (e. g. a-Al2O3 rather than g-Al2O3). Metal or metal oxides
supported on transparent support may be therefore selectively heated. Thus for ex-
ample, Pt sites in Pt/SiO2 catalyst can be selectively heated, in contrast to the trans-
parent, unheated SiO2 support. Hence, selective heating of the active sites to the
temperature higher than that of the support may occur if heat loss from the metal
particles is not too fast. Microwave field does not couple with transparent ceramic
support, but it may strongly couple with metal particles because of their high electri-
cal conductivity. Unfortunately, to get a direct experimental proof of this concept
seems very unlikely, since to measure the temperature of individual active sites is be-
yond current experimental possibilities. Until now, the possibility to heat selectively
the active sites without rising the temperature of the catalyst bed has been only mod-
eled [102] and found to be strongly dependent on the size of catalyst particles and mi-
crowave frequency [103]. A number of experimental studies on catalytic reactions un-
der microwave conditions were concerned with kinetic aspects and showed that the
reaction rates and product distribution correspond to a higher reaction temperature
than was that measured for the bulk of catalyst bed. Both reaction characteristics
were often explained in terms of the higher local temperature of certain active cen-
ters within the catalyst bulk. It is worth mentioning that the results of theoretical cal-
culation reported by Thomas [102] did not substantiate the possibility of achieving
selective heating of supported metal catalysts by microwaves. However, a recent de-
tailed modeling of small-scale, microwave-heated, packed bed and fluidized bed cata-
lytic reactors indicated that under specific conditions, the active sites may be selec-
tively heated in both types of catalyst bed [103, 104].
Hot spots
Hot spots are created by a nonlinear dependence of thermal and electromagnetic
properties of the material being heated on temperature. If the rate at which microwave
energy is absorbed by the material increases faster than linearly with temperature,
then heating does not take place uniformly, and the regions of very high temperature
can create hot spots. Nonhomogenity of electromagnetic field contributes to the crea-
tion of hot zones significantly. Generally, hot spots can be created by localized super-
heating, selective heating as well as by nonhomogenity of electromagnetic field.
In a microwave-heated packed catalyst bed, two different forms of hot spots can be
created.
1. Hot spots as large scale nonisothermalities, which can be detected and measured
by optical pyrometers, fiber optic or IR pyrometers, i. e. these are macro scale hot
zones.
10.3 Microwave Activation of Catalytic Reactions 367
2. Hot spots as a temperature gradient between the metal particles and the support,
which cannot be detected and measured because they are close to micro scale, i. e.
they possess molecular dimensions, they are closed to selective heating of active
sites.
Unfortunately, the microwave radiation effects on molecular level are not well un-
derstood.
Zhang et al. [92] presented a certain evidence for the formation of hot spots in the
microwave experiments and demonstrated that these hot spots need not be exclu-
sively localized on the active sites but may also involve the support material
(g-Al2O3). They also estimated the dimensions of these hot spots to be in the region
of 90±1000 mm. Development of hot spots in the catalyst bed in the course of the gas
phase decomposition of H2S catalyzed by MoS2/Al2O3 was probably the reason for a
significant apparent shift in the equilibrium constant. The temperature of these hot
spots was probably by 100±200 K above the bulk temperature. The formation of hot
spots in the support could be due to the absorptive properties of g-Al2O3 (compared
to low absorptive a-Al2O3). It was concluded that these hot spots also induce a con-
siderable reorganization of the catalyst under microwave conditions. Such selective
or localized superheating to create hot spots as effected with microwave heating can-
not be achieved by conventional heating methods.
If microwave heating leads to enhanced reactions rates, it is plausible to assume
that the active sites on the surface of the catalyst (micro hot spots) are exposed to se-
lective heating which causes some pathways to predominate. In the case of metal
supported catalysts, the metal can be heated without heating of the support due to
different dielectric properties of both catalyst components. The nonisothermal nat-
ure of the microwave-heated catalyst and the lower reaction temperature affects fa-
vorably not only reaction rate but also selectivity of such reactions.
Chen et al. [70] suggested that temperature gradients may have been responsible
for the more than 90 % selectivity of the formation of acetylene from methane in a
microwave heated activated carbon bed. The authors believed that the highly non-
isothermal nature of the packed bed might allow reaction intermediates formed on
the surface to desorb into a relatively cool gas stream where they are transformed via
a different reaction pathway than in a conventional isothermal reactor. The results
indicated that temperature gradients were approximately 20 K. The nonisothermal
nature of this packed bed resulted in an apparent rate enhancement and altered the
activation energy and pre-exponential factor [94]. Formation of hot spots was mod-
eled by calculation and, in the case of solid materials, studied by several authors
[105±108].
It is obvious that nonisothermal conditions induced by microwave heating lead to
very different results from those obtained under conventional heating conditions. In
summary microwave effects like superheating, selective heating and hot spots, can
all be characterized by temperature gradients ranging from macroscopic to molecu-
lar scale dimensions.
Let us now return to the question ªHow to interpret or explain the fact that some
reactions are affected by microwaves and some reactions do notº, as documented by
368 10 Microwave Catalysis in Organic Synthesis
examples 1±7 and 8±13, above. A detailed study of this subject has been performed
by Hjek et al. [58±60] for heterogeneous catalytic liquid phase reactions. Transfor-
mation of 2-t-butylphenol into phenol, 4-t-butylphenol, 2,4-di-t-butylphenol and iso-
butene on montmorillonites as catalysts (KSF, K10) was chosen as a model reaction.
Both the reactant and the catalysts coupled very well with microwaves. KSF and K10
catalysts in the form of a fine powder (10±15 mm) were used to avoid creation of hot
spots (as in the presence of voluminous catalyst pellets, e. g. 5 mm [36, 37]). The re-
sults can be summarized as follows.
When the results obtained in the liquid phase are applied to the reactions 1±7 and
8±13, above, both for the liquid and gas phase, one can conclude that if reactions are
performed under conditions where temperature gradients can be completely elimi-
nated, e. g. on using efficient stirring, nonpolar solvents, a high flow of reactants in
continuous flow system, fluidized bed and then any differences between microwave
and conventional heating conditions may disappear. However, from the synthetic
point of view it is more attractive to perform synthetic reactions using microwaves
under conditions favorable to the production of temperature gradients, because
higher reaction rate and improved selectivities can be obtained more easily com-
pared to conventional heating methods.
10.3.4
Microwave Catalytic Reactors
ing pump and pressure gauge at the inlet end and to a heat exchanger and pressure-
regulating valve at the effluent end. Temperature is monitored outside the cavity at
the inlet and the outlet.
The reactor has facilitated a diverse range of synthetic reactions at temperatures
up to 200 8C and 1.4 Pa. The temperature measurements taken at the microwave
zone exit indicate that the maximum temperature is attained, but they give insuffi-
cient information about thermal gradients within the coil. Accurate kinetic data for
studied reactions are thus difficult to obtain. This problem has recently been avoided
by using fiber optic thermometer. The advantage of continuous-flow reactor is the
possibility to process large amounts of starting material in a small volume reactor
(50 mL, flow rate 1 L h±1). A similar reactor, but of smaller volume (10 mL), has been
described by Chen et al. [117].
It was noted that hydrolysis of sucrose using a strong acidic cation-exchange resin
as heterogeneous catalyst gives better results than with soluble mineral acids [118].
The hydrolysis of sucrose catalyzed by the strong acidic cation-exchange resin Am-
berlite 200C was used by Plazl et al. [48, 49] to test successfully a continuous-flow cat-
alytic reactor with packed catalyst bed made by a modification of a commercial mi-
crowave oven (Panasonic NE-1780). The continuous-flow catalytic reactor was built
by modification of domestic ovens (Panasonic, General Electric) also by other
authors [38, 40]. The continuous-flow microwave reactor based on single-mode sys-
tem was described by Chemat [37] and Marqui [118]. Better isothermal conditions,
the possibility of selective heating of catalyst bed and scaling-up are the main advan-
tages of single-mode continuous-flow reactors compared to multi-mode systems for
catalytic reactions in the liquid phase.
For gas phase heterogeneous catalytic reactions, the continuous-flow integral
catalytic reactors with packed catalyst bed have been exclusively used [61±91]. Con-
tinuous or short pulsed-radiation (milliseconds) was applied in catalytic studies
(see Sect. 10.3.2). To avoid the creation of temperature gradients in the catalyst
bed, a single-mode radiation system can be recommended. A typical example of
the most advanced laboratory-scale microwave, continuous single-mode catalytic
reactor has been described by Roussy et al. [79] and is shown in Figs. 10.4 and
10.5.
The principle aspects of this microwave irradiation system are, briefly:
372 10 Microwave Catalysis in Organic Synthesis
In addition to packed catalyst bed, a fluidized bed irradiated by single and multi-
mode microwave field, respectively, was also modeled by Roussy et al. [120]. It was
proved that the equality of solid and gas temperatures could be accepted in the sta-
tionary state and during cooling in a single-mode system. The single-mode cavity
eliminates the influence of particle movements on the electric field distribution.
When the bed was irradiated in the multimode cavity, the model has failed. Never-
theless, the combination of the fluidization technique with microwave heating pro-
vides significant benefits such as reduced energy costs, appreciable reduction in pro-
cessing time, and higher product quality. A fluidized bed reactor has been success-
fully applied in the destruction of hazardous/toxic organic compounds as trichlor-
oethylene, p-xylene, naphthalene, and some gasoline-like hydrocarbon materials
[121]. Granulated activation charcoal was used as the catalyst. All these organic com-
pounds could be completely decomposed by the combination of these two technolo-
gies.
10.4
Industrial Applications
Since the early 1980s there has been a great deal of interest in the development of in-
dustrial applications of microwave irradiation in catalytic processes. This interest has
been driven by the unique attributes of the selective interactions of microwave radia-
tion with different materials as the basis for the initiation and control of catalytic che-
mical reactions. The microwave mode of energy conversion has many attractive fea-
tures for the chemist because some liquids and solids can absorb in situ and trans-
form electromagnetic energy into heat very effectively [2]. However, microwave tech-
nology has so far only rarely attracted industrial attention. A number of potential che-
mical processes using this technique have been established in research laboratories,
but the large scale-up to commercial practice remains elusive. Microwave technology
is not new to physicists and electrical engineers, but the concept of using microwaves
as an energy source for chemical reactions has only recently been appreciated by che-
mists. In the subsequent stage of the scale-up of the process, some of potential advan-
tages of the microwave induced catalysis proved in laboratory experiments might not
be easy to realize in a commercial scale production. This is because of the necessity
to identify many difficulties and modifications, technical and economical, before the
final choice of a proper system can be designed and installed [122].
Wan et al. [122] demonstrated problems of the scaling-up in the case of the cataly-
tic conversion ethane into ethyne by finding that the optimization of the process for
the large-scale reaction system was quite different from the laboratory experiment. It
is not practical to construct a huge quartz reactor enclosed by a huge microwave
oven. Rather, the microwave radiation can be introduced into a stainless steel reactor.
There are obviously many other engineering problems to be encountered, such as
penetration depth, uniformity of temperature profile of catalyst bed and the like,
which are specific to a microwave reactor. In large scale processes the spatial unifor-
mity may be often critical. Engineering and scale-up aspects of microwave induced
catalytic reactions were summarized by Mehdisadeh [123] of DuPont de Memours
Co. It was recommended that the choice of the catalyst should be made not only for
its chemical merits, but also for its microwave properties. For example, a catalyst
with too low a dielectric loss would be inefficient, and could cause control difficul-
ties. On the other hand, a catalyst with too high a dielectric loss can cause low pene-
tration and uniformity problems.
374 10 Microwave Catalysis in Organic Synthesis
Despite the scaling-up problems, the following industrial processes have been pro-
posed, and disclosed mainly in the patent literature:
In conclusion, one can ask the question ªWhich microwave catalytic process is op-
erated on industrial scaleº? The answer has been frequently looked for at AMPERE
meetings, in particular at conferences in Prague (1998), Valencia (1999), or Antibes
(2000) and Bayreuth (2001). It was stated that relevant information concerning this
matter was missing, because an industrial microwave catalytic process put in opera-
tion had not been disclosed. The reason could be either that such a process has not
yet been successfully realized or that the disclosure of new technology has been
strictly protected by industry. Nevertheless, the potential of the microwave catalysis
for industrial application is quite great, both for gas phase processes and continu-
ous-flow liquid phase reactors where even a small equipment can produce signifi-
cant amounts of products. Moreover, another potential application exists in environ-
mental catalysis, since recent legislative measures have stimulated the development
of new technologies to reduce toxic emissions or to remove and dispose of hazardous
wastes.
References
1 V. Sridar, Curr. Sci. 1998, 74, 446±450. 7 J.L. Kuester, Res. Chem. Intermed. 1994,
2 G. Roussy, J.A. Pearce, Foundations 20, 51±59.
and Industrial Applications of Microwave 8 F.-S. Xiao, W. Xu, S. Qiu, R. Xu,
and Radiofrequency Fields, John Wiley & J. Mater. Sci. Lett. 1995, 14, 598±599.
Sons, New York, 1995. 9 F.-S. Xiao, W. Xu, S. Qiu, R. Xu, Catal.
3 A.C. Metaxas, R.J. Meredith, Indus- Lett. 1994, 26, 209±215.
trial Microwave Heating, Peter Pere- 10 F.-S. Xiao, W. Zhang, M. Jia,Y. Yu,
grinns, 1993. C. Fang, G. Tu, S. Zheng, S. Qiu,
4 G. Bond, R.B. Moyes, D.A. Whan, R. Xu, Catal. Today 1999, 50, 117±123.
Catal. Today 1993, 17, 427±437. 11 Y. Wang, J.H. Zhu, J.M. Cao,Y. Chun,
5 S.G. Deng,Y.S. Lin, Chem. Eng. Sci. Q.H. Xu, Micropor. Mesopor. Mater.
1997, 52, 1563±1575. 1998, 26, 175±184.
6 Y. Liu,Y. Lu, S. Liu,Y. Yin, Catal. Today 12 J. Zhu,Y.Chun,Y. Wang, Q. Xu, Catal.
1999, 51, 147±151. Today 1999, 51, 103±111.
References 375
13 N. Lingaiah, P.S.S. Prasad, P.K. Rao, A.V. Sapre, G.J. Teitman, 1990,
L.E. Smart, F.J. Berry, Appl. Catal. A: US Patent 4968403; W.J. Murphy, 1993,
General 2001, 213, 189±196. US Patent 5205912.
14 F.J. Berry, L.E. Smart, P.S.S. Prasad, 33 P. Veronesi. C. Lionelli, A.B. Cor-
N. Lingaiah, P.K. Rao, Appl. Catal. A: radi, Ceram. Trans. 1997, 80, 249±256.
General 2000, 204,191±201. 34 T.A. Treado, Proc. 30th Int. Microwave
15 P.S.S. Prasad, N. Lingaiah, P.K. Rao, Power Symp., Colorado, USA, 9±12 Sep-
F.J. Berry, L.E. Smart, Catal. Lett. tember 1995, pp. 4±7.
1995, 35, 345±351. 35 F. Chemat, M. Poux, S.A. Galema,
16 S. Rajeshkumar, G.M. Anilkumar, J. Chem. Soc., Perkin Trans. 2 1997,
S. Ananthakumar, K.G.K. Warrier, 2371±2374.
J. Porous. Mater. 1998, 5, 59±63. 36 F. Chemat, M. Poux, J.-L. di Martino,
17 D.H. Yin, D.L. Yin, Micropor. Mesopor. J. Berlan, Chem. Eng. Technol. 1996, 19,
Mater. 1998, 24, 123±126. 420±424.
18 J.A. Menendez, E.M. Menendez, 37 F. Chemat, D.C. Esveld, M. Poux,
M.J. Iglesias, A. Garcia, J.J. Pis, J. L. di Martino, J. Microwave Power
Carbon 1999, 37, 1115±1121. Electromagn. Energy 1998, 33, 88±94.
19 P. Monsef-Mirzai, D.M. Kavanagh, 38 K.G. Kabza, B.R. Chapados, J.E. Gest-
S. Bodman, S. Lange,W.R. McWhi- wicki, J.L. McGrath, J. Org. Chem.
nie, J. Microwave Power Electromagn. 2000, 65, 1210±1214.
Energy 1999, 34, 216±220. 39 S.D. Pollington, G. Bond,
20 Y. Ning, J. Wang, P. Hong, J. Mater. R.B. Moyes, D.A. Whan, J.P. Candlin,
Sci. Technol. 1996, 12, 307±311. J.R. Jennings, J. Org. Chem. 1991, 56,
21 G. Fetter, G. Heredia, L.A. Velazquez, 1313±1314.
A.M. Maubert, P. Bosch, Appl. Catal. 40 A. Breccia, B. Esposito, G. B. Frata-
A: General 1997, 162, 41±45. docchi, A. Fini, J. Microwave Power
22 A. Arafat, J.C. Jansen, A.R. Ebaid, Electromagn. Energy 1999, 34, 3±8.
H. van Bekkum, Zeolites 1993, 13, 162± 41 S. LeskovÉek, A. midovnik, T. Ko-
165. loini, J. Org. Chem. 1994, 59, 7433±
23 C.S. Cundy, Collect. Czech. Chem. Com- 7436.
mun. 1998, 63, 1699±1723 (118 refer- 42 B. Dayal, N.H. Ertel, K.R. Rapole,
ences). A. Asgaonkar, G. Salen, Steroids 1997,
24 H. Kosslick, H.-L. Zubova, U. Lohse, 62, 451±454.
H. Landmesser, Ceram. Trans. 1997, 80, 43 A.K. Bose, B.K. Banik, K.J. Barakat,
523±536. M.S. Manhas, Synlett 1993, 575±576.
25 E. Vileno, H. Zhou, Q.H. Zhang, 44 B.K. Banik, K.J. Barakat, D.R. Wagle,
S.L. Suib, D.R. Corbin, T.A. Koch, M.S. Manhas, A.K. Bose, J. Org. Chem.
J. Catal. 1999, 187, 285±297. 1999, 64, 5746±5753.
26 D.R. Baghurst, D.M.P. Mingos, 45 A. Wali, S.M. Pillai, S. Satish, React.
J. Chem. Soc., Chem. Commun. 1988, Kinet. Catal. Lett. 1997, 60, 189±194.
829±830. 46 J.T. Li, A.W.-H. Mau, C.R. Strauss,
27 S.L. Suib, Cattech 1998, 2, 75±84. Chem. Commun. 1997, 1275±1276.
28 L. Seyfried, F. Garin, G. Maire, 47 I. Plazl, S. LeskovÉek, T. Koloini,
J.-M. Thiebaut, G. Roussy, J. Catal. Chem. Eng. J. 1995, 59, 253±257.
