Utilization of Electromagnetic and Acoustic Irradiation in Enhancing Heterogeneous Catalytic Reactions

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Applied Catalysis A: General 279 (2005) 122

www.elsevier.com/locate/apcata

Review

Utilization of electromagnetic and acoustic irradiation in enhancing


heterogeneous catalytic reactions
B. Toukoniitty, J.-P. Mikkola, D.Yu. Murzin*, T. Salmi
Laboratory of Industrial Chemistry, Process Chemistry Centre, Abo Akademi,
Biskopsgatan 8, 20500 Turku, Finland
Received 8 June 2004; received in revised form 25 October 2004; accepted 27 October 2004
Available online 15 December 2004

Abstract
Process intensification has become a very interesting approach, transforming current practices in chemical engineering and bringing forth
new developments in equipment, processing techniques and operational methods. This development aims at more compact, safe, energy
efficient and environmentally friendly process.
Several unconventional processing techniques rely on alternative forms of energy. Chemistry under extreme and non-conventional
conditions is an actively studied topic in applied research and industry. Alternatives to conventional synthetic procedures promise
enhancement of reaction rates, yields, selectivity and also bear promise of milder reaction conditions in chemical synthesis. During the
last few decades, chemical application of ultrasound (US) and microwave (MW) irradiation has received a lot of attention and widespread
research is going on in these areas. Significant enhancement of selectivities, rates and yields in chemical reactions has been achieved by means
of US and MW irradiation. The popularity of US and MW irradiation as chemical laboratory techniques is rapidly growing, based on the
number of publications, presentations and meetings, demonstrating their vast potential. Other less exploited methods are solar and plasma
reactors. These, however, will not be touched in the present review, which focuses on the use of ultrasound and microwaves as sources of
energy, mainly in catalytic applications.
# 2004 Elsevier B.V. All rights reserved.
Keywords: Ultrasound irradiation; Microwave irradiation; Hetrogeneous catalytic reaction

Contents
1.

Sonochemistry . . . . . . . . . . . . . . . . . . . . . .
1.1. Introduction . . . . . . . . . . . . . . . . . . .
1.2. History . . . . . . . . . . . . . . . . . . . . . .
1.3. Applications of ultrasound . . . . . . . . .
1.4. Acoustic cavitation . . . . . . . . . . . . . .
1.4.1. Homogeneous sonochemistry .
1.4.2. Heterogeneous sonochemistry
1.5. Sonoluminescence . . . . . . . . . . . . . .
1.6. Factors effecting cavitation . . . . . . . .
1.6.1. Ultrasonic frequency. . . . . . .
1.6.2. Acoustic power . . . . . . . . . .
1.6.3. Solvent . . . . . . . . . . . . . . . .
1.6.4. Gas type and content . . . . . .
1.6.5. External temperature. . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

* Corresponding author. Tel.: +358 2 215 4985; fax: +358 2 215 4479.
E-mail address: dmurzin@abo.fi (D.Yu. Murzin).
0926-860X/$ see front matter # 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2004.10.044

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

2
2
2
2
2
3
3
3
4
4
4
4
4
5

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.. 5
.. 6
. 10
. 10
. 10
. 10
. 10

2.

Microwave chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Industrial application . . . . . . . . . . . . . . . . . . . . . . . .
2.4. Microwaves in chemistry. . . . . . . . . . . . . . . . . . . . . .
2.4.1. Microwave spectroscopy . . . . . . . . . . . . . . . .
2.4.2. Dielectric polarization. . . . . . . . . . . . . . . . . .
2.5. How materials interact with microwaves . . . . . . . . . . .
2.6. Microwave heating of liquids and solids . . . . . . . . . . .
2.7. Microwaves specific and athermal effects in chemistry .
2.8. Microwaves application in heterogeneous catalysis . . . .
2.9. Microwave equipment . . . . . . . . . . . . . . . . . . . . . . . .
2.9.1. Multi-mode cavities . . . . . . . . . . . . . . . . . . .
2.9.2. Single-mode cavities. . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

11
11
11
11
11
11
12
12
13
13
14
19
19
19

3.

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

21

1.7.
1.8.
1.9.

1.6.6. External pressure . . . . . .


Heterogeneous catalyzed systems .
Other applications . . . . . . . . . . .
1.8.1. Polymerization reactions .
Ultrasonic systems . . . . . . . . . . .
1.9.1. Ultrasonic bath. . . . . . . .
1.9.2. Horn (probe) system . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

1. Sonochemistry
1.1. Introduction
Ultrasound is simply sound pitched above the frequency
bond of human hearing. It is a part of sonic spectrum that
ranges from 20 kHz to 10 MHz and corresponds to the
wavelengths from 10 to 103 cm.
The application of ultrasound, in connection to chemical
reactions, is called sonochemistry. The range from 20 kHz to
around 1 MHz is used in sonochemistry, since acoustic
cavitation in liquids can be efficiently generated within this
frequency range [1]. However, common laboratory equipment utilizes the range between 20 and 40 kHz.
1.2. History
Chemical application of ultrasound has become an
exciting new field of research during the last decade,
although the interest in ultrasound and cavitational effect
dates back over 100 years. The first report on cavitation was
published in 1895 by Thornycroft and Barnaby when they
noticed that the propeller of their submarine, the H.M.S.
Daring, was pitted and eroded [2]. The first commercial
application appeared in 1917 when Paul Langevin invented
and developed echo-sounder. The original echosounder, later on, became underwater SONAR for
submarine detection during the World War II. The same
year Lord Rayleigh published the first mathematical model

.
.
.
.
.
.
.

.
.
.
.
.
.
.

for cavitational collapse, predicting enormous local temperatures and pressures. In 1927, Richards and Loomis
published the first paper on chemical effects of ultrasound
[3]. In 1980, Neppiras used for the first time the term
sonochemistry in a review of acoustic cavitation [4].
Since that time, sonochemistry has become so popular that
in 1986 the first international meeting on sonochemistry
took place at Warwick University, UK. In the 1980s, the
renaissance of sonochemistry research has accelerated.
1.3. Applications of ultrasound
Household use of ultrasound is widespread: ultrasound is
applied in dog whistles, burglar alarms and jewellery cleaners.
Also, high frequency ultrasound, at around 5 MHz, is
applied for medical purposes: to remove kidney stones
without surgery, to image fetal development during
pregnancy, to treat cartilage injuries (such as tennis
elbow), for dental descaling and in physiotherapy [5].
In industry, ultrasound is important in extraction, impregnation, crystallization, filtration, emulsifying cosmetics and
foods, welding plastics, powder production, cutting alloys,
cleaning and disinfections of medical instruments as well as
in food processing equipment [6].
1.4. Acoustic cavitation
The origin of sonochemical effects in liquids is the
phenomenon of acoustic cavitation. Ultrasound is trans-

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Fig. 1. Cavitation process [5]: formation, growth, and violent collapse of


bubbles, which generates very high pressures and temperatures. This
process is called cavitation.

mitted through a medium via pressure waves by inducing


vibrational motions of the molecules, which alternately
compress and stretch the molecular structure of the medium
due to a time-varying pressure [5]. Molecules start to
oscillate around their mean position and if the strength of the
acoustic field is sufficiently intense, cavities are created in
liquids. This will happen if the negative pressure exceeds the
local tensile strength of the liquid.
The cavities are also called cavitation bubbles and the
process itself is referred to as cavitation. Two types of
cavitation are known: stable and transient. Stable cavities are
bubbles, which form and oscillate around their equilibrium
position over several rarefaction/compression cycles, before
collapsing or never collapsing at all [2].
A transient cavity exists for only a few acoustic cycles. It
means that the bubbles grow during the expansion phase of
the sound wave, followed by a shrinking phase and,
consequently, a violent collapse (Fig. 1). This collapse
generates very high pressures and temperatures. The
chemical and physical effects of ultrasound are the results
of both stable and transient cavities.
Two different theories exist about cavitation: the hot-spot
and the electrical theory. The hot-spot theory assumes that
short-lived, localized hot-spots are generated in liquids.
Such a hot-spot can reach a temperature of roughly around
5000 8C, a pressure of 1000 atm, a lifetime less than a
microsecond and heating and cooling rates above 109 8C s1
[1].
According to the electrical theory, an electrical charge is
created on the surface of a cavitation bubble and enormous
electrical field gradients are formed across the bubble, being
able to break bonds upon collapse [7]. However, in 1996,
Lepoint-Mullie [8] discredited the electrical theory as a valid
mechanism behind sonoluminescence (SL) and sonochemistry. It seems very unlikely that SL arises from the
mechanism of bubble fragmentation during which ejected
microbubbles could accumulate electrical charges as
claimed by Margulis.
1.4.1. Homogeneous sonochemistry
1.4.1.1. Aqueous systems. The effect of ultrasound on
aqueous solutions has been studied for many years.