1994, 148, 281±287. 48 S. LeskovÉek, I. Plazl, T. Koloini,
29 J.-M. Thiebaut, G. Roussy, M.-S. Med- Chem. Biochem. Eng. Q. 1996, 10, 21±
jram, F. Garin, L. Seyfried, G. Maire, 26.
Catal. Lett. 1993, 21, 133±138. 49 I. Plazl, G. Pipus, T. Koloini, AIChE J.
30 M. Hjek, Coll. Czech. Chem. Commun. 1997, 43, 754±760.
1997, 62, 347±354. 50 P.J. Kooyman, G.C.A. Luijkx, A. Ara-
31 G. Roussy, A. Zoulalian, M. Char- fat, H. van Bekkum, J. Mol. Catal. A:
reyre, J.-M. Thiebaut, J. Phys. Chem. Chemical 1996, 111, 167±174.
1984, 88, 5702±5708. 51 F.M. Moghaddam, A. Shariff, Synth.
32 J.A. Herbst, C.L. Markham, Commun. 1995, 25, 2457±2461.
376 10 Microwave Catalysis in Organic Synthesis
89 T.R.J. Dinesen, M.Y. Tse, M.C. Depew, 108 G.A. Kriegsman, P. Varatharajah,
J.K.S. Wan, Res. Chem. Intermed. 1991, J. Appl. Phys. 1993, 36, 221±228.
15, 113±127. 109 M.A.B. Pougnet, Rev. Sci. Instrum.
90 G. Bond, R.B. Moyes, I. Theaker, 1993, 64, 529±531.
D.A. Whan, Topics Catal. 1994, 1, 177± 110 M. Pagnotta, A. Nolan, L. Kim,
182. J. Chem. Educ. 1992, 69, 599±600.
91 A.Y. Klimov, B.S. Balzhinimaev, 111 D.M.P. Mingos, Chem. Ind. 1994, 596±
L.L. Makarshin,V.I. Zaikovskii, 599.
V.N. Parmon, Kinet. Catal. 1998, 39, 112 K.D. Raner, C. R. Strauss, R.W. Trai-
511±515. nor, J.S. Thorn, J. Org. Chem. 1995,
92 X.L. Zhang, D.O. Hayward, D.M.P. 60, 2456±2460.
Mingos, Chem. Commun. 1999, 975± 113 C.R. Strauss, R.W. Trainor, Aust. J.
976. Chem. 1995, 48, 1665±1692.
93 R. Laurent, A. Laporterie, J. Dubac, 114 C.R. Strauss, R.W. Trainor, Aust. J.
J. Berlan, S. Lefeuvre, M. Audhuy, Chem. 1995, 48, 1675, 1674.
J. Org. Chem. 1992, 57, 7099±7102. 115 S. Stinson, Chem. Eng. News 1990, 68,
94 W.L. Perry, J.D. Katz, D. Rees, 33±34.
M.T. Paffer, A.K. Datye, J. Catal. 1997, 116 T. Cablewski, A.F. Faux, C.R. Strauss,
171, 431±438. J. Org. Chem. 1994, 59, 3408±3412.
95 W.L. Perry, D.W. Cooke, J.D. Katz, 117 S.-T. Chen, S.-H. Chiou, K.-T. Wang,
A.K. Datye, Catal. Lett. 1997, 47, 1±4. J. Chem. Soc., Chem. Commun 1990,
96 D.R. Baghurst, D.M.P. Mingos, 807±809.
J. Chem. Soc., Chem. Commun. 1992, 118 J. Marqui, G. Salmoria, M. Poux,
674±677. A. Laporterie, J. Dubac, N. Roques,
97 D.M.P. Mingos, Res. Chem. Intermed. Ind. Eng. Chem. Res. 2001, 40, 4485±
1994, 20, 85±91. 4490.
98 D. Stuerga, P. Gaillard, Tetrahedron 119 G. Roussy, S. Hillaire, J.M. Thiebaut,
1996, 52, 5505±5510. G. Maire, F. Garin, S. Ringler, Appl.
99 P. Gaillard, N. R. Roudergue, Catal. A: Gen. 1997, 156, 169, 170.
D. Stuerga, Proc. 6th Int. Conf. Micro- 120 G. Roussy, S. Jassm, J.-M. Thiebaut,
wave and High Frequency Heating, J. Microwave Power Electromagn. Energy
Fermo, Italy, 9±13 September, 1997, 1995, 30, 178±187.
pp. 82±83. 121 G.C.-J. Jou, Carbon 1998, 36, 1643±
100 M. Hjek, H. Richterov, Proc. 6th 1648.
Int. Conf. Microwave and High Frequency 122 J.K.S. Wan, M.C. Depew, Ceram. Trans.
Heating, Fermo, Italy, 9±13 September, 2001, 111, 241±247.
1997, pp. 79±81. 123 M. Mehdizadeh, Res. Chem. Intermed.
101 J. Dolande, A. Datta, J. Microwave 1994, 20, 79±84.
Power Electromagn. Energy 1993, 28, 58± 124 J.K.S. Wan, Canadian Patent CA
67. 2031608, 1995.
102 J.R. Thomas, Ceram. Trans. 1997, 80, 125 C.L. O'Young, Eur. Pat. Appl. EP
397±406. 634211, 1995.
103 J.R. Thomas, Catal. Lett. 1997, 49, 137± 126 G. Roussy, J.-M. Thiebaut, O.M.
141. Souiri, A. Kiennemann, K.C. Petit,
104 J.R. Thomas, J.F. Faucher, J. Micro- G. Maire, Fr. Pat. Appl. FR 9211676,
wave Power Electromag. Energy 2000, 35, 1994.
165±174. 127 F.G. Dwyer, J.A. Herbst,Y.Y. Huang,
105 J.M. Hill, N.F. Smith, Math. Eng. Ind. H. Owen, P.H. Scipper, A.B. Schwartz,
1990, 2, 267±278. Int. Pat. Appl. PCT 87/02227, 1988.
106 C.J. Coleman, J. Aust. Math. Soc. B 128 G. Roussy, G. Maire, Eur. Pat. Appl. EP
1991, 33, 1±8. 519824, 1992.
107 G.A. Kriegsman, Ceram. Trans. 1991, 129 D.D. Tanner, Q.Z. Ding, P. Kanda-
21, 177±183. narachchi, J.A. Franz, Prep. Symp. ±
378 10 Microwave Catalysis in Organic Synthesis
Am. Chem. Soc., Div. Fuel Chem. 1999, 132 T. Hill, D.R. Weaver, Ger. Pat. Appl.
44, 133±139. DE 2535119, 1976.
130 D.D. Tanner, Q.Z. Ding, P. Kanda- 133 B. Herzog, Eur. Pat. Appl. EP 922,491,
narachchi, J.A. Franz, Int. Pat. Appl. 1999.
PCT 99/16851, 1999. 134 D.J. Harper, D.J. Wheeler, R.M. Hen-
131 W. Cho,Y. Baek, D. Park,Y.C. Kim, son, Eur. Pat. Appl. EP 742189, 1996.
M. Anpo, Res. Chem. Intermed. 1998,
24, 55±66.
379
11
Transition Metal Catalysis and Microwave Flash Heating in
Organic Chemistry
Kristofer Olofsson, Anders Hallberg, and Mats Larhed
11.1
Introduction
The combination of homogeneous catalysis and microwave heating is not only a hot
topic but also a research area that is likely to have an impact on several fields of modern
chemistry [1±11]. Transition metal-catalyzed reactions that typically needed hours or
days to reach completion with standard, thermal heating can now with high reproduc-
ibility be brought to full conversion in only seconds or minutes, consuming only a frac-
tion of the energy normally needed for a standard, oil-bath-heated reaction. What is
more, the application of microwave heating is in many cases associated with advan-
tages not easily achieved by the use of traditional heating. Indeed, transition metal-ca-
talyzed chemistry has on several occasions been proven to be even more efficient under
the action of microwaves than with standard heating [9]. In addition, the preparative
value of this heating method has increased with the attentive development of catalytic
systems that can withstand the sometimes extreme temperatures that are induced
under irradiation [12]. This chapter will discuss some of the factors which make the
combination of organometallic chemistry and microwave heating one of the most pro-
mising research areas in the field of modern synthetic chemistry.
We review herein progress in microwave-assisted homogeneous metal catalysis in
the liquid phase. Most of the work published in this field is, as the reader will see, as-
sociated with palladium chemistry, and this will, by necessity, also be reflected in the
text.
11.2
Safety
mended safety levels tend to move in the direction of lower doses of radiation, so
great care should be taken to minimize microwave leakage.
Although microwave-heated organic reactions can be smoothly conducted in open
vessels, it is often of interest to work with closed systems, especially if superheating
and high-pressure conditions are desired. When working under pressure it is
strongly recommended to use reactors equipped with efficient temperature feedback
coupled to the power control and/or to use pressure-relief devices in the reaction ves-
sels to avoid vessel rupture. Another potential hazard is the formation of electric arcs
in the cavity [2]. Closed vessels can be sealed under an inert gas atmosphere to re-
duce the risk of explosions.
11.3
Metal-catalyzed Reactions
The first examples of the use of palladium as a catalyst for carbon±carbon coupling
reactions were reported almost thirty years ago [14], and over recent decades a mas-
sive effort has been devoted to the extension of the scope of palladium-catalyzed reac-
tions. Organic and organometallic chemists have received extensive input from pal-
ladium-coordination chemistry in the task of understanding the mechanisms behind
these efficient synthetic procedures [14].
Palladium(0)-catalyzed coupling reactions ± i. e. the Heck and Sonogashira reac-
tions, the carbonylative coupling reactions, the Suzuki and Stille cross-coupling reac-
tions, and allylic substitutions (Fig. 11.1) ± have enabled the formation of many
kinds of carbon±carbon attachments that were previously very difficult to make.
These reactions are usually robust and occur in the presence of a wide variety of
functional groups. The reactions are, furthermore, autocatalytic (i. e. the substrate
regenerates the required oxidation state of the palladium) and a vast number of dif-
ferent ligands can be used to fine-tune the reactivity and selectivity of the reactions.
The long reaction times (ranging from hours to days) frequently required for full
conversions have, however, limited the exploitation of homogeneous catalysis in
R
Heck coupling R X +
R' R'
Carbonylative coupling R X + CO + Nu R CO Nu
X Nu
Allylic substitution + Nu
R R R R
Fig. 11.1 Examples of important palladium(0)-catalyzed transformations.
11.3 Metal-catalyzed Reactions 381
11.3.1
Heck Reactions
I
Pd(OAc)2
Et3N, (DMF)
Br Br
Yield
o
∆ 17 h, 100 C 64%
MW 4.8 min, 60 W 63%
Equation 11.1 Heck-coupling of 4-bromoiodobenzene and styrene.
β
Hydrolyzed
α
O product
OTf
Pd(OAc)2, DPPP O
Et3 N, DMF/H2O
t-Bu t-Bu
Yield Selectivity
MW 2.8 min, 55 W 77% α/β = 99/1
Equation 11.2 Heck-coupling of 4-tert-butylphenyl triflate and butyl vinyl
ether in the presence of water (DPPP = 1,3-bis(diphenylphosphino)propane).
β O
α NMe2
OTf O
Pd(OAc)2, PPh3 NMe 2
Et3N, DMF
Yield Selectivity
o
∆ 9 h, 60 C 93% α/β = 1/99
MW 7 min, 35 W 87% α/β = 1/99
Equation 11.3 Chelation-controlled Heck coupling of 2-naphthyl triflate.
O
I
Pd(OAc)2, PPh3 O
KOAc, Bu4NCl, DMF
Yield
o
∆ 24 h, 80 C 76%
MW 6 min, 30 W 58%
Equation 11.4 Heck-coupling of iodobenzene and 2,3-dihydrofuran.
γ
β SiMe3
OTf SiMe 3
Pd(OAc)2, DPPF
K 2CO3, CH3CN
Yield Selectivity
∆ 20 h, 80 oC 60% β/γ = 100/0
MW 5 min, 50 W (GC-MS Yield) 66% β/γ = 93/7
Equation 11.5 Internal arylation of allyltrimethylsilane (DPPF = 1,1-bis(di-
phenylphosphino)ferrocene).
γ
β NMe2
OTf NMe2
Pd(OAc)2, DPPF
K2CO3
Yield Selectivity
o
∆ 20 h, 80 C 71% β/γ = 99.5/0.5
MW 5 min, 15 W 42% β/γ = 99.5/0.5
Equation 11.6 Internal arylation of N,N-dimethylallylamine.
γ
OMe β NHBoc OMe
OTf NHBoc
Pd(OAc)2, DPPF
K2CO3
Yield Selectivity
o
∆ 20 h, 90 C 74% β/γ = 99.5/0.5
MW 6 min, 20 W 64% β/γ = 95/5
Equation 11.7 Internal arylation of the Boc-protected allylamine.
OMe O
I
Pd(PPh 3)2Cl2, TBAB OMe
K2CO3, H2O Yield
HOOC HOOC
MW 10 min, 375 W 92%
Equation 11.8 Arylation of methyl acrylate under phase-transfer conditions.
(TBAB = tetrabutyl ammonium bromide)
β
α
O
1. Pd(OAc)2, DPPP Hydrolyzed
Br product
K2CO3, H2 O, DMF O
2. HCl
Yield Selectivity
∆ 16 h, 80 oC 82% α/β = 99/1
MW 60 min, 122 oC 62% α/β = 99/1
O
HO O
HO
Br Pd(OAc)2, PPh3
i-Pr2EtN, H2O, DMF
Yield
∆ 16 h, 100 oC 66%
MW 10 min, 40 W 69%
Equation 11.10 Arylation of the enol form of 1,2-cyclohexanedione.
O
EtO O
EtO
Br Pd(OAc)2, PPh3
i-Pr2EtN, H2O, DMF
Yield
o
∆ 16 h, 100 C 83%
MW 12 min, 40 W 57%
Equation 11.11 Arylation of the ethyl vinyl ether of 1,2-cyclo-
hexanedione.
Traditional Heck arylation of the corresponding ethyl vinyl ether afforded high yields
with most of the aryl bromides investigated (Eq. 11.11). Under continuous single-
mode microwave treatment the transformations were complete within 10±12 min [25].
Heck reactions without solvent in a domestic microwave oven have been examined
by DÌaz-Ortiz [26]. The reactions were conducted in closed vessels with reported tem-
peratures of 150 8C. A study was performed in which reactions performed with mi-
crowave irradiation were compared with oil-bath-heated reactions with identical reac-
tion times and temperatures. The isolated yields tended to substantially favor the mi-
crowave-heated reactions (Eq. 11.12).
386 11 Transition Metal Catalysis and Microwave Flash Heating in Organic Chemistry
Ph
N Ph
N N
N
Br
Pd(OAc)2, P(o-tol)3
Et3N
Yield
o
∆ 22 min, 160 C 39%
MW 22 min, 450 W 78%
Equation 11.12 Heck reaction with phenyl bromide.
O α
O O
β
Pd(OAc)2, DPPP
Et3N, DMSO
O
TfO
Yield Selectivity
o
∆ 6 h, 60 C 79% α/β = 99/1
MW 5 min, 5 W 64% α/β = 99/1
Equation 11.13 Internal vinylation of butyl vinyl ether with a vinyl triflate.
O
α
O β O
HO
Pd(OAc)2, DPPP
Et3N, DMSO
O
TfO O
Yield Selectivity
o
∆ 20 h, 40 C 59% α/β = 99/1
MW 7 min, 5 W 53% α/β = 99/1
Equation 11.14 Internal vinylation of 2-hydroxyethyl vinyl ether with a vinyl triflate.
230 25 W x 5 min
20 W x 7 min
200
170
T/ oC
140 10 W x 5 min
110
80
50
5 W x 5 min
20
0 1 2 3 4 5 6 7 8 9
t/min
Fig. 11.2 Selected temperature profiles for the DMSO at different power inputs (fluoroptic ther-
single-mode heating of internal vinylations in mometer, Smith prototype).
Exchange of the butyl vinyl ether for 2-hydroxyethyl vinyl ether enabled the facile
transformation of vinyl triflates or bromides into protected a,b-unsaturated methyl
ketones (Eq. 11.14) [27]. One interesting aspect of this reaction is that a masked
methyl ketone can easily be introduced into a structure even in the presence of other
free ketone groups.
Selected temperature profiles for the reactions represented in Eqs (11.13) and
(11.14) are shown in Fig. 11.2. Note that the high tan d value of DMSO (0.825) [5]
results in extensive heating even at a relatively modest power input of single-mode
irradiation.
11.3.2
Carbonylative Couplings
O
Br
BINAP/Herrmann's catalyst N Bu
O + n-BuNH2 O H
Mo(CO)6, K2CO 3 (aq), Diglyme
Yield
o
MW 15 min, 150 C 79%
O
I Pd/(C), Mo(CO) 6, K2CO 3 (aq) OH
Diglyme/Ethylene glycol
Me Me Yield
o
MW 15 min, 150 C 87%
mately 150 8C. This enables the in situ generation of carbon monoxide directly in the
reaction mixture, without the need for external devices.
A variety of amides has been formed in moderate to high yields with aryl bromides
or iodides as aryl precursors and single-mode microwave heating for 15 min at
150 8C (Eq. 11.15) [29]. Under these conditions, aliphatic, unhindered primary, and
secondary amines reacted readily, whereas sterically hindered amines or amines of
low nucleophilicity, e. g. anilines, afforded low yields and incomplete conversions.
Among the homogeneous catalytic systems tested the most suitable for the use with
aryl bromides was a 2 : 1 mixture of BINAP and Herrmann's palladacycle.
Addition of ethylene glycol as cosolvent, as in Eq. (11.16), resulted in efficient for-
mation of the corresponding benzoic acid, instead of the amide [29].
11.3.3
Sonogashira Coupling Reactions
SiMe3
OTf
Pd(PPh 3)2Cl2, CuI
+ H SiMe3
NC Et2NH, DMF, LiCl NMR Yield
NC
o
MW 5 min, 120 C 99%
CuI, PPh3, K2 CO 3 OH
I + H
OH DMF Yield
MW 10 min, 375 W 80%
O O
Cl CuI, Et3 N
+ Ph
Ph
Br Br Yield
MW 10 min, 525 W 93%
used, because the cheaper CuI could be employed with reasonable yields. The cop-
per-catalyzed aroylation reactions were performed in triethylamine and in the ab-
sence of any added solvent (Eq. 11.19) [33]
11.3.4
Cross-coupling Reactions
This modification proved generally applicable and the Suzuki reaction is arguably
the most versatile of modern cross-coupling reactions. The reaction has, for example,
attracted the interest of groups involved in high-throughput chemistry, because a
large variety of boronic acids are commercially available.
The first microwave-promoted Suzuki couplings were reported in 1996 (Eq. 11.20)
[17]. Phenyl boronic acid was coupled with 4-bromotoluene to give a fair yield of pro-
duct after a reaction time of less than 3 min under single-mode irradiation. The
same reaction had previously been conducted with a reported reaction time of 4 h.
Equation (11.21) is an example of a single-mode microwave-promoted Suzuki cou-
pling, delivering a cyclic HIV-protease inhibitor after a subsequent hydrolysis of the
ketal group [35]. Half a dozen compounds were synthesized with the help of micro-
waves and some of the compounds had Ki values in the nanomolar range.
Interestingly, the Suzuki reaction worked smoothly on solid supports and high
yields of a variety of products were reported under these reaction conditions
(Eq. 11.22) [36]. 4-Bromo- and 4-iodobenzoic acid linked to Rink-amide TentaGel re-
Br
Pd(PPh 3)4, EtOH
+
DME, H2O Yield
Me (HO)2B
Me 55%
MW 2.8 min, 55 W
Equation 11.20 Suzuki coupling of phenyl boronic acid with 4-bromotoluene.
Br S S
Br
S B(OH)2
O O O O
S S
N N N N
PhO OPh Pd(PPh3 )4 PhO OPh
Yield
O O O O
MW 4 min, 45 W 48%
(HO)2B
N
SO2Ph
1. Pd(PPh 3)4, Na2CO3 (aq)
H EtOH, DME
RAM N H2N
MW 3.8 min, 45 W
Br
O 2. TFA N
O SO2Ph
Yield
88%
Equation 11.22 Suzuki coupling on a solid support.
11.3 Metal-catalyzed Reactions 391
O
S Pd(OAc)2 , K2CO3
O Br +
PEG
H2O
(HO)2B OMe MW 2 min, 75 W
O
S
PEG O
OMe
Equation 11.23 PEG-supported, aqueous Suzuki coupling.
Me
I Me
PEG-400, PdCl2, KF
+
OH (HO)2B Yield
MW 50 s, 240 W OH 79%
Equation 11.24 Suzuki coupling in PEG as a nontoxic reaction medium.
sulted in more than 99 % conversion of the starting material, and the whole reaction
was brought to completion within 4 min. The high yields suggest the high potential
of the use of microwave-assisted reactions on polymer supports [37].
Another interesting report by Kænig deals with rapid parallel Suzuki reactions in
water with phase-transfer catalysts [38]. The polymeric support used in these reac-
tions was PEG and a variety of aryl-palladium precursors was evaluated (halides, tri-
flates, and nonaflates, Eq. 11.23). The soluble polymer-supported liquid phase strat-
egy with PEG is appealing because, apart from facilitating the purification proce-
dure, PEG also helps solubilize the reagents and it has been suggested that PEG sta-
bilizes the palladium catalyst in the absence of phosphine ligands. The reactions ty-
pically reached full conversion in only 2 min of 75 W heating in a multimode oven,
whereas the parallel reactions needed 4 min to reach completion. The standard heat-
ing techniques required 2 h for the same reactions. The polymers and the esters
were both reported to withstand 10 min of 900 W microwave heating whereas con-
ventional thermal conditions induced substantial (up to 45 %) ester cleavage [38].
Similar Suzuki couplings performed with PEG as a nontoxic reaction medium
have been reported by Varma (Eq. 11.24) [39]. A domestic microwave oven with inver-
ter technology was used for these experiments and it was suggested that this technol-
ogy enabled more realistic control of the microwave power input.
Villemin has reported the use of sodium tetraphenylborate as a stable reactant for
Suzuki couplings performed in water or N-methylformamide (NMF) as solvents
(Eq. 11.25) [40].
A series of potent, linear C2-symmetric HIV-1 protease inhibitors with Ki values in
the nanomolar range was prepared from a diaryl bromide precursor emanating from
a carbohydrate scaffold, by application of Heck, Suzuki, Stille, and cyanation reac-
tions. Included in this series was the first reported microwave-promoted Suzuki
coupling with an alkyl borane [41]. A very high-yielding Suzuki coupling is presented
392 11 Transition Metal Catalysis and Microwave Flash Heating in Organic Chemistry
O O
I
HN Pd(OAc)2, Na2CO3 HN
NaBPh4, CH3NHCHO
O N O N Yield
H H
MW 8 min, 100 W 70%
Br
S
(HO)2B
H O OH O H O
Pd(PPh3)4 , Na2CO3
N N S
N N DME, EtOH, H 2O
O H O OH O H MW 4 min, 45 W
H O OH O H O
Br N N
N N
O H O OH O H
Yield
96%
S
in Eq. (11.26) and demonstrates that even relatively complex structures, such as these
peptidomimetics, do not decompose or are in any other way unsuitable for applica-
tions with flash heating chemistry. These syntheses of HIV-protease inhibitors were
reported to be the first examples in medicinal chemistry of the application of homo-
geneous palladium-catalyzed reactions conducted with flash heating [41].
action times was easily applied to Stille reactions in solution (Eq. 11.27) [17] and on
polymer supports (Eq. 11.28) [36].
We [43] and, more recently, Maleczka [44] have reported one-pot microwave-as-
sisted hydrostannylations and Stille coupling sequences. Eq. (11.29) depicts one of
Maleczka's reactions (see also Eq. 11.32).
Some of the disadvantages of the Stille reaction, e. g. the low reactivity of some sub-
strates, separation difficulties in chromatography, and the toxicity of tin compounds,
have been ameliorated by recent efforts to improve the procedure. Curran has, in a ser-
ies of papers, reported the development of the concept of fluorous chemistry, in which
the special solubility properties of perfluorinated or partly fluorinated reagents and
solvents are put to good use [45]. In short, fluorinated solvents are well known for their
insolubility in standard organic solvents or water. If a compound contains a sufficient
number of fluorine atoms it will partition to the fluorous phase, if such a phase is pre-
sent. An extraction procedure would thus give rise to a three-phase solution enabling
ready separation of fluorinated from nonfluorinated compounds.
Among the many applications of fluorous chemistry is the Stille coupling of tin re-
agents with fluorinated tags in which the products and excess of the tin-containing
reagents can be conveniently removed from the reaction mixture, and recycled. Un-
OTf
Pd2dba3, Ph3As
+
O LiCl, NMP
Bu 3Sn O Yield
∆ 70 h, rt 82%
MW 2.8 min, 50 W 68%
Equation 11.27 Stille coupling in solution with 4-acetylphenyl triflate.
Bu 3Sn
H 1. Pd 2dba3, Ph3As
RAM N MW 3.8 min, 40 W H 2N
I
O 2. TFA O
Yield
85%
Equation 11.28 Stille coupling utilizing a RAM-linker on a polymer support.
O
OTf O PdCl2(PPh3)2
+ Sn(CH2CH2 C6F13)3
O LiCl, DMF O
fortunately, fluorous Stille reactions require long reaction times, typically one day at
80 8C and the solubilities of fluorous compounds in most of the common solvents
used in synthesis are low, which limits the scope of the procedure.
The long reaction times under the action of classic heating were reduced to only a
few minutes with single-mode heating at a power of 50±70 W. One example of the
use of CH2CH2C6F13 (denoted F-13) -tagged organostannanes is presented in
Eq. (11.30) [46].
Occasionally it was apparent that the fluoricity provided with the F-13 tags was not
sufficient to enable full partitioning of the products to the fluorous phase. The use
of more heavily fluorinated tags, e. g. the ±CH2CH2C10F21 (F-21) tag, was investi-
gated, but proved to be preparatively elusive, because the solubility of these com-
pounds was very poor. Heating the reactions to 80 8C in fluorinated solvents resulted
in very sluggish and irreproducible reactions. The use of single-mode heating en-
abled rapid and efficient reactions (Eq. 11.31) [43]. Superheating of the solvent made
solution of the reactants possible and a one-phase system was generated by the heat-
ing procedure. On cooling, the fluorous reagents and products separated from the
mixture again and separation of fluorous from nonfluorous compounds ensued
without complications. The use of highly fluorinated tags could be said to combine
the attractive features of both solid- and liquid phase chemistry, because the ease of
separation is reminiscent of the former and the versatility and reactivity of the
soluble reagents is typical of the latter.
The fluorous Stille procedure has also been applied to a one-pot hydrostannylation
of an acetylenic compound in the hybrid fluorous/organic solvent benzotrifluoride
(BTF). Subsequent Stille coupling of the product in BTF/DMF, as shown in Eq. (11.32),
was easily achieved in 8 min with 60 W single-mode microwave irradiation [43].
To take advantage of the fluorous extraction procedure, or the use of the prepara-
tively simple filtration, the use of highly fluorous reagents is crucial. The usefulness
of the microwave heating technique is obvious in these examples.
11.3 Metal-catalyzed Reactions 395
1. AIBN, BTF O
O O Ph OEt
MW 10 min, 60 W
+ HSn(CH2CH2C10F21)3
EtO OEt 2. PhI, Pd(OAc)2, CuO
EtO E/Z = 5:1
DMF/BTF
O
MW 8 min, 60 W Yield
44%
Equation 11.32 One-pot hydrostannylation and Stille reaction with
F-21-tagged reagents.
11.3.5
Other Palladium-catalyzed Reactions
The synthesis of nitriles from halides is valuable in medicinal chemistry because nitriles
are flexible building blocks readily converted into carboxylic acids, amides, amines, or a
variety of heterocycles, e. g. thiazoles, oxazolidones, triazoles, and tetrazoles. The impor-
tance of the tetrazole group in medicinal chemistry is easily understood if we consider
that it is the most commonly used bioisostere of the carboxyl group.
An improvement of the palladium-catalyzed cyanation of aryl bromides, in which
zinc cyanide was used as the cyanide source, was reported in the middle of the nine-
ties [49]. Typically, the conversion from halide to nitrile takes at least 5 h by this route
and the subsequent cycloaddition to the tetrazole is known to require even longer re-
action times.
A single-mode microwave procedure has been reported for the palladium-cata-
lyzed preparation of both aryl and vinyl nitriles from the corresponding bromides.
The reaction times were short and full conversions were achieved in just a few min-
utes (Eq. 11.33) [50]. The cycloadditions to yield the tetrazoles needed slightly longer
reaction times, from 10 to 25 min, but only 20 W of power was required as a tem-
perature of 220 8C was reached after 10 min heating. The yields in this step ranged
from 36 % to 96 %. This method for transforming halides into tetrazoles has been
used for the synthesis of a novel HIV-protease inhibitor [50].
The coupling and cycloaddition could also be achieved as a one-pot procedure on a
polymer support, as shown in Eq. (11.34), in which a Rink linker on TentaGel was
used. Negligible decomposition of the solid support was reported [50].
Br Zn(CN)2, Pd(PPh3)4 CN
DMF
N N Yield
Equation 11.33 Palladium-cata-
o
lyzed conversion of aryl bromides ∆ 4 h, 97 C 71%
into aryl nitriles. MW 2 min, 60 W 88%
396 11 Transition Metal Catalysis and Microwave Flash Heating in Organic Chemistry
O O O
I O P(OEt) 2
MeO PdCl2(PPh3) 2, HSiEt3 MeO
+ HP(OEt)2
Et 3N, Toluene Yield
OMe
OMe OMe OMe
I
PdCl2(PPh3)2, NaOAc
DMA
The conversion of aryl iodides into aryl phosphonates, a useful precursor to aryl
phosphonic acids, was performed in a Teflon autoclave by Villemin [51]. A domestic
microwave oven was used for these experiments and the reaction times for classic
heating were effectively reduced from 10 h to 4±22 min. The reactivity of iodides
was good whereas the use of bromides resulted in lower yields and reactions with tri-
flates were very slow (Eq. 11.35). It is notable that the reactions were brought to com-
pletion with short reaction times in a non-polar solvent.
The synthesis of kinafluorenones, which are of interest for the synthesis of Strepto-
myces-related antibiotics, has been investigated by Jones [52]. The palladium-
mediated ring closure depicted in Eq. (11.36) resulted in poor yields under conven-
tional heating but irradiation with a domestic microwave oven for 60 s improved the
yield substantially.
11.3.6
Asymmetric Catalysis
has been shown that microwave heating can dramatically reduce the reaction times
in asymmetric catalysis without significantly affecting ee values [54].
This particular field of palladium chemistry has been responsible for attempts to
develop temperature-stable catalyst systems, sometimes with very fascinating re-
sults. The high temperatures of microwave synthesis can be a double-edged sword ±
the high temperatures result in high reaction rates but also increase the risk of cata-
lyst breakdown. It is often the metal ligands that are temperature sensitive, and be-
cause these ligands induce the asymmetry in the reactions, catalyst breakdown is as-
sociated with loss of ee. The development of temperature-stable ligands for asym-
metric reactions has thus been of great importance.
A highly thermostable palladium±phosphine oxazoline catalytic system, shown in
Eq. (11.37), has recently been reported to yield high ee under single-mode micro-
wave irradiation [54, 55]. The use of this P,N-ligand catalytic system resulted in even
higher ee than the P,P-ligand BINAP [56]. The reactions were performed in acetoni-
trile (b.p. 81±82 8C) and superheating increased the temperatures up to 145 8C, as
measured by means of a fluoroptic probe.
The palladium-catalyzed substitution of the less reactive racemic ethyl 3-cyclohexe-
nyl carbonate could, in a similar fashion, be completed with dimethyl malonate,
p-methoxyphenol, or phthalimide as nucleophiles, with satisfactory ee (Eq. 11.38)
[55]. These reactions, when irradiated for 1 min with temperatures up to 100 8C, de-
livered yields (91±96 %) and ee values (94±95 %) identical with those performed in
O O O
(Pd(η3-C3H5)µ-Cl)2
Dimethyl malonate
BSA, CH3CN HPLC Yield ee
MW 30 s, 120 W
98% >99%
Equation 11.37 Asymmetric palladium-catalyzed allylic alkylation
(BSA = N,O-bis(trimethylsilyl)acetamide).
O O
NH HN
PPh2 PPh 2 OMe
OMe
O OEt + O
HO (Pd(η3-C3H5)µ-Cl)2
O Yield ee
MW 1 min, 120 W 96% 95%
closed vessels in an oil bath at 100 8C for 5 min. A related experiment, conducted in
acetonitrile with phthalimide as nucleophile in which the temperature was increased
to 140 8C, resulted in much higher yields when microwave heating was used,
although the enantiomeric excesses remained the same (95±96 %) [55].
The molybdenum-catalyzed asymmetric reaction differs from the palladium-cata-
lyzed reaction in several ways, the most important of which is the different regios-
electivity achieved. Molybdenum-catalyzed reactions favor the most sterically hin-
dered position (Eq. 11.39), in contrast with palladium catalysis. The molybdenum-
catalyzed allylations also suffer from significantly lower reactivity.
It was found by Trost that the low reactivity could be circumvented by the employ-
ment of labile ligands, such as the propionitrile in the Mo(CO)3(EtCN)3 precatalyst
[57]. Instead of directly transferring this procedure to microwave heating applica-
tions, a useful and easily handled microwave procedure was developed for rapid and
selective molybdenum-catalyzed allylic alkylations under noninert conditions
(Eq. 11.39) [12]. The former, more sensitive, two-step reaction was fine-tuned into a
robust one-step procedure employing the inexpensive and stable precatalyst
Mo(CO)6, used in low concentrations. The alkylations were conducted in air and re-
sulted in complete conversions, high yields, and an impressive enantiomeric excess
(98 %) in only 5±6 min. Despite the daunting temperatures, up to 250 8C with THF
O O
NH HN O O
O N N MeO OMe
O OMe
Mo(CO)6 , THF, BSA
Dimethyl malonate Yield ee
O O
NH HN O O
O N N MeO OMe
as solvent (b.p. 65±67 8C), the high enantiomeric purity remained constant, which
suggests that the coordination of the ligand to the metal is not broken, despite of
these stressing reaction conditions [12].
After preparation of several different chiral trans-1,2-bis-(2-carboxyamidopyridine)
cyclohexane ligands, a 4-methoxypyridine derivative (Eq. 11.40) was discovered to
have superior qualities and to deliver a high yield (88 %) of the branched product
with >99 % ee and a regioisomer ratio of 41 : 1 (alkylation at the most, rather than
least, substituted position) [58]. Single-mode microwave irradiation was applied at
the very high power of 200 W for 4 min and superheating caused THF to reach tem-
peratures of 150±180 8C.
11.3.7
Other Metal-catalyzed Reactions
SiMeCl2
MeSiCl2H, CuCl
N TMEDA N
MW 6 x 30 s, 750 W
Cl O
O CCl2COEt
Cu(2-propylamine)2Cl2
+ CCl3COEt
CH3CN
MW 600 W
Equation 11.42 Addition of ethyltrichloroacetate to styrene.
400 11 Transition Metal Catalysis and Microwave Flash Heating in Organic Chemistry
O
O
H RuHCl(CO)(PPh3 )3 OH
+
H OH MW
Equation 11.43 Transfer hydrogenation of benzaldehyde with ruthenium catalysis.
O
1. Cu(OAc)2
N (HO)2B O N
N Molecular sieves
+
H N Pyridine/NMP H2N N
H MW 3 x 10 s, 1000 W
2. TFA:CH 2Cl2 Yield
56%
1:1 mixture
of isomers
Equation 11.44 Solid support aryl/heteroaryl C±N coupling reactions.
Ts
N
O Br CuBr2, CH3CN
+ S N
Na Yield
O
MW 12 min 88%
Equation 11.45 Transition metal-catalyzed aziridation of olefins.
References 401
11.4
Summary
The development in microwave chemistry has been remarkable during the last few
years; from the first reports in which, typically, domestic ovens were used to modern
applications with state-of-the-art single-mode cavities. We believe that it is today pos-
sible today to develop robust microwave-assisted methods for nearly any reaction
that needs an external heat source and we have proved it is possible to perform tran-
sition-metal catalyzed reactions very cleanly and selectively.
Heating by means of single-mode microwave irradiation enables readily adjustable
and controlled bulk heating which can be performed safely and with very low energy
consumption. The synthetic chemist of today can take advantage of the unique car-
bon±carbon bond formation reactions afforded by organometallic chemistry and
make the reaction happen in seconds or minutes, an important feat, because many
transition-metal-catalyzed reactions are time-consuming.