Sonification of aqueous solutions leads to the sonolysis of


the various components in the solution. For example,
sonolysis of water produces two radicals: H2O ! OH + H
[9]. Cavitational implosion produces free radicals in most
solvents and these radicals can recombine or react with other
species, to induce secondary reduction and oxidation
reactions [10].
Several reports dealing with utilization of acoustic
irradiation in hazardous chemical destruction and waste
water treatment were published in the recent years [1115],
since, ultrasonically generated free radicals in solvents can
be used to degrade toxic compounds.
Beneficial impact of acoustic irradiation on chemical
synthesis (heterogeneous or homogeneous) can be utilized in
connection to free radical formation under ultrasound, as it
promotes additional reaction pathways. Serpone and Terzian
[16] investigated ultrasonic induced dehalogenation and
oxidation of 2-, 3-, and 4-chlorophenol in air-equilibrated
media based on sonolysis mechanism of aqueous solution.
1.4.1.2. Organic systems. In case of organic systems, the
enhancing effect of ultrasound is not necessarily directly
related to thermal effects like in aqueous systems, but it is
rather a result of single-electron-transfer (SET) process
acceleration. The SET-step is required as the initial step in
some reactions, for example cycloadditions involving
carbodienes and heterodienes [17]. In systems where the
reaction mechanism does not require a SET-step, ultrasound
has a minor or no effect on the reaction [2].
1.4.2. Heterogeneous sonochemistry
1.4.2.1. Liquidliquid systems. Ultrasound forms very fine
emulsions in systems of two immiscible liquids. It means
that the surface area available for the reaction between two
phases is significantly increased and, therefore, also the
reaction rate increases [2]. Emulsions produced by
sonification are more stable than those formed conventionally [18].
1.4.2.2. Liquidsolid systems. When a cavitation bubble
collapses violently near a solid surface, liquid jets are
produced and high-speed jets of liquid are driven into the
surface of a particle (Fig. 2). These jets and shock waves can
cause surface coating removal, produce localized high
temperatures and pressures as well as improve the liquid
solid mass transfer [19]. Moreover, surface pitting may
result.
1.5. Sonoluminescence
When a liquid is irradiated by ultrasound, light is emitted
in some cases. This light-emission is called sonoluminescence (SL), which is basically emission of light associated
with cavitation. Sonoluminescence was first discovered in
1934 by Frenzel and Schultes in water [21]. There are two
forms of sonoluminescence: single bubble sonolumines-

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Fig. 3. The effect of ultrasonic power on reaction rate of iodine formation


[30].
Fig. 2. Collapse of a cavitation bubble near to a solid surface [20].

cence (SBSL) and multi-bubble sonoluminescence (MBSL)


[22]. The duration of a light flash is much more shorter in the
case of SBSL (100400 ps) than upon MBSL (3000 ps) [23].
No general consensus has been reached as to what is the
cause of this light emission although several theories have
been published. Walton and Reynolds believe that SL is
caused by recombination of free radicals generated within
cavitation bubbles during the collapse [24]. At the same
time, Suslick et al. favors an alternative theory claiming that
the light emission is caused by thermally created chemiluminescence [25].
1.6. Factors effecting cavitation
The ambient conditions of the reaction system (frequency, acoustic power, solvent, ultrasonic intensity, system
vapor pressure, external temperature and pressure, presence
and nature of dissolved gases, sample preparation, buffer)
can effect the intensity of cavitation, which influences the
reaction rates and/or yields [2]. In the following, these
factors are discussed in more detail.
1.6.1. Ultrasonic frequency
Frequency has a significant influence on cavitation since
by increasing the frequency the cavitation effect is
decreased. There are two explanations for this: (1)
rarefaction cycle of the sound wave produces a negative
pressure which is not sufficient in its duration and/or
intensity to initiate cavitation [2], (2) rarefaction and
compression cycles are too short to permit bubbles to reach a
sufficient size to collapse [26]. This is the reason why many
sonochemical reactions are carried out applying frequencies
between 20 and 50 kHz.
1.6.2. Acoustic power
With increasing power input to the reaction mixture, the
reaction rates increase to a maximum and then decrease with
additional power increase [27]. A probable explanation for
the observed decrease of the reaction rate, at high acoustic

powers, is the formation of a dense cloud of cavitation


bubbles near the surface of the horn tip, which acts as a
screen and reduces the amount of acoustic energy
transmitted to the fluid [28]. The optimum power input is
also dependent on the operating frequency [29]. Henglein
and Gutierrez [30] investigated the effect of ultrasonic power
on formation of iodine (Fig. 3) by using a commercial
20 kHz sound generator (KLN system 582, horn diameter
14 mm). Initially, the rate was increasing linearly, reaching a
plateau and sharply dropping at high power.
1.6.3. Solvent
Solvents with low vapor pressure and high surface tension
undergo intense cavitational effects, although cavities are
more easily formed in solvents with high vapor pressure, low
surface tension and low viscosity [2]. Solvents with high
vapor pressure decrease the effect of bubble collapse and
cavitation is inhibited. Since it is required that the negative
pressure in the rarefaction region must overcome the natural
cohesive forces acting in liquid and therefore it is more
difficult to produce cavitation in viscous liquids [26].
The effect of solvent on sonic enantioselective hydrogenation of 1-phenyl-1,2-propanedione (Fig. 4) over 5 wt.%
Pt/silica fiber (SF) was investigated (Table 1). The highest
enhancement of the rate and enantiomeric excess, compared
to silent conditions, was observed in mesitylene (the solvent
with the lowest vapor pressure).
The enantiomeric excess (ee) of B is defined as follows
ee = ([1R]  [1S])/([1R] + [1S]), where [R] and [S] are the
concentrations of (R)- and (S)-1-hydroxy-1-phenylpropanone, respectively.
1.6.4. Gas type and content
In general, dissolved gases with high specific heat ratio
usually give better cavitational effects than a gas with low
specific heat ratio since they convert more energy upon
cavitation [2]. For this reason, monoatomic gases such as
helium, argon and neon are preferred to diatomic gases such
as nitrogen and oxygen [26].

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

formation of the bubbles more difficult. On the other hand, at


lower temperatures higher rates are obtained in sonochemical processes [32]. In certain reactions systems, an optimum
reaction temperature may lead to more favorable results [2].
1.6.6. External pressure
With raising external pressure (Ph) cavitation threshold
and intensity of bubble collapse are increased. It is assumed
that there will no longer be a resultant negative pressure
phase of the sound wave (since Ph  PA > 0), therefore
cavitation bubbles cannot be created. However, a sufficiently
large increase in the intensity (I) of the applied ultrasonic
field could produce cavitation, even at higher overpressures,
since it will generate larger values of Pa (Eq. (1)) making
Ph  PA < 0. In that Pm (the pressure of collapsing bubble)
is approximately Ph + PA, increasing the value of Ph will
lead to a more rapid (Eq. (2)) and violent (Eq. (3)) collapse
[20]:
Fig. 4. Reaction scheme of 1-phenyl-1,2-propanedione hydrogenation: (A)
1-phenyl-1,2-propanedione, (B) (R)-1-hydroxy-1-phenylpropanone, (C)
(S)-1-hydroxy-1-phenylpropanone, (D) (S)-2-hydroxy-1-phenylpropanone,
(E) (R)-2-hydroxy-1-phenylpropanone, (F) (1R,2S)-1-phenyl-1,2-propanediol, (G) (1S,2S)-1-phenyl-1,2-propanediol, (H) (1S,2R)-1-phenyl-1,2-propanediol, (I) (1R,2R)-1-phenyl-1,2-propanediol.