The examples presented indicate that the combined approach of microwave heat-
ing and homogeneous catalysis can be an almost synergistic strategy, in the sense
that the combination in itself has more potential than its two separate parts in isola-
tion. There are, furthermore, still many other catalytic reactions with great potential
for microwave heating [66]. It might, for example, be expected that increasing num-
bers of aqueous biocatalytic reactions might respond well to microwave heating. It is
already clear that modern automatic microwave synthesizers have much to offer in
the high-speed generation of combinatorial libraries.
Acknowledgment
We thank the Swedish Natural Science Research Council, the Swedish Foundation
for Strategic Research, Knut and Alice Wallenberg's Foundation, PersonalChemistry
AB, Medivir AB and Gunnar Wikman.
References
12
Microwave-Assisted Combinatorial Chemistry
C. Oliver Kappe and Alexander Stadler
12.1
Introduction
Combinatorial chemistry, the art and science of synthesizing and testing compounds
for bioactivity en masse, has emerged as one of the most promising approaches to drug
discovery [1±5]. Because of the enormous progress made in genomic sciences, molecu-
lar biology, and biochemistry, a large number of biologically important target proteins
has now become available for screening purposes. This has led to an ever-growing de-
mand for large libraries of novel compounds that are being evaluated for their biological
properties by use of appropriate screening procedures. The discovery of novel biological
targets was paralleled by the development of novel assays and high-throughput screen-
ing (HTS) technologies including miniaturized formats, enabling the testing of thou-
sands of individual compounds per day. Traditional methods of organic synthesis are or-
ders of magnitude too slow to satisfy the increasing demand for these compounds.
Combinatorial chemistry officially dates back to the mid-1980s and since then has
experienced enormous growth and has steadily attracted the interest of researchers
from a variety of different fields. The essence of combinatorial synthesis is the ability
to generate large numbers of chemical compounds very quickly [1±5]. Chemistry in
the past has been characterized by slow, steady, and painstaking work; combinatorial
chemistry has broken many of the preconceptions and resulted in a level of chemical
productivity thought impossible a few years ago. In the past chemists made one com-
pound at a time, in one reaction at a time. For example, compound A would have
been reacted with compound B to give product AB, which would have been isolated
after reaction, workup, and purification by chromatography. In contrast with this ap-
proach combinatorial chemistry offers the potential to make every combination of
compound A1 to An with compound B1 to Bn (Fig. 12.1). The range of combinatorial
techniques is highly diverse and the products could be made individually in a parallel
fashion or in mixtures, by use of either solution or solid-phase techniques [1±6].
Parallel to these developments in combinatorial chemistry, microwave-enhanced
organic synthesis has attracted a substantial amount of attention in recent years. As
evidenced from the other chapters in this book and the large number of review arti-
cles available on this subject [7±14], high-speed microwave-assisted synthesis has
A1 B1
A2 B2
A3 B3
A4 B4 A1-nB1-n
A+B AB
. .
. .
An Bn
(A) (B)
Fig. 12.1 (A) In a conventional synth- ferent building blocks of type A1 ±An are
esis, one starting material A reacts with treated with different building blocks of
one reagent B resulting in one product type B1±n according to combinatorial
AB. (B). In a combinatorial synthesis, dif- principles.
been applied successfully in many fields of synthetic organic chemistry. It is, in fact,
becoming evident that microwave approaches can probably be developed for most
chemical transformations requiring heat, and/or for processes in which small polar
molecules are removed from a chemical equilibrium. Although different hypotheses
have been proposed to account for the rate-enhancements observed under micro-
wave irradiation, a generally accepted rationalization remains elusive (Chapt. 3) [13,
14]. Irrespective of the origin and/or existence of a special microwave effect, rapid
microwave dielectric heating (microwave flash heating) is extremely efficient and ap-
plicable to a broad range of practical synthesis [7±12].
The main benefits of performing reactions under microwave irradiation condi-
tions are the significant rate-enhancements and the higher product yields that can
frequently be observed. Not surprisingly, these features have recently also attracted
interest from the combinatorial and/or medicinal chemistry community, for whom
reaction speed is of great importance [15±18]. Combination of microwave heating
technology and combinatorial chemistry applications is, therefore, a logical conse-
quence of the increased speed and effectiveness afforded by microwave heating. In
that respect, all chemical transformations that can be speeded up by microwave irra-
diation should, by definition, be of interest to practitioners in the field of combinator-
ial and medicinal chemistry [15±18].
This attractive linking of combinatorial processing with microwave heating was re-
cognized in a 1995 summary of microwave-assisted organic reactions [8]. This review
is limited to microwave-assisted processes and techniques that are of direct relevance
to combinatorial chemistry applications, including solid-phase organic synthesis, the
use of polymer-bound scavengers or reagents, synthesis on soluble polymer sup-
ports, fluorous phase techniques, parallel synthesis, and the construction of libraries
in automated format by use of microwave technology. Because some of these applica-
tions require specialized equipment, a section will be devoted to microwave reactor
technology for high-throughput synthesis. In many of the examples of microwave
chemistry presented in this review we mention reported ªrate-enhancementsº in the
microwave-heated reactions compared with conventional heating. The reader should
be aware that such comparisons are inherently troublesome (Chapt. 3) [13, 14] and
do not a priori imply the existence of nonthermal or specific microwave effects.
12.2 Solid-phase Organic Synthesis 407
12.2
Solid-Phase Organic Synthesis
One of the cornerstones of combinatorial synthesis has been the development of so-
lid-phase organic synthesis (SPOS) based on the original Merrifield method for pep-
tide preparation [19]. Because transformations on insoluble polymer supports should
enable chemical reactions to be driven to completion and enable simple product pur-
ification by filtration, combinatorial chemistry has been primarily performed by
SPOS [19±23]. Nonetheless, solid-phase synthesis has several shortcomings, because
of the nature of heterogeneous reaction conditions. Nonlinear kinetic behavior, slow
reaction, solvation problems, and degradation of the polymer support, because of the
long reactions, are some of the problems typically experienced in SPOS. It is, there-
fore, not surprising that the first applications of microwave-assisted solid-phase
synthesis were reported as early 1992 [24].
The earliest published example of microwave-assisted SPOS involved diisopropyl-
carbodiimide (DIC)-mediated solid-phase peptide couplings [24]. Numerous Fmoc-
protected amino acids and peptide fragments were coupled with glycine-preloaded
polystyrene Wang resin (PS-Wang) in DMF, using either the symmetric anhydride or
preformed N-hydroxybenzotriazole active esters (HOBt) as precursors (Scheme 12.1).
Reactions were performed in an unmodified domestic microwave oven employing
a custom-made solid-phase reaction vessel under atmospheric pressure. Efficiencies
were significantly improved by use of microwave irradiation for all the peptide cou-
plings tested, the rate enhancement being at least 2±3-fold compared with conven-
tional couplings at room temperature. In general, peptide bond formations were
completed within 2±6 min, compared to 30 min for reactions without microwave ir-
radiation at room temperature. As with so many of the early publications in micro-
wave-assisted chemistry, the exact reaction temperature during the irradiation period
was not determined, presumably because of lack of suitable instrumentation
(although the temperature at the end of the reaction could have been measured).
The reasons for the observed rate-enhancements, and the possible involvement of
so-called nonthermal microwave effects (Chapt. 3) [13, 14] therefore remain unclear.
In a more recent study using dedicated multimode microwave reactors for chemical
synthesis, which enable temperature and power control, it was demonstrated that mi-
crowave irradiation could be effectively employed to couple aromatic carboxylic acids
to polystyrene Wang resin [25], if the symmetrical anhydride procedure was used, and
not the three-component O-acylisourea activation method [19]. Almost quantitative
loading was achieved in 1-methyl-2-pyrrolidone (NMP) at 200 8C within 10 min under
(PhCO)2O, O
a) OH MW, 10 min
O Ph
PS-Wang NMP, 200 °C
RCO2 H, Cs2CO3 O
b) Cl MW, 5-15 min
O R
PS-Merrifield/Wang NMP, 200 °C
33 examples
Scheme 12.2 Attachment of carboxylic acids to Wang and Merrifield resins.
I Ph
H Bu3SnPh, [Pd] H
N MW, 3.8 min N
a)
O NMP
O
TentaGel-Rink
X Ar
H ArB(OH)2, Na2CO3, [Pd] H
N MW, 3.8 min
N
b)
O X = Br, I DME, EtOH, H2O
O 14 examples
TentaGel-Rink
1. Zn(CN)2, [Pd] N
N
MW, 2 min N
I
H 2. NaN3, NH4Cl H N
MW, 20 min H
N N
c)
NMP
O O
TentaGel-Rink
p-Tol
H
H N H N
N p-TolB(OH)2, Cu(OAc)2, [Pd] N
N MW, 0.5 min N
d)
pyridine-NMP
O O
TentaGel-PAL
proved by use of microwave irradiation (e. g. Scheme 12.3 d) [29]. These experiments
were performed with a domestic microwave oven and use of several intermittent ir-
radiation cycles and multiple additions of excess reagent. Compared with the con-
ventional thermal procedure (80 8C) the reaction time was effectively reduced from
48 h (overall) to less than 5 min and afforded a variety of N-arylated heterocycles in
high yield and purity after cleavage from the solid support.
Another metal-catalyzed microwave-assisted transformation performed on a poly-
mer support involves the asymmetric allylic malonate alkylation reaction shown in
Scheme 12.4. The rapid molybdenum(0)-catalyzed process involving thermostable
chiral ligands proceeded with 99 % ee on a solid support. When TentaGel was used
as as support, however, the yields after cleavage were low (8±34 %) compared with
the corresponding solution phase microwave-assisted process (monomode cavity)
which generally proceeded in high yields (>85 %) [30].
Multicomponent reactions (MCR), in which three or more reactions combine to
give a single product, have lately received much attention. The Ugi four-component
condensation in which an amine, an aldehyde or ketone, a carboxylic acid, and an iso-
cyanide combine to yield an a-acylamino amide, is particularly interesting, because
410 12 Microwave-assisted Combinatorial Chemistry
O
O O O O
Ph O OMe
O OMe Mo(CO)6, ligand O OMe
MW, 10 min
TentaGel-OH Ph
THF
Scheme 12.4 Solid-phase molybdenum(0)-catalyzed allylic alkylation.
of the wide range of products obtainable by variation of the starting materials [31].
Although conventional solid-phase Ugi reactions require several days for completion,
it has been demonstrated that by use of microwave irradiation (monomode instru-
ment, sealed Pyrex vessel), products of high purity were obtained in moderate to ex-
cellent yields (24±96% yield) within 5 min irradiation time (Scheme 12.5) [32].
A polymer-bound amino component was used and a mixture of dichloromethane
and methanol (DCM±MeOH 2 : 1) as solvent which both absorbed microwave energy
and solvated the resin. A library comprising 18 members was obtained after cleavage
of the resin-bound material from the support with TFA : DCM 19 : 1.
Another condensation process that has been performed by microwave heating is
the Knoevenagel reaction (Scheme 12.6). Resin-bound nitroalkenes, for example,
R1CHO,
R2CO2H, R 3NC R1 H
MW, 5 min N
NH2 N R3
DCM/MeOH Scheme 12.5 Ugi 4-component
O
TentaGel-RAM O R2 reactions on a solid support.
O
NO2
HO
ArCHO, NH4OAc
DIC, HOBt, rt, 17 h O O
MW, 20 min
NO 2 NO2
THF O THF O
Ar
5 examples
OH
PS-Wang
O O
ArCHO
R1O R
O O piperidinium acetate O O
MW, 170 °C, 1-10 min MW, 125 °C, 60 min
O R O R
1,2-Dichlorobenzene 1,2-Dichlorobenzene
7 examples Ar
21 examples
Scheme 12.6 Knoevenagel condensations on solid support.
12.2 Solid-phase Organic Synthesis 411
R R2XH (X = O, NR) R
N N Cl MW, 6 min N N X
L L R2
N N NMP or DMSO N N
PP or cellulose
membrane X X
R1 R1
Scheme 12.7 High-speed nucleophilic substitution reactions on polypropylene
or cellulose membranes.
412 12 Microwave-assisted Combinatorial Chemistry
O ArNCO O O
a) MW, 12 min
H Ar
O N O N N
DCM
R R H
PS-Wang
5 examples
O O O
e.g.
O
O SiO2 /TaCl5 O
b) MW, 5-7 min
NH2 N
O O
no solvent
PS-Merrifield O
6 examples
O O O OH
MW, 4-6 min
c) O
O R
R DMF
PS-Merrifield
8 examples
Scheme 12.8 Miscellaneous solid-phase reactions performed under
the action of microwave irradiation.
12.2 Solid-phase Organic Synthesis 413
O TFA/DCM 1:1 O
MW, 120 °C (7 bar), 30 min
Scheme 12.9 Microwave- O Ph
assisted cleavage of benzoic HO Ph
acid from Merrifield resin. PS-Merrifield
O2 O2 O R1R2NH O
H H
a) S S N Ph MW ,~140 °C, 15 min R1 N Ph
NH2 N N
O DMSO R2 O
Me Me
PS NC
1) p-NO2C6H4OCOCl
S DCM, 0°C S
2) RNH2, DMF O
R
PS OH O N
H
R1R2NH O
MW, 1-20 min R2 R1
N N
NMP H
R1
Scheme 12.11 Microwave-assisted aminolysis (Marshall linker).
12.3
Polymer-supported Reagents, Scavengers, and Catalysts
Apart from traditional solid-phase organic synthesis (SPOS), the use of polymer-sup-
ported reagents (PSR) has gained increasing attention from practitioners in the field of
combinatorial chemistry [49±52]. The use of PSR combines the benefits of SPOS with
the advantages of solution-phase synthesis. The most important advantages of these re-
agents are simplification of reaction workup and product isolation, the workup being
reduced to simple filtration. PSR can also be used in excess without affecting the purifi-
cation step. By using this technique, reactions can be driven to completion more easily
than in conventional solution-phase chemistry. So far only a few applications of micro-
wave-assisted PSR strategies have been reported in the literature.
One recent process involves the use of a polymer-supported Burgess reagent for
the synthesis of 1,3,4-oxadiazoles from 1,2-diacylhydrazines (Scheme 12.12). Irradia-
tion of a variety of 1,2-diacylhydrazines with a polyethylene glycol (PEG 750)-sup-
ported Burgess reagent in THF provided the corresponding 1,3,4-oxadiazoles in
75±96 % yield and high purity after only 2±8 min irradiation [53]. Under conventional
reflux conditions a 40 % conversion was obtained after 3 h. Apart from filtration
through silica gel to remove the soluble polymer-supported reagent, and evaporation
of the solvent, no purification was necessary. Reactions were performed in dedicated
monomode instruments under sealed vessel conditions (no temperatures given). In
a variation of this procedure the same authors also described the cyclodehydration of
diacylhydrazines to oxadiazoles by use of the insoluble polystyrene-supported re-
agents A and B (Scheme 12.12) [54]. Similar reduced reaction times as with the solu-
ble polymer support (5±10 min) were also achieved by use of either of the two poly-
styrene-supported reagents in THF with microwave irradiation in sealed vessels.
Another example of microwave-assisted PSR chemistry involves the rapid conver-
sion of amides to thioamides by use of a polystyrene-supported Lawesson-type thio-
nating reagent. By use of microwave irradiation at 200 8C in sealed vessels (mono-
mode reactor), a range of secondary and tertiary amides was converted within
O2 O
S
Et3N N O
PEG R1 O R2
O O
R2 MW, 2-8 min
R1
N N N N
THF
H H 16 examples
O2 O Et2N NtBu
S Me P
N N O N N
Me2N PS PS
A B
Scheme 12.12 Synthesis of 1,3,4-oxadiazoles by use of polymer-supported
reagents.
416 12 Microwave-assisted Combinatorial Chemistry
O S
MW, 200 °C, 15 min
R NR1R2 R NR1R2
toluene, ionic liquid
9 examples
15 min to the corresponding thioamides, in high yield and purity (Scheme 12.13)
[55]. Compared with classical reflux conditions, these thionation reactions were
much faster and reaction times were reduced from 30 h to 10±15 min. Interestingly,
even heating at these elevated temperatures caused no damage to the polymeric sup-
port. Because toluene is not an optimum solvent for absorption and dissipation of
microwave energy [48], a small amount of ionic liquid (1-ethyl-3-methyl-1H-imidazo-
lium hexafluorophosphate) was added to the reaction mixture to ensure even and ef-
ficient distribution of heat.
In a related example, the catalytic transfer hydrogenation of olefinic substrates
using a polymer-supported hydrogen donor has been reported. By use of microwave
irradiation (monomode reactor) high yields of products (80±95 %) could be obtained
by irradiation of a mixture of an Amberlite-derived supported formate, Wilkinson's
catalyst (RhCl(PPh3)3), and the olefinic substrate in the minimum quantity of
DMSO (Scheme 12.14) [56]. After separation of the Amberlite at the end of the reac-
tion the polymer-supported formate salt could easily be regenerated and used in
further hydrogenation reactions. In total, five reaction±regeneration cycles were pos-
sible before an appreciable decrease in the reaction yield was noted.
Very recently a novel one-pot three-step Wittig reaction using microwave irradia-
tion and polymer-supported triphenylphosphine has been reported [57]. By use of
PS
N HCOO
RhCl(PPh 3)3
E MW, 30 s E Scheme 12.14 Microwave-medi-
R R ated hydrogenation using a polymer-
DMSO
supported hydrogen
8 examples donor.
PS Ph
P
Ph
O MW, 150 °C, 5 min R2
R1 + Br R2 R1
H MeOH, K2CO3
15 examples
12.4
Soluble Polymer-supported Synthesis
ArB(OH)2
Pd(OAc)2, K2CO3 O
O
MW, 2 min S
S Ar
O Br O
H2O
PEG
5 examples
ArB(OH)2
Pd(OAc)2, K2CO3 O
O
MW, 2-4 min
O O
H2O
PEG
X Ar
X = I, OTf, ONf 12 examples
O RX, Cs2CO3 O
MW, 30-60 min
N Ph N Ph
O no solvent O
PEG Ph R Ph
8 examples
Scheme 12.17 Microwave-assisted alkylations on PEG support.
O O
AlkO R
O O
MW, 10 min
O R Scheme 12.18 Derivatization of
OH no solvent
poly(styrene-co-allyl alcohol) with
R = Me, Ph, OEt b-dicarbonyl compounds.