The sonochemical effect also depends on thermal


conductivity of a gas: the greater the thermal conduction
of a gas, the more heat (formed in the bubble during the
collapse) will be dissipated to the surrounding liquid
resulting into an increased heat loss [31].
Gases with very high solubility in the reaction mixture
may reduce the intensity of a cavitation by possible
redissolving of the cavitational bubble before collapse [2].
1.6.5. External temperature
Generally, an increase in the ambient reaction temperature results in overall decrease in the sonochemical effect.
With increasing reaction temperature, the equilibrium vapor
pressure increases, which leads to easier bubble formation.
However, the cavitation bubbles which are formed, contain
more vapor [2]. The vapor reduces ultrasonic effect by
decreasing violent cavitation bubble collapse. Lower
temperature causes a higher viscosity, which makes the

Pa PA sin2pft

(1)

where Pa is the applied acoustic pressure, t the time, f the


frequency and PA is the pressure amplitude
 1=2 

Pvg
r
t 0:915Rm
1
(2)
Pm
Pm
where t is the collapse time, Rm the radius of the cavity at the
start of collapse, r the density of the medium and Pm is the
pressure in the liquid


Pm K  1 K=K1
Pmax P
(3)
P
where Pmax is the maximum pressure developed in the
bubble, Pm the pressure in the liquid at the moment of
transient collapse, P the pressure in the bubble at its maximum size and K is the polytropic index of the gas mixture.
Moulton et al. [33] observed, when investigating
hydrogenation of soybean oil at high hydrogen pressure
(14 bar), negligible enhancement of the catalyst activity
while at lower hydrogen pressure (8.5 bar) the ultrasonic
effect was significantly increased.
To ro k et al. [34] observed a similar trend when studding
cinamaldehyde hydrogenation over Pt/SiO2 catalyst. At
30 bar hydrogen pressure the enhancement of the catalyst
activity under ultrasound was almost negligible. When the

Table 1
Solvent effect on the initial hydrogenation rate and on enantiomeric excess (ee)
Solvent

Methyl acetate
Toluene
Mesitylene

Pvp (15 8C)


(bar)

DvapH (25 8C)


(kJ/mol)

Silent
Initial rate
( 103 mol/L gcat min)

ee (%)

Initial rate
( 103 mol/L gcat min)

ee (%)

0.1814
0.0220
0.0016

32.29
38.01
47.50

1.6
2.6
0.3

40
18
35

2.1
3.3
1.3

39
37
55

Ultrasound

Pvp is the vapor pressure of solvent at reaction temperature 15 8C; DvapH the molar enthalpy (heat) of vaporization at temperature 25 8C; ee the enantiomeric
excess (after 2 h of reaction).

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

pressure was decreased to 1 bar the catalyst activity was


significantly enhanced by ultrasound.
Comprehensive study of the influence of several
parameters such as hydrostatic pressure, nature of dissolved
gases frequency in degradation of phenol was reported by
Berlan et al. [15].
1.7. Heterogeneous catalyzed systems
Heterogeneous catalytic reaction systems are industrially
the most important class of reactions. A large number of
products ranging from bulk chemicals to alimentary and
pharmaceuticals are produced by means of heterogeneous
catalysis. This is the reason why quite many heterogeneous
systems could be found in the literature, in order to probe the
effect of ultrasound. Table 2 contains several representative
examples, which demonstrate the positive effects of
ultrasound on activity and selectivity. Some of the results
are commented below.
Lindley et al. investigated Ullmann [35] coupling of
2-iodonitrobenzene over Cu-powder. The positive effect of
copper powder presonification was reported. Under ultrasound the yield increased from 1.5 to 70.4%.
Li et al. [36] carried out a reaction of chalcones with
ethyl cyanoacetate catalyzed by KF/basic alumina in
anhydrous ethanol. The mixture was irradiated in a water
bath of an ultrasonic cleaner, at 2534 8C. Shorter reaction
time and higher yield was observed compared to silent
conditions.
Shin and Han [37] hydrogenated diphenylacetylene over
palladium hydrazine. The mixture was exposed to sonification in common ultrasound laboratory cleaning bath at
25 8C. A higher yield was observed under ultrasound.
To ro k et al. [38] investigated sonochemical enantioselective hydrogenation of ethyl pyruvate over 3 wt.% Pt/SiO2
catalyst in ethanol and toluene. The hydrogenation reaction
was carried out at room temperature (25 8C). The catalyst
was activated in a closed system by stirring it for 1 h or in an
ultrasonic bath for 30 min under atmospheric hydrogen
pressure. After the activation period ethyl pyruvate was
injected into the reactor. The experimental results indicated
that the acoustic irradiation of catalyst modifier system
priori to the reaction is a highly efficient method to enhance
enantioselectivity. However, such enhancement was very
solvent specific, as the catalyst activity was increased by
sonification in ethanol solvent, only.
Cains et al. [39] carried out hydrogenation of oct-1-ene
over 10% Pd/C catalyst, at 25 8C, in absolute ethanol.
Presonification of this catalyst led to a considerable
enhancement in activity resulting in 100% conversion after
only 30 min for the sonicated system. Only 66% conversion
was achieved under silent conditions.
Sulman [40] studied hydrogenation of a carbonyl group
of ethyl 9-(2,3,4-trimethoxy-6-methylbenzoyl)nonanoate on
Pt/sibunit catalyst. The catalyst was subjected to the
ultrasonic irradiation (ultrasonic flux density 1 W cm2)

for 30 s. A yield enhancement was observed by use of


ultrasonic pretreated catalyst. The increase in the ultrasonic
flux density from 0.1 to 1 W cm2 favored the catalyst
activity, while its further increase to 3 W cm2 resulted in
the decrease of activity.
Ando et al. [41] investigated ultrasonic irradiation of a
mixture of benzyl bromide, potassium cyanide and alumina
in an aromatic solvent. When benzyl bromide was treated
with potassium cyanide and alumina in toluene under
mechanical agitation at 50 8C, only a mixture o- and pbenzyltoluene was obtained in 75% yield. Unexpectedly,
alumina had not promoted the cyanide substitution reaction
but enhanced the FriedelCrafts reaction of benzyl bromide
with toluene. In contrast, when the same reaction mixture
was irradiated by ultrasound (ultrasonic bath) at 50 8C, the
substitutions did take place to afford benzyl cyanide in 71%
yield. Thus, ultrasound completely switched the reaction
course from the FriedelCrafts reaction to nucleophilic
substitutions.
Han and Boudjouk [42] studied Reformatsky reaction
over Zn at 2530 8C under ultrasonic irradiation (ultrasonic
bath). A significant enhancement of the yield under
ultrasound was observed. Several examples of different
Reformatsky reactions and solvents comparisons were
reported in this paper.
Fuentes et al. [43] investigated synthesis of chalocones by
ClaisenScmidt condensation over activated barium hydroxide C-200, in ethanol at 25 8C. Substantial improvement of
the yield by catalyst activation in the presence of ultrasound
(ultrasonic bath) was noticed.
Boudjouk and Han [44] carried out hydrogenations of
olefins using formic acid as the hydrogen transfer agent over
10% Pd/C, at 25 8C, with toluene solvent under ultrasound
irradiation (ultrasonic cleaning bath). They reported
significant enhancement of the hydrogenation rate in the
presence of sonic waves.
Szo llo si et al. [45] investigated selective hydrogenation
of cinnamaldehyde to cinnamyl alcohol, at a reaction
temperature of 60 8C and hydrogen pressure of 1 bar, in
isopropyl alcohol. Several heterogeneous Pt catalysts on
different supports were tested: 5% Pt/C, 5% Pt/Al2O3,
EuroPt-1 (6.3% Pt/SiO2), and 3% Pt/SiO2. The catalysts
were presonificated in an ultrasonic bath, at 25 8C, for
10 min. As the data indicates, the reaction rates decreased
over 5% Pt/C and 5% Pt/Al2O3, while the increase was
negligible over EuroPt-1 (6.3% Pt/SiO2) and significant over
3% Pt/SiO2. Parallel, the cinnamyl alcohol became the
major product of the hydrogenation over 3% Pt/SiO2.
Han and Boudjouk [46] studied sonically accelerated
hydrosilation of alkenes (several reactions were reported).
The experiments were carried out in a single-necked flask
submerged to an ultrasonic laboratory cleaner (Bransonic
model 220). Their results showed that ultrasonic waves
dramatically increased the activity of Pt/C in the hydrosilation reactions, giving yields comparable to those
obtained with a homogeneous catalyst. In each case, the

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Table 2
Organic synthesis over heterogeneous catalysts studied in the presence of ultrasound
Reference Reaction

Conventional conditions

Ultrasonic conditions

[35]

Ullman coupling of 2-iodonitrobenzene:

Stirring; temperature: 63 8C;


reaction time: 2 h; yield: 1.5%

Probe system;
power: 135 W;
temperature: 63 8C;
reaction time: 2 h;
yield: 70.4%

[36]

(copper power presonificated for 15 min)


Michael reaction:

[37]

(catalyst: KF/basic alumina)


Hydrogenation of diphenylacetylene:

[38]

Enantioselective hydrogenation of ethyl pyruvate:

[39]

(solvent: ethanol, toluene; catalyst: 3 wt.% Pt/SiO2)


Hydrogenation of oct-1-ene:

[40]

Bath system; temperature:


2534 8C; reaction time:
1.5 h; yield: 86%

Magnetic stirring; temperature: 25 8C;


reaction time: 1 h; yield: 24%

Bath system; temperature:


25 8C; reaction time:
1 h; yield: 100%

Stirring; temperature: 25 8C;


solvent: ethanol; yield: 95% (600 min);
ee: 20.2%; solvent: toluene;
yield: 95% (73 min); ee: 24.8%

Bath system; temperature:


25 8C; solvent: ethanol;
yield: 100% (52 min);
ee: 34.2%; solvent:
toluene; yield: 100%
(110 min); ee: 50.5%

(catalyst: 10 wt.% Pd/C)

Stirring; temperature: 25 8C;


solvent: ethanol; reaction time:
30 min; yield: 66%

Hydrogenation of ethyl
9-(2,3,4-trimethoxy-6-methylbenzoyl)nonanoate:

Stirring; no catalyst pretreatment;


reaction time: 80 min; yield: 55%

Probe system; temperature:


25 8C; solvent: ethanol;
reaction time: 30 min;
yield: 100%
Ultrasonic catalyst
pretreatment; reaction
time: 80 min; Y: 85%
(1 W cm2); Y: 65%
(0.1 W cm2); Y: 75%
(3 W cm2)

[41]

(catalyst: Pt/sibunit)
Change in pathway from the FriedelCrafts reactions to
nucleophilic substitutions:

[42]

Reformatsky reaction:

Stirring; temperature: 50 8C;


product: o- and p-benzyltoluene;
yield: 75%

Bath system; power:


200 W; product:
benzyl cynanide;
yield: 71%

Stirring; temperature: 80 8C;


reaction time: 12 h; yield: 50%

Bath system;
temperature: 2530 8C;
power: 150 W; reaction
time: 30 min; yield: 98%

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Table 2 (Continued )
Reference Reaction

Conventional conditions

Ultrasonic conditions

[43]

Stirring; temperature: 25 8C;


solvent: ethanol; catalyst mass:
1.0 g; reaction time: 60 min; yield: 5%

Bath system;
temperature: 25 8C;
solvent: ethanol;
catalyst mass: 0.1;
reaction time: 10 min;
yield: 76%
Bath system;
temperature: 25 8C;
solvent: ethanol;
reaction time: 1 h;
yield: 100%

Synthesis of chalcones by ClaisenSchmidt condensation:

(catalyst: activated barium hydroxide C-200)


[44]

Hydrogenation of olefins:

[45]

(catalyst: 10 wt.% Pd/C)


Selective hydrogenation of cinnamaldehyde to cinnamyl alcohol:

(solvent: isopropyl alcohol; catalyst: 5% Pt/C,


5% Pt/Al2O, EuroPt-1, 3% Pt/SiO2)

[46]

Hydrosilation reaction:

[47]

(catalyst: Pt/C)
Dechlorination of polychlorinated biphenyls:

[48]

(catalyst: 10% Pd/C)


FriedelCrafts acylation of 2-methoxynaphtalene:

Stirring; temperature: 25 8C;


reaction time: 1 h; yield:
maximum 70%

Stirring; temperature: 25 8C;


rate (5% Pt/C):
1
60  103 mol g1
Pt h ; rate (5%Pt/Al2O3):
3
1
40  10 mol gPt h1; rate
1
(EuroPt-1): 25  103 mol g1
Pt h ;
1
rate (3%Pt/SiO2): 31  103 mol g1
Pt h ;
selectivity (5% Pt/C): 42 mol%; selectivity
(5%Pt/Al2O3):47 mol%; selectivity
(EuroPt-1): 58 mol%; selectivity
(3%Pt/SiO2): 40 mol%

Bath system; presonification 10 min;


temperature:
25 8C; rate (5% Pt/C):
1
36  103 mol g1
Pt h ;
rate(5%Pt/Al2O3):
1
32  103 mol g1
Pt h ;
rate (EuroPt-1):
1
29  103 mol g1
Pt h ;
rate (3%Pt/SiO2):
1
72  103 mol g1
Pt h ;
selectivity (5% Pt/C):
37 mol%; selectivity
(5%Pt/Al2O3):47 mol%;
selectivity (EuroPt-1):
78 mol%; selectivity
(3%Pt/SiO2): 79 mol%
Bath system;
temperature:
30 8C; reaction time:
1.5 h; yield: 94%

temperature: 25 8C; solvent: hexane;


reaction time: 45 min; yield:
maximum 100%

Bath system;
temperature:
25 8C; solvent: hexane;
reaction time: 15 min;
yield: 100%

Conventional method; Amberlyst36:


conversion (20 min): 25%;
Amberlyst15: conversion (20 min):
22%; Indion-130: conversion (20 min):17%

Probe system;
temperature: 25 8C;
Amberlyst36:
conversion (20 min):
65%; Amberlyst15:
conversion (20 min):
50%; Indion-130:
conversion (20 min):
60%
Probe system;
temperature: 70 8C;
solvent: isopropanol;
reaction time: 50 min;
yield: 60%

(catalyst: Amberlyst-36, Amberlyst-15,


Indion-130, K-10 clay, Filtrol-24)
[49]

Hydrogenation of citral to citronellal:

(catalyst: RaNi)

Stirring; temperature: 70 8C;


solvent isopropanol; reaction time:
50 min; yield: maximum 50%

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Table 2 (Continued )
Reference Reaction

Conventional conditions

Ultrasonic conditions

[50]

Stirring; temperature: 15 8C;


1
rate (5%Pt/SF): 0.3  103 mol g1
Pt h ;
ee (2 h): 35%; rate (5% Pt/Al 2O3):
1
1.7  103 mol g1
Pt h ;
ee (5%Pt/Al2O3): 52%

Probe system;
temperature: 15 8C;
rate (5%Pt/SF):
1
1.3  103 mol g1
Pt h ;
ee (2 h): 55%; rate
(5% Pt/Al2O3):
1
2.0  103 mol g1
Pt h ;
ee (5%Pt/Al2O3): 52%

Enantioselective hydrogenation of 1-phenyl-1,2-propandione:

(solvent: mesitylene; catalyst:


5 wt.% Pt/ silica fiber (SF), 5 wt.% Pt/Al 2O3)

reaction proceeded smoothly and without vigorous


exotherm.
Gonzalo Rodrguez and Lafuente [47] found a new
heterogeneous catalytic method for the complete dechlorination of PCBs using an ultrasound source. They studied
dechlorination of 2,4,5-trichlorobiphenyl (PCB-29) with
solid hydrazine hydrochloride and sodium carbonate over
Pd/C catalyst in hexane, at 60 8C (silent conditions) and
25 8C (ultrasonic irradiation). The dechlorination of PCB-29
by ultrasonication at 487 kHz showed remarkably high
efficiency.
Yadav and Mujeebur Rahuman [48] investigated sonic
intensification of FiedelCrafts acylation of 2-methoxynaphtalene with acetic anhydride under several acid treated
clays (such as K-10, Filtrol-24) and cation exchange resins
(such as Amberlyst-36, Amberlyst-15 and Indion-130) [48].
The reaction was carried out in a probe system at a
temperature of 25 8C. The results indicated that ultrasound
effectively promotes the acylation of 2-methoxynaphtalene
in the presence of a solid acid catalyst. Sonification did not
alter the selectivity but only intensified the reaction rates.
Mikkola et al. [49] studied hydrogenation of citral over
RaNi catalyst in isopropanol, at the operating pressure
50 bar and at the temperature of 70 8C. The results indicated
that the reaction rate, as well as the selectivity to the desired
end products; citronellol and citronellal were enhanced by
the use of ultrasound irradiation.
Toukoniitty et al. [50] carried out enantioselective
hydrogenation of 1-phenyl-1,2-propanedione over cinchona
alkaloid modified 5 wt.% Pt/silica fiber (SF) and 5 wt.% Pt/
Al2O3 catalyst. The reaction was carried in mesitylene at
15 8C and at the hydrogen pressure of 10 bar. In the case of
5 wt.% Pt/silica fiber (SF) catalyst, an acceleration of the
reaction rate was observed and the enantiomeric excess was
significantly higher under ultrasound. However, with Pt/
Al2O3 application of ultrasound resulted only in a moderate
enhancement of the reaction rate.
In general, it can be concluded, that acoustic irradiation
was found to be a beneficial tool for enhancing the reaction
rate and selectivity. There are several explanations for
observed phenomenon. The catalyst activity is increased due
to the deformation of the catalyst surface that exposes fresh,
highly active surface and reduces the diffusion length in the
catalyst pores (Figs. 5 and 6). Furthermore, ultrasound is
capable of cleaning and smoothening the catalyst surface
(Figs. 7 and 8). The local turbulent flow associated with

Fig. 5. SEM image of used RaNi catalyst treated under ultrasound.

Fig. 6. SEM image of used RaNi catalyst without sonification.

acoustic streaming improves mass transfer between the


liquid phase and the surface, thus increasing the observed
reaction rates [1]. Finally, sonification of aqueous solutions
leads to the sonolysis of the various components in the

Fig. 7. SEM image of used 5 wt.% Pt/silica fibre (SF) catalyst treated under
ultrasound.

10

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Fig. 8. SEM image of used 5 wt.% Pt/silica fibre (SF) catalyst without
sonification.

solution and free radicals are formed. Ultrasonically


generated free radicals in solvents can be used to degrade
toxic compounds and, therefore, acoustic irradiation is
widely used in waste water treatment and oxidation
reactions.
Acoustic irradiation can also be utilized in the process of
catalyst preparation, in order to improve the catalyst activity
and enhance the crystallization process.
Synthetic zeolites are prepared hydrothermally by
heating aqueous solutions of sodium silicate and sodium
aluminate, at a temperature range of 25300 8C for few
hours to several days. Lindley [51] found out that in
synthesis of zeolite NaA, the use of ultrasound led to
substantial reductions in nucleation time (1 h, probe
system) and overall completion times (3.4 h, probe system)
compared to control reactions (5 h nucleation time and 10 h
completion time).
Bianchi et al. [52] investigated the activity of ultrasonically prepared Co/SiO2 catalyst in FisherTropsch
synthesis. The catalyst was prepared by direct addition of
the aqueous solution of cobalt nitrate to the silica and
followed by sonification for 1 h at room temperature and
then reduced in flowing hydrogen. Catalyst prepared under
) and higher metal
US showed smaller crystallite size (38 A
dispersion compared to the catalyst prepared by classical
incipient wetness impregnation method (crystallite
). The catalyst prepared by using US showed
size = 69 A
quite good stability deactivating by less than 10% in 4 days.
On the contrary, the catalyst prepared by a classical
impregnation method deactivated rapidly with the loss in the
conversion of 40% in 4 days.
Li and Inui [53] studied the effect of acoustic irradiation
on copper/zinc/alumina oxides catalyst, during co-precipitation and aging. The catalytic activity of the catalyst was
evaluated in methanol synthesis. Irradiation of the suspension during co-precipitation and aging steps, in the
preparation of copper/zinc/alumina catalyst, appreciably
enhanced the activity for methanol synthesis. This enhancement is proposed to stem from the promotion by the
ultrasound of the formation of the hydrotalcite-like phase

(56.4%) in the precursor of the catalyst compared to the


classical method (37.5%).
Pillai et al. [54] exposed vanadium phosphorus oxide
(VPO) to acoustic irradiation in one of the steps of the
catalyst preparation (precipitation) and studied its phase
composition and activity for hydrocarbons oxidation. It
was demonstrated that active VPO catalyst could be
prepared in a relatively short time (36 h) by employing
ultrasound compared to a conventional method (CM)
(20 h). At the same time, there was no deterioration of the
catalytic properties: both catalysts had similar compositions
2 1
and surface areas (US = 10.3 m2 g1
cat , CM = 10.2 m gcat )
and displayed comparable oxidation activity for liquid
phase hydrogenperoxide oxidation of cycloalkanes in
acetonitrile.
1.8. Other applications
1.8.1. Polymerization reactions
The effect of ultrasound on polymerization reactions has
been studied for about 30 years. Most studies deal with the
radical polymerization of vinyl monomers where sonification makes the addition of thermal initiators unnecessary and
allows for better control over the molecular weight, tacticity
and polydispersity [55].
Ultrasound accelerates emulsion polymerization because
of its ability to create fine emulsions.
Not only emulsion, but suspension and polymerization as
well, is enhanced by sonification since ultrasound prevents
agglomeration of monomer droplets [56].
Price [57] has shown that the reaction rate in step-growth
reaction can be accelerated by the use of high intensity
ultrasound. In this case, the source of enhancement seems to
be related to local heating produced by collapsing cavitation
bubbles.
1.9. Ultrasonic systems
For laboratory research purposes, two basic methods of
applying acoustic power to liquid loads are used: low
intensity systems 12 W cm2 (an ultrasonic bath) and high
intensity systems affording hundreds of W cm2 (a horn
system).
1.9.1. Ultrasonic bath
Ultrasonic baths were originally produced for cleaning
purposes. However, nowadays, ultrasonic baths are widely
used in sonochemistry because lower acoustic intensities are
often required in liquidsolid heterogeneous systems. For
these reaction systems, common ultrasonic cleaning baths
are sufficient [58]. They are easily available and rather
inexpensive.
1.9.2. Horn (probe) system
Horn systems (Fig. 9) are getting more popular in
sonochemical laboratory research. These systems are able to

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

11

wavelengths of 1 cm to 1 m. Microwave heaters use specific,


fixed frequencies 2.45 GHz (wavelength 12.2 cm) or
900 MHz (wavelength 33.3 cm), in effort to avoid interferences with RADAR (wavelength from 1 to 25 cm) and
telecommunication applications. All domestic microwave
ovens operate at the frequency of 2.45 GHz.
In comparison to the conventional heating, energy
transfer does not occur by convention and conduction but
by dielectric loss in case of microwave heating.
2.2. History

Fig. 9. High-pressure autoclave with in situ ultrasonic irradiation system


(horn system).

deliver large energy intensities of ultrasound power directly


to the reaction mixture, which can be regulated by varying
the amplitude delivered by the transducer [2]. A disadvantage of using a horn system is the possibility of
contamination of the reaction solution caused by erosion
and pitting of the horn tip [59].