12.5 Fluorous-phase Synthesis 419
O E2 E1
E2 E1
NaOMe, rt, 18 h
H
MeOH MeO
O
O
5 examples
12.5
Fluorous-phase Synthesis
[Pd], LiCl
MW, 1.5-2 min
Ara X + (C6F13CH2CH2)3Sn-Arb Ara Arb
DMF
X = Br, I, OTf (14 examples)
Scheme 12.20 Fluorous Stille couplings under the action of microwave irradiation.
12.6.
Parallel Synthesis
Parallel processing of synthetic operations has been one of the cornerstones in com-
binatorial chemistry for years [1±6]. In the parallel synthesis of combinatorial li-
braries, compounds are synthesized using ordered arrays of spatially separated reac-
tion vessels adhering to the traditional ªone vessel-one compoundº philosophy. The
defined location of the compound in the array provides the structure of the com-
pound. A commonly used format for parallel synthesis is the 96-well microtiter plate,
and today combinatorial libraries comprising hundreds to thousands of compounds
can be synthesized by parallel synthesis, often in an automated fashion [6].
As demonstrated in the previous sections of this review, microwave-assisted reac-
tions allow rapid product generation in high yield under uniform conditions. There-
fore, they should be ideally suited for parallel synthesis and/or combinatorial chem-
istry applications. The first example of parallel reactions performed under micro-
wave irradiation conditions involved the nucleophilic substitution of an alkyl iodide
with 60 diverse piperidine or piperazine derivatives (Scheme 12.22) [71]. Reactions
were performed in a multimode microwave reactor in individual sealed polypropy-
12.6. Parallel Synthesis 421
X
I X NH
N
N MW, 4 h
NH2 N
S MeCN X = NR, CR1R2 NH2
S
60 examples
Scheme 12.22 Nucleophilic substitution reactions performed in parallel.
O O Ar O
Ar bentonite clay
R2 MW, 5 min R2 R2
+ + NH4NO3
O H no solvent
R1 O R1 N R1
> 96 examples
Scheme 12.23 Microwave-assisted Hantzsch pyridine synthesis.
lene vials using acetonitrile as solvent (no further details given). Screening of the re-
sulting 2-aminothiazole library in a herpes simplex virus-1 (HSV-1) assay led to three
confirmed hits, demonstrating the potential of this method for rapid lead optimiza-
tion.
In a key 1998 publication, the concept of microwave-assisted combinatorial chem-
istry was introduced for the first time. Using the three-component Hantzsch pyri-
dine synthesis as a model reaction, libraries of substituted pyridines were prepared
in a high-throughput parallel fashion. In this variation of the Hantzsch multicompo-
nent reaction, ammonium nitrate was used as the ammonium source as well as oxi-
dizing agent, employing bentonite clay as inorganic support (Scheme 12.23) [72]. Mi-
crowave irradiation was performed in 96-well filter-bottom polypropylene plates, in
which the corresponding eight 1,3-dicarbonyl compounds and twelve aldehyde build-
ing blocks were dispensed using a robotic liquid handler. Microwave irradiation of
the 96-well plate containing aldehydes, 1,3-dicarbonyl compounds, and ammonium
nitrate±clay in a domestic microwave oven for 5 min produced the expected pyridine
library directly, after the desired products were extracted from the solid support by
organic solvent and collected in a receiving plate. HPLC±MS analysis showed that
the reactions were uniformly successful across the 96-well reactor plate, without any
starting material being present. The diversity of this method was further extended
when mixtures of two different 1,3-dicarbonyl compounds were used in the same
Hantzsch synthesis, potentially leading to three distinct pyridine derivatives (not
shown).
Using a similar format, dihydropyrimidines were obtained in a microwave-expe-
dited version of the classical Biginelli three-component condensation (Scheme 12.24)
[73]. Neat mixtures of b-ketoesters, aryl aldehydes and (thio)ureas with polyphosphate
ester (PPE) as reaction mediator were irradiated in a domestic microwave oven for
1.5 min. The desired dihydropyrimidines were obtained in 61±95 % yield after aqu-
422 12 Microwave-assisted Combinatorial Chemistry
Ar O Ar
O
PPE H
O H R1O N
R1O NH2 MW, 1.5 min
+
R2 O HN X no solvent R2 N X
R3 R3
15 examples
Scheme 12.24 Microwave-assisted Biginelli dihydropyrimidine
synthesis.
R1
R1
O
H NH 2 Montmorillonite K-10 clay N
a) + MW, 4 min R2
R2 NC N X N X
no solvent
Y Y
X = Y = CH 14 examples
X = CH, Y = N
X = N, Y = CH
R2
R2
R1 O H2N NH4OAc, Al2O3 R1 N
b) MW, 20 min
+ O R3
R1 O no solvent R1 N
R3
H
4 examples
O Lawesson's reagent S
R1 MW, 8 min R1
Ar N Ar N
c) R2 no solvent R2
25 examples
Ar O Ar
O
R H O CN R CN
MW,15-20 min N
d) N
+ Me
Me no solvent X N N Ph
X N NH2 O Ph H
X = O, S; R = H, Me 9 examples
sualization procedures for the results of the reaction. In this particular example, the
synthesis of an arylpiperazine library (Scheme 12.26) was described, but the simpli-
city and general utility of the approach for the rapid screening of solvent-free micro-
wave reactions may make this a powerful screening and reaction optimization tool.
Other microwave-assisted parallel processes, for example, involving solid-phase or-
ganic synthesis (SPOS) have already been discussed in Sect. 12.2. (Schs. 12.6, 12.7,
and 12.10). In the majority of cases described so far, however, domestic multimode
microwave ovens have been used as heating devices, without utilizing specialized re-
actor equipment. Since reactions in household multimode ovens are notoriously dif-
ficult to reproduce due to the lack of temperature and pressure control, pulsed irra-
424 12 Microwave-assisted Combinatorial Chemistry
12.7
Equipment for High-throughput Microwave-assisted Synthesis
only one mode is present and the electromagnetic irradiation is focused directly
through an accurately designed wave guide on to the reaction vessel mounted in a
fixed distance from the radiation source. For combinatorial chemistry applications,
the key difference between the two types of reactor systems is that in multimode cav-
ities several reaction vessels can be irradiated simultaneously, whereas in monomode
systems, only one vessel can be irradiated at a time. One instrument designed for
high-throughput organic synthesis is the Ethos SYNTH microwave labstation (Mile-
stone Inc.) [80, 83]. This multimode instrument features a built-in magnetic stirrer,
direct temperature control of the reaction mixture with the aid of fiber-optic probes
and software that enables online temperature and/or pressure control by regulation
of microwave power output (1000 W maximum). The flexible modular platform al-
lows the use of standard glassware for reactions performed at atmospheric pressure,
sealed reaction vessels for performing synthetic transformations at elevated tempera-
ture and pressure, and the application of various parallel reactors [80, 83]. For all con-
figurations, the operator has full on-line access to all control parameters such as tem-
perature, pressure, microwave power, and irradiation time. For this reactor so-called
multiPREP rotors have been developed that house 36, 50, or 80 reaction vessels that
fit into the large multimode cavity. These continuously moving rotors operate at at-
mospheric pressure with ca. 20 mL glass, TFM, or PFA vessels. For applications re-
quiring sealed vessel conditions a rotor system with 36 glass vessels (Ethos multi-
PREP-36/P, Fig. 12.3) with a maximum operating temperature and pressure of
200 8C and 15 bar, respectively, has also been developed. In general, the temperature
in a parallel reaction is measured in one reference vessel either by fiber-optic sensor
or with the aid of a shielded thermocouple. During the reaction the temperature can
additionally be controlled by an external IR sensor in the wall of the instrument,
monitoring the surface temperature of the individual vessels as they move by the
sensor. Magnetic stirring of each vessel ensures homogeneous mixing of the sample
and even temperature distribution. Published applications of the multiPREP 50-ro-
tor system (open vessels) have included the generation of a 21-member library by
parallel solid-phase Knoevenagel condensations (Scheme 12.6) [34]. Here the tem-
perature was monitored with the aid of a shielded thermocouple inserted into one of
the reaction containers. It has been confirmed by standard temperature measure-
ments performed immediately after the irradiation period that the resulting end
temperature in each vessel was the same within ±2 8C [34]. In a different study invol-
ving the multiPREP 36/P rotor (Fig. 12.3) utilizing sealed glass vessels, the unifor-
mity of the reaction conditions in such a parallel set-up was investigated. For that
426 12 Microwave-assisted Combinatorial Chemistry
100 b
80 b Aldehydes a-f
c
Yield (%)
60 a c a f
e f d e
40 d
20
0
1 6 11 16 21 26 31 36
Reaction Vessel
Fig. 12.6 Monomode microwave reactor with and reaction vials (bottom right) are also dis-
integrated robotics interface for automated use played (Emrys Synthesizer, Personal Chemistry
(left). Details of the cavity/gripper (top right) AB) [86].
the outside of the reaction vessel. The instrument processes up to 120 reactions per
run with a typical throughput of 12±15 reactions h±1. Magnetic stirring of each indi-
vidual process vial (250 8C, 20 bar maximum temperature and pressure, respectively)
and unattended operation are some of the features of this microwave reactor. In con-
trast to the parallel synthesis application in multimode cavities this, approach allows
the user to perform a series of optimization reactions with each reaction separately
programmed.
The use of this setup for automated sequential microwave-assisted library synth-
esis was first reported in the context of preparing a series of dihydropyrimidines by
the Biginelli reaction (see Scheme 12.24). A diverse set of 17 CH-acidic carbonyl
compounds, 25 aldehydes, and 8 urea and/or thioureas was used in the preparation
of a dihydropyrimidine library [88]. Of the 3400 theoretically possible dihydropyrimi-
dine derivatives, a representative subset of 48 analogs was prepared using automated
addition of building blocks and subsequent sequential microwave irradiation of each
process vial. For most building block combinations, 10 min of microwave flash heat-
ing at 120 8C using AcOH±EtOH, 3 : 1, and 10 mol % Yb(OTf)3 as solvent±catalyst
system proved to be successful, leading to an average isolated yield of 52 % of
DHPM with > 90 % purity. For some building block combinations the general condi-
tions were modified, by changing the solvent, catalyst, reaction temperature, or irra-
diation time. The unattended automation capabilities of the microwave synthesizer
enables a library of this size to be prepared within 12 h.
In a related example involving the use of the same instrument (Fig. 12.6) in the
Hantzsch multicomponent condensation, the serial synthesis of 24 dihydropyridine
12.7 Equipment for High-throughput Microwave-assisted Synthesis 429
R O R O
O O
O H MW, 140-150 °C, 10-15 min R1 R1
R1 R1
+
R2 O O R2 aqueous NH3 (25%) R2 N R2
H
24 examples
Scheme 12.27 Generation of a dihydropyridine library using automated sequential
microwave processing.
derivatives was reported, involving the microwave assisted reaction of 6 different al-
dehydes with 4 different b-ketoesters or 1,3-dicarbonyl compounds and aqueous am-
monia (Scheme 12.27) [89]. Reactions were run in sealed vessels (see above) at
140±150 8C for 10±15 min, providing the desired products in 39±89 % yield and low
to excellent purities (53±99 %).
Additional applications of this technology for rapid lead discovery and lead optimi-
zation have been reported [87, 90±93]. It should also be noted that a variety of chemi-
cal transformations, in particular in the area of transition-metal catalyzed reactions,
have been performed with this or related equipment (Chapt. 11) [25]. Other mono-
mode microwave reactors using related concepts to introduce high-throughput were
recently introduced by CEM Corp. (Discover or Explorer line of products, Fig. 12.7.)
[81]. At the time of writing this review no published synthetic applications using this
microwave reactor were available.
The issue of parallel versus sequential synthesis using multimode or monomode
cavities, respectively deserves special comment. While the parallel setup allows for
considerable throughput that can be achieved in the relatively short timeframe of a
microwave-enhanced chemical reaction, the individual control over each reaction
vessel in terms of reaction temperature and/or pressure is limited. In the parallel
mode, all reaction vessels are exposed to the same irradiation conditions. In order to
ensure similar temperatures in each vessel, the same amount of the identical solvent
should be used in each reaction vessel because of the dielectric properties involved
[48]. An alternative to parallel processing the automated sequential synthesis of li-
braries can be a viable strategy. Irradiating each individual reaction vessel separately
gives better control over the reaction parameters, and allows for the rapid optimiza-
tion of reaction conditions. For the preparation of relatively small libraries, where de-
licate chemistries are to be performed, the sequential format may be preferable.
12.8
Conclusion
References
13
Microwave-Enhanced Radiochemistry
John R. Jones and Shui-Yu Lu
13.1
Introduction
13.1.1
Methods for Incorporating Tritium into Organic Compounds
The standard work of Evans [2] as well as a survey of the papers produced in the Jour-
nal of Labeled Compounds and Radiopharmaceuticals over the last 20 years shows that
the main tritiation routes are as given in Tab. 13.1. One can immediately see that un-
like most 14C-labeling routes they consist of one step and frequently involve a cata-
lyst, which can be either homogeneous or heterogeneous. One should therefore be
able to exploit the tremendous developments that have been made in catalysis in re-
cent years to benefit tritiation procedures. Chirally catalyzed hydrogenation reactions
(Knowles and Noyori were recently awarded the Nobel prize for chemistry for their
work in this area, sharing it with Sharpless for his work on the equivalent oxidation
reactions) immediately come to mind. Already optically active compounds such as
tritiated l-alanine, l-tyrosine, l-dopa, etc. have been prepared in this way.
The development of phase transfer catalysis, of supercritical fluids, of ionic liquids
and of course, new reagents, should also have considerable potential in the labeling
area. Furthermore there is the possibility of combining these approaches with en-
ergy-enhanced conditions ± in this way marked improvements can be expected.
13.1.2
Problems and Possible Solutions
The tritiation procedures given in Tab. 13.1 all have serious limitations/disadvan-
tages. Thus for all three hydrogen isotope exchange reactions HTO is used as the do-
13.1 Introduction 437
Reaction Example
catalyst
X T X = Br, Cl, I
4 Methylation N N
NaH
N CT3 I N
H CT3
5 Borohydride reduction O H HO H
NaBT4 T
the temperature, one effect more than compensating for the other ± however, the
time required is frequently very long, extending into days. Ion exchange resins, both
acid and base forms, can be used to overcome separation problems [12, 13].
Since the pioneering work of Garnett and Long [14, 15] much progress has been
made in increasing the selectivity of one-step metal-catalyzed hydrogen isotope ex-
change reactions. RhCl3, and iridium(I) catalysts of the type [Ir(COD)(L)2]PF6 (COD:
cis,cis-1,5-cyclooctadiene) have been successfully used by, amongst others, Heys,
Hesk, Lockley, Salter and their coworkers [16±19]. In some of these studies HTO has
been replaced by T2 as donor so that compounds of very high specific activity can be
obtained. Myasoedov and colleagues [20, 21] have also made extensive use of high
temperature solid state catalytic isotope exchange (HSCIE) for the tritiation of a wide
range of organic compounds at high specific activity.
Until recently hydrogenation reactions with T2 were performed on glass gas lines
but this is now frowned upon by the environmental and health and safety inspecto-
rate. Fortunately there are two commercial instruments available, one manufactured
in Switzerland and the other in the USA, which are entirely metallic and use an ura-
nium ªgetterº for storing T2 gas; gentle heating allows a predetermined volume of gas
to be transferred to the reaction vessel and on completion of the reaction any excess
can be returned to a secondary bed for storage and reuse. T2 gas is relatively inexpen-
sive and available at 100 % isotopic incorporation (specific activity of 56 Ci mmol±1).
The main disadvantage now is that it is sparingly soluble in many organic solvents
with the result that the catalyzed reactions, under both homogeneous and heteroge-
neous conditions, are frequently very slow.
Aromatic dehalogenation suffers from the disadvantage that only 50 % of the tri-
tium is incorporated, the rest appearing as waste. This situation is even more
marked for borotritide reductions but the problem can be overcome by using some
of the new tritide reagents that have recently become available as a result of the
synthesis of carrier-free lithium tritide (Scheme 13.1) [22]. Their reactivity can be
fine-tuned through the elements (e. g. B, Al, Sn) to which the tritium is attached and
by the electronic and steric nature of the substituents at the central atom.
TMEDA
n-BuLi + T2 LiT + n-BuT
hexane
Li(s-Bu)3BT LiPh3BT
s-Bu3B Ph3B
Et3B Bu3SnCl
LiEt3BT LiT Bu3SnT
B(OMe)3 Al(OBu-t)3
Scheme 13.1 Preparation of
LiB(OMe)3T Li(OBu-t)3AlT tritide reagents from LiT.
13.1 Introduction 439
Tritiated methyl iodide has the advantage that three tritiums can be incorporated
in one step so that compounds with a specific activity close to 80 Ci mmol±1 can be
prepared. CT3I is available from commercial sources and being a low boiling liquid
needs very careful handling. It is stable for short periods, consequently there is a
need for new methylating agents that offer greater flexibility.
13.1.3
The Use of Microwaves
. recoil labeling
. electrical discharge method
. low pressure method
. ion beam method
. microwave discharge activation (MDA) method
. adsorbed tritium method
tively simple to construct and the MDA method was seen to be superior to the other
radiation-induced methods of labeling, requiring a short reaction time and using a
small amount of tritium gas. However, the chemistry within this kind of microwave
plasma is complex with various tritium species being formed e. g. T +, T3+, T and T ±.
Consequently there are several reaction mechanisms taking place concurrently so
that the overall tritiation is characterized by low selectivity; extensive purification is
also necessary. Being a radiation-induced method the work can not be applied to
deuteration studies, in sharp contrast to the more recently developed microwave-en-
hanced methods. An excellent account of radiation-induced methods of labeling has
been given by Peng [33].
The exponential growth in microwave-enhanced synthetic organic chemistry (see
Fig. 1 in Refs. [8] and [34]) which Gedye and Giguere and their colleagues initiated
in 1986 may come to represent one of the most significant events in the develop-
ment of chemistry over the last half-century. Whilst microwave dielectric heating has
been widely used in the food processing area for many years chemists have been
slow in recognizing the potential, partly because the necessary theory is usually
taught as part of a physics degree course and partly because of the lack of suitable in-
strumentation for making quantitative, as distinct from qualitative, measurements.
This second problem will soon not exist as several powerful computer-controlled in-
struments have come to market within the last year.