During the World War II, Randall and Booth working at


the University of Birmingham, in connection to the
development of RADAR, designed a magnetron to generate
microwaves [61].
As with many other great inventions, the microwave oven
was a by-product of another technology. In 1946, Percy
Spencer realized that a candy bar in his pocket melted during
the tests of a vacuum tube called magnetron [62]. The
widespread use of domestic microwave ovens appeared
during 1970s, in the United States and Japan.
Microwave ovens have been used for cooking and
defrosting of food for past three decades. Gedye et al. [63]
published the first pioneering report about utilizing
microwave irradiation in chemical synthesis, in 1986.
During the last decade, microwave heating has been
increasingly applied in carrying out organic synthesis.

2. Microwave chemistry
2.3. Industrial application
2.1. Introduction
Microwave (MW) irradiation is a form of electromagnetic energy. Microwaves consist on an electric component
as well as a magnetic one (Fig. 10). The microwave region of
the electromagnetic spectrum is situated between infrared
radiation and radio frequencies. Microwave irradiation
ranges from 30 GHz to 300 MHz and corresponds to the

Most of the industrial applications of electromagnetic


heating are found when a change of state of non-conductive
matter is involved (e.g. defrosting, dehydration with the
change of state of water) [64]. Other important use of MW is
sintering and fusion of solids. Microwave heating is also
used in food industry for drying fruits, berries, corn, etc.
[65]. In medicine, microwaves are utilized in resonance
therapy and microwave thermography.
2.4. Microwaves in chemistry
Microwave irradiation, in connection to, chemistry is
applied in two different areas, namely microwave spectroscopy and microwave dielectric heating.

Fig. 10. Electromagnetic irradiation: electric and magnetic components


[60].

2.4.1. Microwave spectroscopy


There is a growing interest in new analytical instruments
to quantify and identify the constituents of gas samples
obtained from automobile exhausts, factory emissions, and
process streams. Microwave spectroscopy offers high
sensitivity, accurate quantification, and unambiguous species identification and, furthermore, it is compact and easy to
use.
Microwave spectroscopy is a major method in physical
chemistry to study the molecules in the gas phase.

12

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Microwave spectra arise when the rotational energy of a


compound is changed. These energy changes are rather
small, only few joules per mole. The absorption peaks are
unusually narrow and frequency measurements can routinely be carried out with an accuracy of as much as 710
significant digits. From these moments, it is possible to
calculate the geometrical structure of a molecule more
accurately than with any other method.
Microwave spectroscopy provides excellent fingerprinting technique for identifying molecules in the gas
phase. The most important analytical application of
microwave spectroscopy so far has been to confirm the
presence of a wide range of molecules in outer space
[61].
2.4.2. Dielectric polarization
Microwave dielectric heating depends on the ability of an
electric field to polarize charges in materials and their
inability to follow rapid changes of an electric field space
[61].
The total polarization is a sum of several components:
at ae aa ad ai

(4)

where ae is the electronic polarization, aa the atomic polarization, ad the dipolar polarization and ai is the interfacial
polarization (MaxwellWagner effect).
The time scale of the electronic and atomic polarization/
depolarization is much smaller than microwave frequencies,
therefore they do not contribute to microwave dielectric
heating.
Microwave energy can effect molecules in two principal
ways: (a) dipolar polarization and (b) by ionic conduction. A
third mechanism (c) interfacial polarization, can also occur,
although it has often limited importance.
The dielectric loss tangent defines the ability of a material
to convert electromagnetic energy into heat energy, at
certain frequency and temperature:
tan d

e00
e0

(5)

where e0 is the dielectric constant, describes the ability of a


molecule to be polarized by electric field and e00 is the
dielectric loss, describes the efficiency with which the
energy of the electromagnetic irradiation can be converted
into heat.
Both parameters, the dielectric loss and dielectric
constant are measurable properties.
2.4.2.1. Dipolar polarization. When microwave energy is
passing through the matter, molecules of the matter having
dipole moments will rotate and try to align themselves with
the electric field [61]. Polar molecules have stronger
interactions with the electric field. Polar ends of the
molecules tend to align themselves and oscillate in step with
the oscillating electric field of microwaves. Collisions and
friction between the moving molecules results in heating.

Generally, the more polar a molecule is, the more effectively


it will couple with microwaves.
2.4.2.2. Ionic conduction. Ionic conduction is just minimally different from dipole interaction. In the presence of
the electric field, ionic species will migrate in one or the
other direction, depending on the electric field. During this
migration, energy is transferred from the dielectric field
causing ionic interactions that speed up the heating of the
solution [66]. Ionic conduction increases with temperature,
thus allowing ionic solutions to become stronger absorbers
of microwave energy as they are heated.
For example, if pure water is heated in microwave oven,
mainly the polarization mechanism occurs and the heating
rate is significantly lower than in a case when a salt is added
and both mechanisms occur and contribute to the heating
effect.
2.4.2.3. Interfacial polarization. A suspension of conducting particles in a non-conducting medium is an inhomogeneous material whose dielectric constant is frequency
dependent. The loss relates to the build up of charges
between interfaces and is known as the MaxwellWagner
effect. As the concentration of conducting phase is
increased, a point is reached when individual conducting
areas contribute: this has been developed by Maxwell
Wagner in the two-layer capacitor model [66].
2.5. How materials interact with microwaves
Materials interact with microwaves in three ways
(Fig. 11). Metals are good conductors because they tend
to reflect microwave energy and they do not warm up that
well. Transparent materials are good insulators because they
are transparent to microwave energy and they do not warm
up. Absorbing materials receive microwave energy and are
heated [66]. These different materials interactions with
microwaves enable selective heating.
The advantage of microwave irradiation as an energy
source for heterogeneously catalyzed systems is that
microwaves do not in many cases substantially heat up
the adsorbed organic layers but interact directly with the
metal sites on catalyst surface and hot-spots might be
created. The temperature of the reactive sites was calculated
to reside 918 K above the bulk temperature [67].

Fig. 11. Interaction of transparent (insulator), absorbing (dielectric),


reflecting (conductor) materials with microwave energy [68].

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122


Table 3
Temperatures of several liquids (50 mL) after heating from room temperature for 1 min at 560 W and 2.45 GHz [61]
Liquids

t (8C)

Bp (8C)

Water
Methanol
1-Propanol
Acetic acid
Ethyl acetate
Acetone
Hexane
Heptane
CCl4

81
65
78
110
73
56
25
26
28

100
65
78
119
77
56
68
98
77

2.6. Microwave heating of liquids and solids


The rate of temperature increase due to the dielectric field
of microwave radiation is determined by the following
equation [61]:
2
dT
e00 fEr:m:s:
k
dt
rcp

(6)

2
is the r.m.s. field intensity, r the density, cp the
where Er:m:s:
specific heat capacity, k the constant, e00 the dielectric loss
and f is the frequency.
Depending on the parameters in Eq. (6) the temperature
increase can be quite substantial.
The main advantage of microwave heating is the
instantaneous heating of solids and liquids, which is not
the case of conventional heating (see Tables 3 and 4).
Organic and inorganic solvents with low molecular
weights and high dipole moments (e.g. H2O, MeOH, EtOH)
couple effectively with microwaves at 2.45 GHz. At the
same time, non-polar organic solvents have negligible
dielectric losses and therefore they do not effectively couple
with microwaves [61]. If polar solvents are used, the main
interactions occur between microwaves and polar molecules
and it would be expected that any specific microwave effects
related to the reactants and products would be masked by the
solventmicrowave interaction. The reaction rates should,

Table 4
Temperatures of several solids (25 g) heated from room temperature at
1 kW oven and 2.45 GHz with 1000 mL vented water load [61]
Solids

t (8C)

Time (min)

Al
C
CO2O3
Ni
SnCl2
CaO
PbO2
TiO2
WO3

577
1283
1290
384
476
83
182
122
532

6
1
3
1
2
30
7
30
0.5

Authors comment [61]: most of these values were obtained in multi-mode


cavity with operation cycles (domestic microwave oven) and cannot therefore, be considered to provide more than indicative values of the temperatures reached in a single-mode equipment with a homogeneous and
continuous field strength.