A detailed account of the theory behind the interaction of microwaves with matter
is beyond the scope of this chapter but there are several excellent and recent reviews
available [35±37 and Chapt. 1], especially by Mingos and coworkers. Whilst micro-
waves are known to be electromagnetic waves in the frequency region 0.3 to
300 GHz current instrumentation tends to operate at a fixed frequency of
2450 MHz, corresponding to a wavelength of 12.2 cm, as the major part of the mi-
crowave region is assigned to radar and telecommunications. In the absence of free
electric charges the interaction between the microwave field and a molecule can be
expressed in terms of an interaction between an electric field and an electric dipole.
The polarization, P, represents the response:
P e0 we E
where E is the electric field, we the electrical susceptibility and e0 the electrical permit-
tivity in vacuum. The applied field creates a current which is the vector sum of the
polarization current and the conduction current. At high frequencies the molecular
inertia causes the polarization vector P to lag behind the applied field E and this is
the origin of the term ªdielectric lossº, expressed as tan d. Microwave energy is not
therefore transferred by convection or conduction but by dielectric loss ± a high value
for tan d indicates a high susceptibility to microwave energy. Polar solvents and so-
lids have high tan d values and are therefore favored for microwave-enhanced reac-
tions. Increasing the ionic strength of the reaction medium can also provide bene-
fits. In the case of nonpolar solvents, or under solvent-free conditions, it is claimed
[34] that when polar mechanisms are involved that nonthermal and specific micro-
wave effects may arise.
13.1 Introduction 441
13.1.4
Instrumentation
The instruments used for microwave-enhanced chemistry fall into two categories:
the multimode type that we commonly use in the kitchen and the mono-mode (or
single-mode) type that is specially designed for chemistry laboratories. The main dif-
ference between them comes down to how the microwave power is delivered and
controlled. In the multimode oven a time pulse at a fixed power level is used and the
waves within the cavity are not directed in any particular direction. The ovens do not
possess well defined electric fields and consequently the heating pattern within the
areas of the cavity is not homogeneous. In contrast, the mono-mode instrument has
a waveguide which focuses the microwave field on the sample. The power control
enables more precise energy to be delivered and this can be varied according to the
needs of the reaction. It also has a temperature sensor, and some are equipped with
a pressure sensor, which allow chemists to monitor or to control the heating process.
These devices give better reproducibility, shorter reaction time and sometimes better
chemical yields because the interaction between the microwave field and the sample
can be optimized.
Most of the early work was carried out using the multimode oven or commercial
domestic microwave oven. At Surrey we first used a Matsui M169BTunit (750 W), la-
ter on a Prolabo Synthewave S402 focused microwave reactor (300 W) was acquired.
In our earlier studies, the reagents were contained in a thick walled glass tube sealed
at the top but experience showed that a small conical flask (20 cm3) fitted with a sep-
tum stopper was well suited for most of the reactions. It was placed in a beaker con-
taining vermiculite and slowly rotated during the course of the irradiation. To avoid
undue pressure buildup the vessel can also be left unstoppered. For parallel reac-
tions, the Radley's RDT 24 place PTFE carousel reaction station together with Pyrex
glass tubes (25 cm3, o.d. = 15 mm) was placed on the turntable of the microwave
oven. The reaction vessels for the Synthelabo S402 are quartz tubes of different sizes.
For smaller scale reaction specially designed Pyrex tubes (o.d. = 9 mm) can be in-
serted inside the standard quartz tube (o.d. = 25 mm).
More and more researchers are using the mono-mode instruments and the manu-
facturers are bringing the second generation of microwave reactors (e. g. The Smith-
Synthesizer or the SmithCreator by Personal Chemistry, the Discover focused micro-
wave synthesis system by CEM, and the Ethos Synth microwave labstations by Mile-
stone) to the market; these are more effective and user friendly. New standard glass-
ware, capable of withstanding high temperature and pressure is also being devel-
oped and marketed.
442 13 Microwave-Enhanced Radiochemistry
13.2
Microwave-enhanced Tritiation Reactions
13.2.1
Hydrogen Isotope Exchange
This is one of the most versatile reactions known as it can be catalyzed by acids,
bases and metals as well as induced by photochemical and other means. It is also
amongst the simplest of organic reactions so that studies in this area have greatly
improved our understanding of reaction mechanisms. It is also one of the best meth-
ods for introducing deuterium/tritium into organic compounds. The knowledge
gained from basic research studies can therefore be put to good use within a fairly
short time interval.
Koves [38±40] was the first to show that acid-catalyzed aromatic hydrogen isotope
exchange could benefit from the use of microwaves and shortly afterwards we
showed that both heterogeneous and homogeneous metal-catalyzed reactions could
also be greatly accelerated [41, 42]. It was during these studies that we recognized
that further benefits could accrue by converting the organic compound (usually neu-
tral) to an ionic form e. g. by protonating the ±NH2 group. These preliminary studies
were then extended to take in a large number of nitrogen-containing heterocyclic
compounds ± here protonation on an NH2 group or the ring-based nitrogen was pos-
sible. This work was prompted by the earlier findings of Werstiuk employing a high
temperature, dilute acid, approach to the deuteration of many organic compounds.
Extensive labeling could be obtained provided the heating times were long (typically
12±50 h). Rarely can compounds of pharmaceutical interest withstand such demand-
ing conditions. Furthermore, the development of new, combinatorial chemistry, re-
quires that labeling reactions should be rapid so that high sample throughput can be
achieved.
The results for mono-, di- and fused ring substituted pyridines showed, as for the
earlier study on the microwave-enhanced deuteration of o-toluidine, that by first
forming the hydrochloride salt extensive labeling could be achieved within 20 min.
Furthermore the reaction products were easily isolated and of high purity, a fre-
quently noticed feature of these reactions. The reaction mechanism is represented in
Scheme 13.2, the deuterium being inserted in the ortho and para positions. The time
taken to reach a predetermined temperature depends greatly on the polarity of the
solvent used and as the results in Fig. 13.1 show this can be altered either by chan-
ging the solvent or adding a cosolvent. For labeling purposes a D2O (or HTO)±
NH3 Cl-
NH2 +
ND3 Cl-
+
ND3 Cl-
+
CH3
HCl CH3 D O CH3 D2O D CH3
2 microwaves
H2O
D
Scheme 13.2 Deuteration of o-toluidine hydrochloride under microwave conditions.
13.2 Microwave-enhanced Tritiation Reactions 443
250
1 Solvent b.p (oC) ε s δ
Tanδ
200
Temperature ('C)
2
1 DMSO 189 47 0.825
150 3
2 DMF 153 36.7 0.161
4, 5 3 BuOH 118 n/a n/a
100
6
4 H2 O 100 80.4 0.123
50 7
8 5 EtOH 78 24.6 0.941
0
6 CHCl3 61 4.8 0.091
0 50 100 150 200 250
7 CH2Cl2 40 9.1 0.042
Microwave irradiation time (sec) 8 1,4-dioxane 101 n/a n/a
DMSO mixture has great attraction as the heating time can now be reduced to a mat-
ter of 1±2 min.
The choice of solvent can also be beneficial in another respect. This possibility
was highlighted by the findings of Cioffi on the Raney Nickel catalyzed hydrogen±
deuterium exchange of a model carbohydrate [1-O-methyl-b-d-galactopyranoside]
but under ultrasonic irradiation (Tab. 13.2) [43]. Extensive deuteration at C-4 position
occurred for a series of ethereal solvents, the C-3 position was deuterated by seven
solvent systems and the C-2 position deuterated less extensively, also by seven sol-
vent systems. For 1,4-dioxane-D2O no labeling at the C-2 position occurred and for
1,2-dimethoxyethane-D2O no C-3 labeling was observed.
We chose the microwave-enhanced Raney Nickel catalyzed hydrogen isotope ex-
change of indole and N-methylindole as our substrates and D2O, CD3COCD3,
CD3OD and CDCl3 as the solvents. The thermal reaction had already been the sub-
ject of a recent study [44]. The microwave-enhanced method was some 500-fold fas-
ter than the corresponding thermal reaction (at 40 8C). Furthermore the pattern of la-
beling (Scheme 13.3) varied with the choice of solvent. Thus in the case of indole it-
Tab. 13.2 Pattern of labeling for a model carbohydrate under ultrasonic irradiation.
CH2OH CH2OH
OH O OMe Ni / ultrasound OH O OMe
OH H OH R
Solvent / R2O
H H R H
H OH R OH R = 2H, 3H
C2 27 28 20 25 <1 15
C3 73 84 59 60 62 <1
C4 89 93 59 76 75 29
D (36%)
D (97%)
(40%) D
D 2O
D (98%)
(44%) D N
D (94%)H
D (5%)
Raney Ni CD3COCD3
(47%) D
N microwaves D (93%)
H N
D (80%)H
D (26%)
CDCl3
N
H
Scheme 13.3 Pattern of deuteration at different sites for indole
using various solvents and microwave irradiation.
self the more polar solvents such as D2O and CD3OD give rise to general labeling
whilst CDCl3, for example, gave very regiospecific labeling. There is at this stage no
indication that the pattern of labeling in the microwave-enhanced reactions is differ-
ent in any way from that observed for the thermal reactions. However inherent in
both sets of results are mechanistic features worthy of more detailed investigation.
The benefits of using ionic compounds in microwave-enhanced reactions led us to
explore the possibility of using ionic solvents i. e. ionic liquids, as donors for both
deuterium and tritium. Whilst D2O is now relatively inexpensive and available at
high isotopic enrichment, tritiated water is usually employed, for safety reasons, at
low isotopic incorporation (we typically use HTO at 5 or 50 Ci mL±1 specific activity
corresponding to 0.2±2 % isotopic incorporation). This is a serious limitation when
there is a need to provide compounds at high specific activity.
Ionic liquids [45±47] are liquids containing only ions, and are fluid at or close to,
room temperature; those with higher melting points are called molten (or fused)
salts. They are nonvolatile and many are both air- and moisture-stable as well as
being good solvents for a wide range of both inorganic and organic molecules; they
frequently permit unusual combinations of reagents to be brought into the same
phase.
Many ionic liquids are based on N,N-dialkylimidazolium cations (BMI) which
form salts that exist as liquids at, or below, room temperature. Their properties are
also influenced by the nature of the anion e. g. BF4±, PF6±. The C-2(H) in imidazole is
fairly labile but the C-4(H) and the C-5(H) are less so. Under microwave-enhanced
conditions it is therefore possible to introduce three deuterium atoms (Scheme 13.4).
As hydrogen isotope exchange is a reversible reaction this means that the three deu-
terium atoms can be readily exchanged under microwave irradiation. For storage pur-
pose it might be best to back-exchange the C-2(D) so that the 4,5-[2H2] isotopomer
can be safely stored as the solid without any dangers of deuterium loss. The recently
13.2 Microwave-enhanced Tritiation Reactions 445
toluene
N N + n-BuCl N + N
-
Cl NaBF4
acetone
D D
CD3OD
N + N N + N
-
- Raney Ni BF4
D BF4 microwaves
13.2.2
Hydrogenation
Neither tritium or deuterium gas, with zero dipole moments, can be expected to in-
teract positively with microwave radiation. Their low solubilities are seen as a
further disadvantage. Our thoughts therefore turned towards an alternative proce-
dure, of using solid tritium donors and the one that has found most favor with us
is formate, usually as the potassium, sodium or ammonium salt. Catalytic hydro-
gen transfer of this kind is remarkably efficient as the results for a-methylcinnamic
acid show [50]. The thermal reaction, when performed at a temperature of 50 8C,
takes over 2 h to come to equilibrium whereas the microwave-enhanced reaction is
complete within 5 min. A further advantage is that more sterically hindered al-
kenes such as a-phenylcinnamic acid which are reduced with extreme difficulty
when using H2 gas and Wilkinson's catalyst are easily reduced under microwave-
enhanced conditions.
Formate can only donate one hydrogen, the other presumably coming from traces
of water present in the solvent or on the surface of the reaction vessel. With the need
to increase the specific activity of the product our thoughts turned to the preparation
446 13 Microwave-Enhanced Radiochemistry
of a diformate salt and as a first example we synthesized the diformic acid salt of
tetramethylethylenediamine (TMEDA).
The tritiated version could be prepared from tritiated formic acid which we had
prepared at high specific activity (2.5 Ci mmol±1) by a metal-catalyzed hydrogen±tri-
tium exchange procedure using T2 gas. The material can be stored either as a solid
or as a solution; if the latter any release of tritium by back exchange can be easily
monitored by 3H NMR spectroscopy. In our experience very little exchange occurs
over several weeks of storage [51].
The use of tritiated formates have further benefits. Firstly the exact amount neces-
sary for 100 % hydrogenation can be added so the problems associated with the use
of excess T2 gas are avoided and virtually no radioactive waste is produced. Secondly,
the pattern of labeling can be easily varied, as illustrated for the case of cinnamic
acid (using deuterium rather than tritium). When formate is used in D2O there are
three possible combinations: H2O + DCO±2, D2O + HCO±2 and D2O + DCO±2 (or the
DCO2D salt of TMEDA on its own). The 2H NMR spectra show that three iso-
topomers can be prepared ± C6H5CHDCH2COOH, C6H5CH2CHDCOOH, and
C6H5CHDCHDCOOH.
As of now no details of the synthesis of optically active tritiated compounds produced
under microwave-enhanced conditions have been published. Another area of consider-
able interest would be the study of solvent effects on the hydrogenation of aromatic
compounds using noble-metal catalysts as considerable data on the thermal reactions is
available [52]. Comparison between the microwave and thermal results could then pro-
vide useful information on the role of the solvent, not readily available by other means.
13.2.3
Aromatic Dehalogenation
As far as preparing tritiated (and deuterated) compounds are concerned aromatic de-
halogenation is second only to hydrogenation in importance. Furthermore it suffers
from the same disadvantages e. g. slow rates which are caused in part by the poor so-
lubility of the T2 gas in many organic solvents. The situation is however worse as
only 50 % of the tritium is introduced into the desired product. Once again we modi-
fied the classical dehalogenation by replacing D2/T2 gas with labeled formates [53].
There were some previous examples [54±56] of the use of formates in dehalogena-
tion reactions but none of these describe the utility of these agents for labeling or-
ganic compounds under the influence of microwave irradiation.
The N-4-picolyl-4-halogenobenzamide system (Scheme 13.5) was chosen as the ba-
sic substrate structure because:
H O H O
N R1COOK N
N N
microwaves
solvents 1
X R
X = Cl, Br, I
R1 = 2H, 3H
Scheme 13.5 Dehalogenation of N-4-picolyl-4-halobenzamide compounds.
. they yield strong pseudomolecular ions in both positive and negative ion HPLC/
MS; and
. they have simple NMR spectra.
13.2.4
Borohydride Reductions
One of the most attractive features of borohydride reductions is that under micro-
wave-enhanced conditions they can be performed in the solid state, and rapidly. We
were attracted by the work of Loupy [57], and in particular Varma [58, 59] who has
shown that irradiation of a number of aldehydes and ketones in a microwave oven in
the presence of alumina doped NaBH4 for short periods of time led to rapid reduc-
tion (0.5±2 min) in good yields (62±93 %). In our study [60] seven aldehydes and four
ketones were reduced (Tab. 13.3). Again reduction was complete within 1 min, the
products were of high purity (> 95 %), of high isotopic incorporation (95 %, same as
the NaBD4) and the reactions completely selective.
So far this remains the only microwave-enhanced borohydride deuteration study.
Corresponding tritiation studies can be anticipated in due course, especially with the
wider range of tritiated reducing agents, referred to previously, becoming available.
Significant reductions in radioactive waste can be anticipated.
Many borohydrides are highly unstable and have to be used as freshly prepared
ethereal solutions. However there are instances where the polymer-supported ver-
sions are more stable e. g. an Amberlyst anion exchange resin supported borohydride
and cyanoborohydride [61], polyvinylpyridine supported zinc borohydride [62] and
the corresponding zirconium borohydride [63]. Such compounds, in their labeled
forms, should turn out to be very useful.
448 13 Microwave-Enhanced Radiochemistry
Tab. 13.3 Borodeuteride reduction of carbonyl compounds under microwave conditions (750W,
1 min).
NaBD4/alumina
R
1
750 W, 1 min R1 OH
O
2 R
2 D
R
O NaBD4/alumina HO D
750W, 1 min
Aldehydes, R2 = H
13.2.5
Methylation Reactions
Methyl iodide is the most widely used methylating agent and is the favored route to
CT3- and CD3- containing compounds. As it can be rapidly synthesized (from CH4)
the [11C]-isotopomer also finds wide application (Sect. 13.4) in the rapidly expanding
positron emission tomography (PET) area [64]. Nevertheless it is generally agreed
that [11C]-methyl triflate is a better methylating agent ± it is more reactive and less
volatile so that one does not require cooling for trapping or heating for reaction to
take place.
We realized that using a low boiling liquid in a microwave environment, even on a
small scale, did not constitute ªbest practiceº and as for hydrogenations our thoughts
turned to using formates in a modified Eshweiler±Clarke reaction [65±67] and suc-
cessfully methylated a number of primary and secondary amines under microwave
conditions (Scheme 13.6) [68].
The methylation of secondary amines works better than for primary amines be-
cause there is no competition between the formation of mono- or dimethylated pro-
ducts. The best results for the microwave-enhanced conditions were obtained when
the molar ratios of substrate/formaldehyde/formic acid were 1 : 1 : 1, so that the
amount of radioactive waste produced is minimal. The reaction can be carried out in
neat form if the substrate is reasonably miscible with formic acid/aldehyde or in
DMSO solution if not. Again the reaction is rapid ± it is complete within 2 min at
120 W microwave irradiation compared to longer than 4 h under reflux. The reaction
mechanism and source of label is ascertained by alternatively labeling the formalde-
hyde and formic acid with deuterium. The results indicate that formaldehyde contri-
13.2 Microwave-enhanced Tritiation Reactions 449
CH2D
N
HCHO N
DCOOD
H CHD2
N N
DCDO N
N 120W, 1 min HCOOH
CD3
N
DCDO
DCOOD N
butes two deuterium atoms and the carbon whilst formic acid contributes one deu-
terium atom; there is no exchange between the formaldehyde and formic acid.
13.2.6
Aromatic Decarboxylation
O O
N-ethylmorpholine
D 2O
COOD D (100%)
750W(level II), 4min
MeO MeO
Scheme 13.7 Examples of microwave-enhanced decarboxylation.
acid as a D+/T+ donor several labeled imines have been prepared. Under microwave-
enhanced conditions the reactions would be expected to be much faster which again
would be very useful as tritiated imines are frequently used to label b-lactams and
other biologically interesting compounds such as a-aminophosphate [78]. It is also
worth mentioning that the corresponding phosphites (R2HP=O, R = OEt, OMe) are
cheap, nontoxic hydrogen-atom donors and attractive alternatives to organic tin hy-
drides [79] and have been identified as effective radical reducing agents for organic
halides, thioesters and isocyanides. The tritiated version of these reagents thus pro-
vide new opportunities, especially when coupled to microwave irradiation.
The potential of the microwave-enhanced decarboxylation route in the radioactive
waste area is immediately apparent ± washing the tritium waste with a protic solvent
leads to exchange of the labile tritium. The solvent can then be used with one of the
carboxylic acids mentioned above and after the microwave-enhanced decarboxylation
the waste is now in the form of a solid (greatly reduced volume) which may have
some further use.
13.2.7
The Development of Parallel Procedures
D
D D
O O
D2, catalysts
CH3COOEt D
O O
D D
N N D
15 catalysts screened
Scheme 13.8 Parallel screening and ranking of catalysts for the
reduction of isobutenyl groups.
A similar study [84, 85] was performed for ortho-directing hydrogen isotope ex-
change reactions of substituted aromatics. The initial screening showed catalytic ac-
tivity to reside exclusively with the Group VIII metals, especially the salts and com-
plexes of Ru, Rh and Ir. Iridium based catalysts are superior to those previously used
± they are more active, operate at lower temperatures and are applicable to a wider
range of substrates. Eventually CODIrAcac (acac = acetylacetone or 2,4-pentane-
dione) was identified as the catalyst displaying the best activity; further optimization
of activity was achieved by varying the ligand structure.
In our own preliminary studies [86] on parallel procedures under microwave-en-
hanced conditions, we have used the Radley's RDT 24 place PTFE carousel reaction
station on the turntable of the Matsui M 169BT microwave oven. In this way, we have
studied the catalytic activity of RhCl3 and Pd(OAc)2 towards the reduction or dehalo-
genation of 4-bromocinnamic acid and structurally similar compounds. A nine-reac-
tion matrix was used under microwave-enhanced conditions as illustrated in
Scheme 13.9 ± greatly reduced reaction times and easy optimization of reaction con-
ditions are immediate benefits. As robotics come to play an increasingly important
role in chemistry, one can immediately see more sophisticated labeling experiments
being undertaken.
13.2.8
Combined Methodology
D H
R2 R
1 2
Hydrogenation
R
Dehalogenation
H D
X X D
b c
metathesis
or no reaction
X = I, Br, Cl R2
R1 = H, COOH
R2 = H, COOK
X
a
COOH
Br
b (98%)
a (100%) c (95%)
RhCl3
No Cat Pd(OAc)2
RhCl3 RhCl3
b (98%) b (98%)
No Cat Pd(OAc)2
a (100%) c (95%)
COOH
CH2
Br
Br
Scheme 13.9 A 9-reaction parallel matrix under microwave-enhanced
conditions and schematic representation of carousel arrangement.
within 1 min when the reaction was performed under microwave-enhanced condi-
tions (Scheme 13.10) [87]. Therefore one-pot M+3 deuterium labeling in a multi-
functional molecule is achieved using a microwave-enhanced combined hydrogena-
tion/aromatic dehalogenation procedure. Isotopic purity in all three positions was
>95 %.
The screening of the catalytic activity in these reactions is made possible by the
readily available library of various heterogeneous and homogeneous transition metal
catalysts. The use of microwaves ensures that two reactions, each requiring different
times to reach equilibrium under thermal conditions, can now be completed within
a very short interval.
13.3 Microwave-enhanced Detritiation Reactions 453
H D
COOH COOH
RhCl3
D H
Br Br
RhCl3
Pd(OAc)2
Pd(OAc)2
H D
COOH COOH
D H
D D
Conditions for all reactions: catalysts, DCOOK/D2O, DMSO, level II of 750W, 60sec.
13.3
Microwave-enhanced Detritiation Reactions
The first of these can be expensive as also can the last, consequently there is in-
creasing interest in developing procedures which are more efficient than hitherto so
that less radioactive waste is produced in the first place. Some radioactive waste can
be disposed of via the drains but less so to the atmosphere, consequently there is
also increasing interest in the possibility of converting the waste to a suitable reagent
which can then be used in subsequent syntheses even though it is likely that the spe-
cific activity will be somewhat reduced. This philosophy can clearly be seen in the
operation of both the Trisorber and Tritech tritium gas units which have replaced the
old glass gas lines. In both instruments, the tritium gas is stored on an uranium
ªgetterº which on warming, releases a predetermined amount of T2 into the reaction
vessel. On completion of the reaction, any surplus gas is returned to a secondary bed
where it can be stored before use in a subsequent reaction.
Of all the methods used to tritiate organic compounds hydrogen isotope exchange
stands on its own by virtue of the fact that it is the only truly reversible reaction. Con-
sequently the benefits that have emerged from the study of microwave-enhanced hy-
drogen±tritium exchange should be immediately transferable to tritium±hydrogen
exchange. By performing the reaction in a good microwave solvent such as water, a
tritiated compound or mixture of compounds could be ªdecontaminatedº and the
HTO formed used in further tritiation studies, albeit at a lower level of specific activ-
454 13 Microwave-Enhanced Radiochemistry
HTO
1 140 62 b 48 60 562
2 1128 89 c 48 87 562
3 2.1 60 c 66 43 562
4 2.2 70 c 66 71 562
a
Microwave power set at level I (25 %) of 750 W
b
at 120 8C
c
at 180 8C
ity. If the specific activity requirements are not too demanding the whole tritiation/
detritiation cycle could be repeated several times, thereby making much better use of
the tritiated water.
The first example [88] that we encountered was of an oil that had been exposed to
a harsh tritium rich environment for a considerable time and had a level of radioac-
tivity in the 2.1±2.2 kBq mg±1 range. An inactive oil, as represented by its 1H NMR
spectrum, had an identical composition and was used as a model compound. Based
on previous work on shale oils [89] and engine basestocks [90] we knew that Raney
Nickel could catalyze the hydrogen±tritium exchange reactions although the thermal
reactions required high temperature (>180 8C) and long reaction times (48±66 h).
However under microwave-enhanced conditions 5 6 2 min pulses was sufficient to
exchange 60±90 % of the tritium in the oils; repeating the procedure led to further
detritiation (Tab. 13.4). Although these studies were performed on a small scale there
is no reason why, as for other microwave-enhanced reactions, the scale of the opera-
tion can not be increased, or, alternatively a flow system designed.
13.4
Microwave-enhanced PET Radiochemistry
rapid, simple to perform and the products easy to purify. Automation provides
further benefits whilst the need to use reactants on a micro scale requires a good ap-
preciation of both the reaction stoichiometry and the factors that influence the rates
of chemical reactions.
Because of these requirements, and in particular, the need to perform the reac-
tions rapidly, it is not surprising that microwaves were used in the PET synthesis
area at a very early stage [92]. However, the subsequent increase has not been as dra-
matic as one might have anticipated. Two recent reviews [8, 9] give an up-to-date pic-
ture whilst here we present some noteworthy examples.
13.4.1
11
C-Labeled Compounds
The most widely used synthetic routes follow on closely from those adopted for 14C
syntheses ± a small number of key precursors such as 11CO2, 11CN±, 11CH3I and
H11CHO are sufficient to label a wide range of compounds. O-, N- and S-alkylation
can be achieved using 11CH3X (X = I, OTf) [93, 94]. Through the use of microwaves
the time required to synthesize [N-methyl-11C]Flumazenil can be reduced from 5 to
0.5 min, accompanied by a 20 % improvement in yield. Further improvements can
be achieved by changing the solvent, in this case from acetone to dimethylforma-
mide (Scheme 13.11) [93].
[11C]-Cyanide is a secondary precursor, produced from 11CO2, but nevertheless as
a result of microwave-enhanced accelerations can be used to label a wide range of
amines, acids, esters and alcohols. Such an example is illustrated in Scheme 13.12
[95±98]. Reactions with less reactive substrates can be achieved by increasing the po-
larity of the reaction medium through the addition of various salts.
[11C]-formaldehyde has been widely used for reductive methylation reaction but
because of the marked fluctuation in the reported yields as well as impurities formed
in its preparation from 11CO2 the tendency has been to use 11CH3I for direct methy-
lation. However, the recent development [99, 100] of a low temperature no-carrier-
added method for preparing H11CHO, coupled to the microwave-enhanced Eschwei-
ler±Clarke reaction [65±68] has led to a resurgence of interest in the use of H11CHO.
N N N
CO2Et CO2Et CO2H
N 11
N N
CH3I NaOH
F N F N F N
H 11 11
O O CH3 O CH3
In acetone
5 min, 70 C, 45% 2 min, 70 C, 40%
0.5 min, 100W, 65% 0.5 min, 100W, 60%
in DMF
0.5 min, 50W, 80% 0.5 min, 50W, >95%
11
Scheme 13.11 Synthesis of [N-methyl- C]flumazenil acid.
456 13 Microwave-Enhanced Radiochemistry
O Cl
11
CN- LAH, THF
EtOH HO 11
CH2NH2 2 min, r.t. 90%
O
H2SO4
O 11 HO 11 0.25 min, 75W
CN COOH
or 5 min, 120 °C
0.25 min, 100W 90%
or 2 min, 80 °C HCl(g), EtOH
91% HO 11
COOEt 0.5 min, 75 W
or 10, 80 °C
LAH, THF 93%
In their study Roeda and Crouzel shower that the LAH reduction of 11CO2 produced
34 % H11CHO plus 59 % H11COOH. Our experience in using a mixture of DCHO
and DCOOH for the N-methylation of both primary and secondary amines suggests
that the corresponding [11C]-mixture would be ideal for the 11C-labeling of amines
under microwave-enhanced conditions.
13.4.2
18
F-labeled Compounds
O2N 18
(CH3(CH2)4N F
DMSO
H
O
H H
O
18F Oil bath 148 °C 45 min 75%
Microwave 250W 45s 70-83%
OCH3 N
N
N N NO2
O
+18 -
[K2.2.2] F
DMSO
OCH3 N
N
N N 18 F
O
Heating block 150 °C 20 min 16%
Microwave 500W 3 min 40%
H
N S
N NO2
N
O
1 8 F-/DMF O
K2.2.2/K 2 CO3
H
N S
N 1 8F
N
O
O
microwave 575 W 3min 23%
Oil bath 130 °C 30 min
458 13 Microwave-Enhanced Radiochemistry
+
X= Cl Br I NO2 N (CH3)3
50W, 4 min 26% 68% 8% 76% 96%
100W, 2 min 22% 71% 14% 88% 90%
O
O
R
[18F]KF-K2.2.2 R
DMSO
X 18
F
700W, 2.5 min
+ - + - + -
X= 4-NO2 2-NO2 4-N (CH3)3TfO 2-N (CH3)3TfO 4-I 4-N (CH3)3TfO
R= H H H H H CH3
63% 77% 3%
Scheme 13.15 Substituent effects in the microwave-enhanced
nucleophilic fluorination of bifunctional radiopharmaceutical
intermediates.
18F-
O O K2CO3/K2.2.2 O
X X DMF 18F
OEt N N
F F F
F F F
X = Br, F H+
O
18F
OEt
F
F
[109]. Formation of the oxazoline reduced undesirable 18F±19F exchange and the de-
sired product was rapidly (15s) isolated in good yield (45±55 %) with a specific activ-
ity (1.1 GBq mmol±1) that was 20 times higher than that obtained when using direct
debromination route. In another example, the synthesis of 2-deoxy-2-18F-fluoro-d-
glucose [110] the high purity of the product and the improved specific activity is at-
tributed to the short reaction times that are made possible via the use of micro-
waves.
References 459
13.5
Conclusion
From an early stage in the development of modern chemistry, it has been customary
practice to simplify complex problems by the selective use of isotopes. For example,
in the area of catalysis, relatively simple reactions such as hydrogen isotope ex-
change or hydrogenation reactions, have been investigated in order to delineate
some of the finer details of reaction mechanisms. This information has then been
applied so as to optimize the procedures that have been developed for labeling or-
ganic compounds. The last fifty years has seen the emergence of several, now large
companies, specializing in these areas, as well as many smaller versions.
Now, with the emergence of new, microwave-enhanced technology, the process is
destined to be repeated. In this case, faster, more selective and efficient procedures
will emerge, where levels of radioactive waste will be greatly reduced and a more fa-
vorable, environmentally friendlier image of the chemical industry achieved. The
products so produced can then go on to be used, in the words of Gordon Dean,
sometime Chairman of the US Atomic Energy Commission, ªto treat the sick, to
learn more about disease, to improve manufacturing processes, to increase the pro-
ductivity of crops and livestock, and to help man to understand the basic processes
of his body, the living things about him, and the physical world in which he exists.º
Acknowledgments
The tritium work at Surrey has been generously funded over many years by EPSRC
(previously SERC and at the beginning, SRC), the EU, NATO and the chemical in-
dustry. The task of writing the current chapter was undertaken as part of the EU
sponsored D10 COST Program (Innovative Methods and Techniques for Chemical
Transformations). We are also grateful to Dr Norman De'ath (Radleys Discovery
Technologies) for the loan of the RDT 24 place PTFE carousel reaction station.
References
1 D. Dalvie, Curr. Pharm. Design 2000, 6, 5 T. J. Simpson, Biosyn. Top. Curr. Chem.
1009. 1998, 195, 1.
2 E. A. Evans, Tritium and Its Compounds, 6 L. Y. Lian, D. A. Middleton, Prog.
2nd edn. Butterworths, London, 1974. Nucl. Mag. Res. Spectr. 2001, 39, 171.
3 a) J. A. Elvidge, J. R. Jones (eds.), Iso- 7 K. W. Turteltaub, J. S. Vogel, Curr.
topes: Essential Chemistry and Applications Pharm. Design 2000, 6, 991.
I,The Chemical Society, London, 1980; 8 N. Elander, J. R. Jones, S-Y Lu,
b) J. R. Jones (ed.), Isotopes: Essential S. Stone-Elander, Chem. Soc. Rev.
Chemistry and Applications II,The Royal 2000, 29, 239.
Society of Chemistry, London, 1988. 9 S. Stone-Elander, N. Elander,
4 F. J. Winkler, K. Kuhnl, R. Medina, J. Labelled Compd. Radiopharm. 2002,
R. Schwarzkaske, H. L. Schmidt, Iso. 45, 715
Environ. Health Studies, 1995, 31, 161.
460 13 Microwave-Enhanced Radiochemistry
14
Microwave Photochemistry
Petr Kln and VladimÌr CÌrkva
14.1
Introduction
UV radiation, certainly not sufficient to disrupt the bonds of common organic mole-
cules. We therefore assume that, essentially, photoinitiation is responsible for a che-
mical change and MW radiation subsequently affects the course of the reaction. The
objective of microwave photochemistry is frequently, but not necessarily, connected
to the electrodeless discharge lamp (EDL) which generates UV radiation when
placed in the MW field.
This chapter presents a complete picture of current knowledge about microwave
photochemistry. It provides the necessary theoretical background and some details
about synthetic and other applications, the technique itself, and safety precautions.
Although microwave photochemistry is a newly developing discipline of chemistry,
recent advances suggest it has a promising future.
14.2
Ultraviolet Discharge in Electrodeless Lamps
The electrodeless discharge lamp (EDL) [22] consists of a glass tube (ªenvelopeº)
filled with an inert gas and an excitable substance and sealed under a lower pressure
of a noble gas. A high-frequency electromagnetic field (radiofrequency or MW: 300±
3000 MHz) can trigger gas discharge causing the emission of electromagnetic radia-
tion. This phenomenon has been studied for many years [23] and was already well
understood in the nineteen-sixties [24]. The term ªelectrodelessº means that the
lamps lack the electrodes within the envelope. Meggers [24] developed the first EDL
using the mercury isotope 198Hg in 1942 (Fig. 14.1) and its earliest application was
in absorption spectroscopy [25]. EDL is usually characterized by a higher emission
intensity than that of hollow cathode lamps, lower contamination, because of the ab-
sence of the electrodes [26], and a longer lifetime [27].
electrostatic EDL
plasma waves
(b) (d)
14.2.1
Theoretical Aspects of the Discharge in EDL
The theory of EDL operation, as it is currently understood, is shown in the block dia-
gram in Fig. 14.2 [28]. Free electrons in the fill (i. e. electrons that have become sepa-
rated from the environment because of the ambient energy) accelerate as a result of
the electromagnetic (EM) field energy. They collide with the gas atoms and ionize
them to release more electrons. The repetitive effect causes the number of electrons to
increase significantly over a short period of time, an effect known as an ªavalancheº.
The electrons are generated by processes including collisional or collisionless transfor-
mation of EM waves, and normal or nonlinear wave absorption [26]. The energetic elec-
trons collide with the heavy-atom particles present in the plasma, exciting them from
a ground state to higher energy levels. The excitation energy is then released as EM
radiation with spectral characteristics which depend on the composition of the envel-
ope. The excited molecular or atomic species in the plasma can emit photons over
very broad region of the EM spectrum, ranging from X-rays to the IR [29].
14.2.2
The Fundamentals of EDL Construction and Performance
The EDL system is modular and consists of two basic parts, a gas filled bulb and a
power supply with waveguides or external electrodes. A typical EDL consists of a
scaled (usually quartz) tube envelope, which contains an inert gas (such as a noble
gas) and an excitable substance (e. g. Hg, Cd, Na, Ga, In, Tl, Sc, S, Se, or Te) [30]. The
envelope material must be impermeable to gases, an electrical insulator, and chemi-
cally resistant to the filling compounds at the temperature of operation.
Historically four basic methods have been used to excite discharges without elec-
trodes [31±35]. In the first method, referred as the capacitive coupling, the electric
field lines of the applied EM signal (usually 915 MHz) originate from one external
electrode, pass through the gas-filled bulb containing the discharge, and terminate
at a second external (coaxial) electrode. This discharge is similar to arc discharge in
an electrode lamp, but needs a higher current. The second method of exciting EDL
466 14 Microwave Photochemistry
with MW power (typically 2450 MHz) is to place the bulb in the path of radiation
from a directional antenna. The microwave discharge is excited by both electric and
magnetic components of the EM field. Because free propagation of the MW power
occurs, however, emission is often inherently inefficient. This method is used for ex-
citation of EDL inside a MW oven. The third method is called the traveling wave dis-
charge ± a gap between the external electrodes provides the electric field that
launches a surface wave discharge. The fourth method uses the inductive coupling of
EDL and the system can be compared with an electrical transformer. An alternating
current in the coil generates a changing magnetic field inducing the electric field
that drives a current into the plasma. The operating frequency is limited to approxi-
mately 50 kHz [36].