13

therefore, be basically the same as those observed under the


conventional conditions. This hypothesis has been confirmed recent study of esterification reactions, where no
enhancement by microwaves irradiation was observed
compared with the conventional reaction at the same
reaction temperature [69].
In 1992, Baghurst and Mingos showed that microwaves
heat solvents above their normal boiling points. Water
heated by microwaves reaches 105 8C before boiling. Even
higher temperature difference (38 8C above the normal
boiling point) can be achieved with acetonitrile, since
superheating to 120 8C was reported. This superheating is
one of the reasons why reactions proceed faster [70].
Microwave effects are most likely to be observed under
solvent free reaction conditions [71] as absorption of
microwave irradiation is limited only to the reactive species.
The possible specific effects will not be moderated or
impeded by solvents.
Microwave effects depend on microwavematerial
interactions, the more polar is a molecule, more significant
is the microwave effect when the rise in temperature is
considered. In terms of kinetics and reactivity, the specific
heat has therefore be considered according to reaction
mechanism and, particularly, how polarity of the system is
altered during the progress of the reaction [72]. Specific
microwave effects can be expected for a polar mechanism,
when the polarity of reaction is increased during the reaction
from the ground state toward the transition state [73]. The
outcome is essentially dependent on the medium and
reaction mechanism.
Several articles contain examples of increased selectivity
in which the chemo- or regioselectivity can be increased
under the microwave dielectric heating compared to
conventional heating [74,75]. It can be foreseen that
microwave effects could be important in determining the
selectivity of some reactions. When competitive reactions
are involved, the ground state is common for both processes.
The mechanism occurring via the more polar transition state
could, therefore, be favored under the dielectric heating [72].
Finally, it is important to note, that microwave heating is
rapid because energy is absorbed directly by the material.
The temperature of the reaction mixture is uniform,
especially in comparison with conventional heating [76].
2.7. Microwaves specific and athermal effects in chemistry
Specific microwave effects could basically be of two
origins: those, which are not purely thermal, and a special
thermal effect connected with possible intervention of hotspots [72]. One of the few theoretical papers trying to
explain acceleration under microwaves has been published
by Miklavc [77]. The author claims that large increases in
chemical reaction rates occur due to the effects of rotational
excitation and on collision geometry.
Concerning microwave dielectric heating, there is a
controversy, whether specific athermal effects exist. Several

14

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

questions still await detailed theoretical consideration. Are


electromagnetic fields able to enhance or to modify
collisions between reactants? Do reactions proceed at the
same reaction rate under microwave or conventional heating
for the same bulk temperature? What changes are caused by
electromagnetic field on the molecular level? Since the
beginning of the use of microwaves in chemical synthesis,
athermal effects have been assumed. However, recent
experiments with proper temperature measurements indicate
a failure of these athermal effects. Many of these prejudices
have arisen from inaccurate knowledge of the temperature
distribution due to the difficulty in measuring temperature in
irradiated medium [78].
The pioneering report claiming athermal effects was
published by Sun et al. [79] on hydrolysis of phospho
anhydride bonds in nucleotide triphosphates. The reaction
temperature seemed to be the same for microwave and
conventional heating. However, the hydrolysis rate was
almost 15 times higher under microwaves. These results
could prove that it is possible to obtain localized absorption
of microwaves without significant increase of temperature.
But later on, Jahngen et al. [80] referred that hydrolysis rate
enhancement observed by Sun et al. was caused by thermal
gradients and not special athermal effects.
Raner and Strauss [81] clearly pointed out the problem of
kinetic investigation. In order to perform kinetic studies the
exact temperature must be known, and thermal gradient
must be modeled. Esterification reaction was carried out
under microwave dielectric heating and conventional
heating (oil bath). Additional results showed [81] that the
final yield of ester depended only on temperature but not on
the mode of heating.
These examples clearly demonstrate that it is very
important to carefully monitor reaction temperature directly
in the electromagnetic field.
According to Gedye et al. [82,83] it is possible to observe
different product composition under microwave and conventional heating. However, probable explanation for this
phenomenon is that microwave heating significantly
increases the reaction temperature and it is possible that
the reaction temperature in case of dielectric heating could
exceed the ignition temperature for an additional reaction,
which is not possible at the lower temperatures achieved by
conventional heating.
Although it seems that at present moment there are no
athermal effects in connection to microwaves dielectric
heating, electromagnetic irradiation in chemical reactions
still remains largely unknown and all the potentials of
microwaves in chemical and catalytic applications have to
be discovered in the future.

2.8. Microwaves application in heterogeneous catalysis


Several examples, highlighting application of microwaves in heterogeneous catalysis are presented in Table 5.

Pillai et al. [84] studied MW-assisted olefin epoxidation


over hydrotalcite catalyst, using excess of acetonitrile.
Apparently, the selective epoxidation of a variety of olefins
occurs more rapidly upon microwave irradiation than by
conventional heating in an oil bath. This is presumably due
to the polar nature of the reaction intermediates that couple
efficiently with microwaves and causes the dramatic rate
acceleration. This approach significantly minimizes the
longer reaction times required in conventional heating of
olefins with H2O2 and avoids the use of large excess of
volatile organic solvent usually employed.
Ha jek et al. [85] investigated rearrangement of 1,1,3triphenylpropargyl alcohol under microwave irradiation
(700 W) as well as conventional conditions in oil bath
(100 8C), in the presence of 10 g neutral alumina (Al2O3).
Conversion of 1,1,3-triphenylpropargyl alcohol in nonstirred sample (temperature gradient 18 8C) was higher
(three times) under microwave exposure compared to the
conventional heating. Conversion in stirred sample was
almost the same under microwave and conventional heating.
Local superheating effect is responsible for the acceleration
of the reaction as indicated by these results.
Huber and Jones [86] studied the acceleration of an OrthoEster Claisen rearrangement by clay-catalyzed microwave
thermolysis. Conventional thermolysis was carried out in
three-necked round bottom flask. The assembly was flushed
with dry nitrogen gas and then triethyl orthoacetate was
introduced, followed by 2-cyclohepten-1-ol and, finally,
propionic acid. The mixture was heated using an oil bath to
140 8C. In the case of microwave theromolysis, montmorillonite KSF was suspended in dry DMF in a flame dried
sealed tube. 2-Cyclohepten-1-ol was added under stream of
nitrogen, followed by triethyl orthoacetate. The mixture was
exposed to irradiation for 9 min (500 W) in a commercial
microwave oven. Under traditional conditions, only moderate
yields of the product were recovered, contaminated with
varying amounts of starting alcohol and their derived acetates
even after forcing conditions (12 h, 180 8C). By microwave
thermolysis, high yields of clean product were obtained.
Application of microwave heating accelerates many ring
opening reactions. For example, Villemin and Labiad [87]
investigated the opening of phenylthio dichlorocyclopropanes
in the presence of AgBF4 in ethanol, under conventional
heating and, alternatively, without solvent (dry reaction),
catalyzed by AgBF4/Al2O3 under microwave irradiation
(microwave oven, 650 W). Microwave heating allows us
working under solvent free conditions and has potential for
significant acceleration of the reaction rates.
Calinescu et al. [88] carried out catalytic dechlorination
of chlorobenzene in a solution of NaOH in 2-propanol, at
boiling temperature. It was found that Pd (supported alumina
or aluminum silicate) is a very active catalyst. The
microwave heating had very favorable effect: the conversions were almost double in comparison to classical heating.
Banik et al. [89] carried out hydrogenation of b-lactam in
ethylene glycol and ammonium formate over 10% Pd/C

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

15

Table 5
Heterogeneous catalyzed organic reactions enhanced by microwave irradiation
Reference Reaction

Conventional conditions

Microwave conditions

[84]

Olefin epoxidation:

Oil bath; solvent:


acetonitrile; reaction
time: 5 h;
conversion: 60%

Multi-mode; power:
720 W; solvent:
acetonitrile; reaction time:
5 min; conversion: 96%

[85]