The operating conditions affecting electrodeless lamp performance are a combina-
tion of many factors [30]:
. The inert gas. The arc chamber contains a buffer noble gas (usually Kr, Xe, or Ar)
which is inert to the extent that it does not adversely affect the lamp operation. He-
lium has a higher thermal conductivity than other noble gases and, therefore,
higher thermal conduction loss is observed [37]. The inert gas easily ionizes at low
pressure but its transition to the thermal arc is slower and the lamp requires a
longer warm-up time. Ionization is more difficult at higher pressures and it re-
quires a higher input power to establish the discharge. The pressure in the EDL at
the operating temperature is usually much higher (5±20 atm) than that of a con-
ventional electrode lamp.
. The choice of the fill material initiating the discharge is very important. Together
with a standard mercury fill it is often desirable to incorporate an additive in the
fill material that has a low ionization potential [38, 39]. One category of low-ioniza-
tion-potential materials is the group of alkali metals or their halides (LiI, NaI) but
some other elements, such as Al, Ga, In, Tl [40, 41], Be, Mg, Ca, Sr, La, Pr, or Nd
[23, 37, 42], can be used.
. The dimensions and properties of the lamp envelope are based on the discovery that
the volume of Hg is critical for the effective UV operation [43]. Higher Hg pres-
sures result in the need to use higher MW power levels. To focus the MW field effi-
ciently into the EDL a special Cd low-pressure lamp with a metal antenna (a mo-
lybdenum foil) was developed by Florian and Knapp [44].
. The nature and characteristics of the EM energy-coupling device are discussed in
Sect. 14.2.3.
. The frequency and intensity of EM energy is determined by the type of a device.
A standard MW power source operates around 915 or 2450 MHz.
14.2.3
Spectral Characteristics of EDL
shown in Fig. 14.3) and molecular fills give continuous spectra [45]. The total emis-
sion output of Hg-EDL in the region of 200±600 nm is approximately the same as
that of the electrode lamp with the same power input [46]. The distribution of radia-
tion is, however, markedly different, as a result of much higher Hg pressure and the
greater number of atoms that can participate in the plasma. EDL emit over three
times as much UV and less than a half as much IR as a conventional lamp [47]. It
has been noted that EDL and electrode lamps furnish different spectra when the fill
includes a rare-earth material but similar spectra when a nonrare-earth fills are used
[48]. Addition of some elements (Sn, Pb, Ga) has a very significant effect on the spec-
tral distributions of the EDL [46]. Most lamps emit less efficiently below 280 nm
than a standard Hg lamp (Tab. 14.1). The advantages of EDL for microwave photo-
chemical applications are discussed in the Sect. 14.3.
14.3
Microwave Photochemical Reactor
Hg (Ar) Hg 185, 254, 365, 405, 436, 546 30, 31, 43, 44, 48±50
Cd (Ar) Cd 229, 326 30, 44, 51
SnI2 (Ar) SnI2 400±850, 610 52
FeCl2 (Ar) Fe 248 30
Zn (Ar) Zn 214 30, 44, 53
CuCl (Ar) Cu 325 30
NaI (Xe, Kr) Na 589 54, 55
Mg, H2 (Ar) MgH 518, 521, 480±560 56
AlBr3 (Ne) AlBr 278 57
AlCl3 (Ne) AlCl 261 58, 59
Ga, GaI3 (Ar) Ga 403, 417, 380±450 49, 53
InI3 (Ar) In 410, 451 53
TlI (Ar) Tl 277, 535 30, 53
PCl4 (Kr) P2 380 60
S (Ar) S 350±850, 525 40, 61, 62
Se (Ar, Xe) Se 370±850, 545 62±64
Te (Xe) Te 390±850, 565 62, 64
Ar (Ar) Ar2 126, 107±165 29, 65
Ar, Cl2 (Ar) ArCl 175 29, 65
Xe, Cl2 (Xe) XeCl 308 29, 65
B2O3, S (Kr) B2S3 812 66
I2, HgI2 (Ar) I 206 67
produce desired photoproducts (the EDL was positioned inside the MW cavity,
although outside the reaction vessel). Several similar reactors have been proposed
for UV sterilization [80±82] or for treatment of waste water containing organic pollu-
tants [83±85].
CÌrkva and Hjek have proposed a simple application of a domestic microwave
oven for microwave photochemistry experiments [86]. In this arrangement, the EDL
(the MW-powered lamp for this application was specified as a microwave lamp or
MWL) was placed in a reaction vessel located in the cavity of an oven. The MW field
generated a UV discharge inside the lamp that resulted in simultaneous UV and
MW irradiation of the sample. This arrangement provided the unique possibility of
studying photochemical reactions under extreme thermal conditions (e. g. Ref. [87]).
Kln et al. published a series of papers that described the scope and limitations of
this reactor [88±91]. In a typical design (Figs. 14.5 and 14.6) four holes were drilled
into the walls of a domestic oven ± one for a condenser tube in the oven top, another
in the side for an IR pyrometer, and two ports for a glass tube with circulating water.
Part of the oven bottom was replaced with an aluminum plate to enable magnetic
stirring. The opening for the IR pyrometer could also serve for an external (addi-
tional) source of UV radiation. The vessel was connected to a very efficient water-
cooled condenser by means of a long glass tube. The circulating cool water or a MW-
absorbing solid material (dummy load ± basic Al2O3, molecular sieve, etc.) were
used when a small quantity of a non- or poorly absorbing sample was used for mi-
crowave photochemical experiments. The material removed excess microwave power
and prevented the magnetron from being destroyed by overheating.
The EDL had always to be placed in a position, in which the solvent cooled it effi-
ciently, because lamp overheating might cause failure of lamp emission. Intense IR
output from the lamp triggered immediate boiling of all solvents including nonpolar
(MW-transparent) liquids [89, 90]. Polar solvents, on the other hand, absorbed most
of MW radiation, resulting in reduction of UV output efficiency. Tab. 14.2 depicts the
most important advantages and disadvantages of EDL applications.
Chemat and his coworkers [92] have proposed an innovative MW±UV combined
reactor (Fig. 14.7) based on the construction of a commercially available MW reactor,
the Synthewave 402 (Prolabo) [9]. It is a monomode microwave oven cavity operating
at 2.45 GHz designed for both solvent and dry media reactions. A sample in the
quartz reaction vessel could be magnetically stirred and its temperature was moni-
tored by means of an IR pyrometer. The reaction systems were irradiated from an ex-
ternal source of UV radiation (a 240-W medium-pressure mercury lamp). Similar
photochemical applications in a Synthewave reactor using either an external or inter-
nal UV source have been reported by Louerat and Loupy [93].
A microwave-assisted, high-temperature, and high-pressure UV digestion reactor
has been developed by Florian and Knapp [44] for analytical purposes. The appara-
tus consists of the immersed electrodeless discharge lamp operating as a result of
the MW field in the oven cavity (Fig. 14.8). An antenna fixed to the top of EDL en-
hanced the EDL excitation efficiency. Another interesting MW±UV reactor has
14.3 Microwave Photochemical Reactor 471
Tab. 14.2 Advantages and disadvantages of EDL applications in photochemistry. Adapted from
Ref. [90].
Advantages
Disadvantages
charge through an oxygen stream [96]. The combined effect of MW and UV irradia-
tion could increase the singlet oxygen concentration in the MW cavity, particularly in
the presence of a photosensitizer.
14.4
Microwave Photochemistry
14.4.1
Interactions of Ultraviolet and Microwave Radiation with Matter
Although microwave chemistry has already received widespread attention from the
chemical community, considerably less information is available about the effect of mi-
crowave radiation on photochemical reactions. Photochemistry is the study of the in-
teraction of ultraviolet or visible radiation (E = 600±170 kJ mol±1 at l = 200±700 nm)
with matter. The excess energy of electronically excited states significantly alters spe-
cies reactivity ± it corresponds, approximately, to typical reaction activation energies
helping the molecules overcome activation barriers. The microwave region of the elec-
tromagnetic spectrum, on the other hand, lies between infrared radiation and radio
frequencies. Its energy (E = 1±100 J mol±1 at n = 1±100 GHz) is approximately 3±6 or-
ders of magnitude lower than that of UV radiation (a typical MW kitchen oven oper-
ates at 2.45 GHz). Microwave heating is not identical with classical external heating,
at least at the molecular level. Molecules with a permanent (or induced) dipole re-
spond to an electromagnetic field by rotating, which results in friction with neighbor-
ing molecules (thus generating heat). Additional (secondary) effects of microwaves in-
clude ionic conduction (ionic migration in the presence of an electric field) and spin
alignment.
14.4 Microwave Photochemistry 473
M* I P
M hν
Fig. 14.10 A simplified model of the ∆T (MW)
synergic effect of UV and MW radia- ∆T (MW) M* I I∆
tion on a chemical reaction, where
depicts ªhotº molecules, and kr and kr M∆ hν' kr kr∆
are the rate constants of the pro- P
474 14 Microwave Photochemistry
14.4.2
Photochemical Reactions in the Microwave Field ± Thermal Effects
Baghurst and Mingos have hypothesized that superheating of polar solvents at atmo-
spheric pressure, so that their average temperatures are higher than the correspond-
ing boiling points, occurs as a consequence of dissipation of the microwaves
throughout the whole volume of a liquid [97]. With the absence of nucleation points
necessary for boiling heat loss occurs at the liquid/reactor wall or at liquid/air inter-
faces. Many reaction efficiency enhancements reported in the literature have been
explained as the effect of superheating when the reactions were essentially per-
formed in sealed vessels without stirring [98±102]; this effect is also expected in mi-
crowave photochemistry experiments in condensed media. Gedye and Wei [15], for
example, have observed enhancement of the rate of several different thermal reac-
tions by factors of 1.05±1.44 in experiments accomplished in a domestic-type MW
oven but not in a variable-frequency microwave reactor. The enhancement was inter-
preted as a consequence of solvent superheating or hot-spot formation rather than
nonthermal effects. Stadler and Kappe reported similar results in an interesting
study of the MW-mediated Biginelli reaction [14].
Kln et al. [91] successfully evaluated MW superheating effects in polar solvents
by use of a temperature dependent photochemical reaction. It is known that quan-
tum efficiencies of the Norrish type II reaction [103] for a mixture of some p-substi-
tuted valerophenones depend on the presence of a weak base, because of specific hy-
drogen-bonding to the biradical OH group (BR; Sch. 14.1) [88, 91]. This reaction pro-
vided linear temperature dependence of the efficiency over a broad temperature re-
gion and the system served as a photochemical thermometer at the molecular level,
even for the MW-heated mixtures. The magnitude of the photochemical change in
the MW field suggested the existence of superheating effect (4±11 8C) for three ali-
1. h ν
O H H 2. ISC
* O H H
3
R CH3 R CH3
R = H, CH3 solvent
H
O H OH H
∆T
R CH3 R CH3
solvated BR BR
HO
O
R +
R
Scheme 14.1
14.4 Microwave Photochemistry 475
Scheme 14.2
O
UV or MW OH
O
bentonite
O O
O
phatic alcohols and acetonitrile. The results were in a perfect agreement with mea-
surements by use of a fiber-optic thermometer.
Chemat and his collaborators [92] reported the UV- and MW-induced rearrange-
ment of 2-benzoyloxyacetophenone, in the presence of bentonite, into 1-(o-hydroxy-
phenyl)-3-phenylpropane-1,3-dione in methanol at atmospheric pressure (Sch. 14.2).
The reaction, performed in the reactor shown in Fig. 14.7, was subject to a signifi-
cant activation effect under simultaneous UV and MW irradiation; this corresponded
at least to the sum of the individual effects (Fig. 14.11). The rearrangement, however,
was not studied in further detail. Such competitive processes can be described by the
diagram in Fig. 14.9, because the product obtained from both types of activation was
the same.
CÌrkva and Hjek [86] studied a photochemically or microwave-induced addition
of tetrahydrofuran to perfluorohexylethene (Sch. 14.3). Whereas the thermal reaction
was too slow, photochemical activation was very efficient, with no apparent thermal
effects of MW radiation. Combined UV and MW radiation has principally been used
to initiate EDL operation in the reaction mixture (Fig. 14.5). Another illustration of
the MW±UV-assisted reaction has been demonstrated by Nçchter et al. [1] on dehy-
drodimerization reactions of some hydrocarbons. Zheng et al. [104, 105] reported
microwave-assisted heterogeneous photocatalytic oxidation of ethylene using porous
TiO2 and SO2± 4 /TiO2 catalysts. Significant enhancement of the photocatalytic activity
was attributed to the polarization effect of the highly defected catalysts in the MW
field. Louerat and Loupy studied some photochemical reactions (e. g. stilbene iso-
merization) in homogenous solutions and on solid supports such as alumina [93].
476 14 Microwave Photochemistry
UV (MW )
H2C CH (CF2) 5-CF3 + CH2 CH2 (CF2) 5-CF3
O O
Scheme 14.3
14.4.3
Nonthermal Microwave Effects ± Intersystem Crossing in Radical-recombination
Reactions
Radical pairs and biradicals are extremely common intermediates in many organic
photochemical (and some thermal) reactions. A singlet state intermediate is
formed from the singlet excited state in reactions that conserve spin angular mo-
mentum whereas the triplet intermediate is obtained via the triplet excited state.
Radical pairs in solution coherently fluctuate between singlet and triplet electronic
states [106, 107] and the recombination reactions are often controlled by electron±
nuclear hyperfine interactions (HFI) on a nanosecond time scale [108, 109]. Only
a pair of the neutral radicals with singlet multiplicity recombines. A triplet pair
intersystem crosses into the singlet pair or the radicals escape the solvent cage
and react independently at a later stage (Fig. 14.12) [110]. The increasing effi-
ciency of triplet-to-singlet interconversion (`mixing' of states) leads to a more rapid
recombination reaction and vice versa. It is now well established that a static mag-
netic field can effect intersystem crossing in biradicals (magnetic field effect,
MFE) and the effect has been successfully interpreted in terms of the radical pair
mechanism [111, 112]. This concept enabled explanation of nuclear and electronic
spin polarization phenomena during chemical reactions, e. g. chemically induced
dynamic nuclear polarization (CIDNP) or reaction yield-detected magnetic reso-
nance (RYDMAR).
An external magnetic field that is stronger than the hyperfine couplings inhibits
(because of Zeeman splitting) singlet±triplet interconversions by isolating the tri-
plets T+1 and T±1 from the singlet (S), which can, therefore, mix only with T0
(Fig. 14.12 a, b). For the triplet-born pair, the magnetic field reduces the probability of
radical recombination. The microwave field, which is in resonance with the energy
gaps between the triplet levels (T+1 or T±1) and T0, transfers the excess population
from the T+1 or T±1 states back to a mixed state. Application of a strong magnetic
field to the singlet-born radical pair leads to an increase in the probability of recombi-
nation that can, however, be also controlled by microwave irradiation [112].
14.4 Microwave Photochemistry 477
'in cage'
R-R recombination
UV
* [R-R]1 [R R]
ISC ISC
radical
* [R-R] 3 [R R] escape
MW
T+1
X
ISC T+1
S T0 S T0
T-1 X
T-1
'in cage' radical
recombination escape
(a) zero external magnetic field (b) applied external magnetic field
Another study of the electron-hole recombination of radical ion pairs (pyrene anion
and dimethylaniline cation) in solution has appeared recently [126]. Triplet±singlet in-
terconversions as a result of HFI are relatively efficient in a zero magnetic field (to be
more precise, in the Earth's field of ~50 mT). All the states are almost degenerate, as-
suming separation of the radicals is such that their electronic exchange interaction is
negligible [111]. Jackson and his coworkers [127] suggested that the resonance energy
of the oscillating field should be tuned to the HFI in one of the radicals. With a typical
value of HFI in the radicals of 0.1±3.0 mT, the oscillating magnetic field effect, enhan-
cing the conversion of the singlet state to the triplet (as was observed for weak static
fields) is expected in the radiofrequency region (3±80 MHz) [124]. Canfield et al. [128±
130] calculated the effects and proved them experimentally for the radical pairs in-
volved in coenzyme B12-dependent enzyme systems. Other theoretical studies have
appeared in the past 5 years [126, 131, 132]. Whether electromagnetic fields influence
animal and human physiology is still an open question. It has, for example, been sug-
gested that radiofrequency fields might disorient birds [133]. Detailed experimental
studies of OMFE in the microwave region have not yet been performed. Hoff and Cor-
nelissen [134] reasoned in their paper that triplet state kinetics could be affected rather
by a pulse of resonant microwaves than by equilibrium methods in the zero static
field.
According to the OMFE model a weak oscillating magnetic field (the magnetic in-
teractions are much smaller than the thermal energy of the molecule [131]) has no
impact on equilibrium constants or activation energies; it can, however, have im-
mense kinetic control over the reaction of the radicals [131]. The simplified kinetic
scheme in Fig. 14.13 shows the excitation of a starting material R±R' into the singlet
state, which intersystem crosses to the triplet (kisc) and is followed by cleavage (kcl) to
the triplet radical pair. The oscillating magnetic field influences the state of mixing
of the radical pair (kTS and kST). The probability that the triplet radical pair will form
the products is given by the efficiency of radical escape from the solvent cage (kesc)
and of triplet-to-singlet intersystem crossing (kTS). The recombination reaction is
very fast when the tight radical pair reaches the singlet state.
hν
* [R-R'] 1 R-R'
kisc
3
* [R-R'] krec
kcl
kesc kTS
products [R R'] [R R']
kST Fig. 14.13 The oscillating magnetic field
OMFE effect (OMFE) in the triplet state radical pair
reaction.
14.4 Microwave Photochemistry 479
14.4.4
Analytical and Environmental Applications of Microwave Photochemistry
14.5
Industrial Applications
14.6
Concluding Remarks
Acknowledgments
We would like to thank Milan Hjek and JaromÌr Literk for their participation in
our research projects, and Andr Loupy for fruitful discussions. We also acknowl-
edge the Grant Agency of the Czech Republic (203/02/0879) for financial support.
References
148 V. A. Danilychev, US Pat. Appl. (1999) 150 J. D. Michael, US Pat. Appl. (2000) US
US 5931557. 6162406.
149 J. D. Michael, US Pat. Appl. (2001) US
6171452.
487
Index
wavelength 3 x
Weflon 427 X-ray diffraction 235
Wilkinson's catalyst 400, 416, 445
Willgerodt 50 z
Willgerodt reactions 46 zeolite 197, 219, 445
Williamson synthesis 44, 153 zeolite±HY 355
Wittig olefination reactions 191 zinc chloride 194
Wolff±Kishner reduction 78, 193 zwitterionic intermediates 335