(catalyst: hydrotalcite (Mg6Al2CO3(OH)16


4H2O)
Rearrangement of 1,1,3-triphenylpropargyl alcohol:

Oil bath; stirring;


temperature: 100 8C;
reaction time: 15 min;
yield: 6.2%

Multi-mode; no stirring;
power: 700 W; temperature:
100 8C; reaction time:
15 min; yield: 21.5%

[86]

Ortho-Ester Claisen rearrangement:

[87]

(catalyst: montmorillonite KSF clay)


Opening of cyclopropanes:

[88]

(catalyst: AgBF4 on Al2O3)


Dechlorination of chlorobenzene:

[89]

(catalyst: Pd/AlSi, Pd/Al2O3)


b-Lactam formation:

[90]

(catalyst: 10% Pd/C)


Dehydrogenation of citronellal:

(catalyst: Y-zeolite)

Oil bath; temperature: 140 8C; Multi-mode; power:


solvent: DMF; reaction time: 500 W; solvent: DMF;
012 h; yield: 68%
reaction time: 9 min;
yield: 100%

Solvent: water and ethanol;


reaction time: 24 h;
yield: 20%

Multi-mode; power:
650 W; in absence of
solvent; reaction time:
9 min; yield: 75%

Temperature: 80 8C;
reaction time: 60 min;
conversion: 62%

Multi-mode; reaction time:


60 min; conversion: 100%

Multi-mode; solvent:
ethylene glycol;
reaction time: 2 min;
yield: 90%

Temperature: 140 8C;


reaction time: 5 h;
conversion (2): 50%;
conversion (3): 22%;
conversion (4): 4.5%

Multi-mode; power:
500 W; reaction time:
3 min; conversion (2):
89%; conversion (3):
6%; conversion (4): 5%

16

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Table 5 (Continued )
Reference Reaction

Conventional conditions

Microwave conditions

[91]

Deprotection of benzyl esters:

Oil bath; temperature:


135 8C; solvent:
dichloromethane; reaction
time: 20 min; yield: 92%

Multi-mode; temperature:
130 ! 140 8C; solvent:
dichloromethane; reaction
time: 7 min; yield: 92%

[92]

Reduction of SO2 and NOx:

Multi-mode; SO2 removal:


99%; NO removal: 99%

[93]

Oligomerization of methane (proposed mechanism)


CH4 ! :CH2 + H2; :CH2 ! :CH + :H; 2CH4 ! C2H4 + 2H2;
2CH4 ! C2H4 + 3H2; C2H4 + H2 ! C2H6; 3C2H2 ! C6H6
(catalyst: RaNi)

[94]

Esterification of propionic acid with ethanol:

Single-mode (continuous
flow operating quartz
reactor); power: 400 W;
reaction time: 10 min;
conversion: 24%
Single-mode; temperature:
105 8C (500 W); reaction
time: 13 h; Concentration
(ethyl propion.): 0.63 mol/L;
Keq. = 4.43

(catalyst: ion-exchange resin (Amberlyst 15))


[95]

Selective oxidation of alkylarenes:

[96]

(catalyst: KMnO4/montmorillonite K10)


Oxidation of alcohols:

[67]

(catalyst: clayfen [K10/Fe(NO3)3])


Esterification reaction:

[97]

(catalyst: Fe2(SO4)3 with montmorillonite)


Condensation of 1,4-diacetylpiperazine-2,5-dione with aldehydes:

(catalyst: KF/Al2O3)

Conventional heating;
temperature: 105 8C;
reaction time: 13 h;
Concentration
(ethyl propion.):
0.64 mol/L; Keq. = 4.9
Conventional conditions:
Not specified; solvent:
stirring; temperature: 25 8C;
in absence reaction time:
reaction time: 20 h; yield:
10 min; yield: 87%
90%; sonic conditions: reaction
time: 2 h; yield: 86%

Oil bath; temperature:


65 8C; reaction time:
18 h; yield: 100%

Multi-mode; in absence
of solvent; reaction time:
30 s; yield: 87%

Temperature: 140 8C;


reaction time:
2 h; yield: 83%

Single-mode; temperature:
140 8C; reaction time:
2 h; yield: 97%

Conventional condition;
temperature: 25 8C;
solvent: DMF; reaction
time: 16 h; yield: 90%

Multi-mode; power:
240 W; in absence of
solvent; reaction time:
15 min; yield: 94%

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

(1 g). This reaction mixture was irradiated for 2 min, at low


power setting, and then worked up after removing the
catalyst by filtration. The filtrate was diluted with water and
extracted with ethyl acetate. Evaporation of the solvent from
organic layer gave about 90% yield of the product.
Ipaktschi and Bru ck [90] carried out citronellal dehydrogenation over Y-zeolite under microwave irradiation
(microwave oven, 500 W) and conventional heating. The
reaction rate and selectivity were dramatically enhanced
under microwave exposure, compared to the reaction under
conventional thermal conditions. Several other Claisen
rearrangements under microwaves were studied, as well.
Varma et al. [91] studied deprotection of benzyl benzoate
over acidic alumina. The catalyst was added to a solution of
benzyl ester, dissolved in a minimum amount of dichloromethane (12 mL), at room temperature, and the reaction
mixture was thoroughly mixed using a vortex mixer. The
adsorbed material was dried in air and placed in an alumina
bath inside a microwave oven. Pure benzoic acid was obtained
(yield = 92%). The observed yield was three times higher
under milder reaction conditions by using microwave irradiation, compared to the conventional heating in an oil bath.
Tse et al. [92] investigated decomposition of SO2 and NOx
over several nickel and copper catalysts under microwave
irradiation. For the majority of the gas phase experiments
either 5% SO2 in air or 25% NO in an inert gas was passed
through a packed bed of commercially available supported
catalyst pellets while the system was exposed to microwave
irradiation. For SO2, the most efficient catalyst was Ni-1404;
more than 99% of the SO2 was removed from the effluent gas
after reaction over this catalysts over prolonged periods of
use. Experiments with 25% NO in He over several catalysts
indicated efficient decomposition under microwave exposure.
Interestingly, some copper containing catalysts, which had
shown activity in the SO2 experiments removed essentially
none of the NO, while Ni catalyst afforded 99% of NO
removal. These results indicate potential of the microwave
induced catalytic technology to the problems of removal of
high concentrations of undesirable pollutant gases.
Conde et al. [93] studied the oligomerization of methane
under microwave heating over Raney nickel catalyst. The
maximum conversion achieved in this reaction was 24%, at
400 W and 10 min of irradiation. Oligomers, such as
ethylene, benzene and ethane have been prepared selectively
under different conditions. Ethylene was the major product
and observed selectivity for ethylene was 70% at 530 W and
5 min, whereas benzene formation was favored (13%) when
an intermediate power (400 W) was used. Pretreatment of
the catalyst with hydrogen reduction increases the formation
of ethane (54%), at low and intermediate power.
Toukoniitty et al. [94] investigated esterification reaction
of propionic acid with ethyl alcohol over an ion-exchange
resin (Amberlyst 15) as a catalyst. The experiments were
carried out under microwave irradiation and conventional
heating in microwave loop reactor (Fig. 16). No significant
acceleration of reaction rate, neither shift in equilibrium

17

Fig. 12. SEM image of spent ion-exchange resin catalyst (Amberlyst 15)
subjected to conventional heating.

Fig. 13. SEM image of spent an ion-exchange resin catalyst (Amberlyst 15)
subjected to microwave irradiation.

concentration was observed under microwaves compared to


conventional heating. The fact that the equilibrium remained
unaffected was, nevertheless, expected since the laws of
thermodynamics should prevail. Spent catalysts (Amberlyst
15) were characterized by scanning electron microscopy
(SEM). Slight melting of catalyst particle occurred under
microwave irradiation compared to conventional heating
(Figs. 12 and 13). Cracking of the catalyst surface was
observed under microwave irradiation, as well, which could
not be obtained under conventional heating (Figs. 14 and
15).

Fig. 14. SEM image of spent ion-exchange resin catalyst (Amberlyst 15)
subjected to conventional heating.

18

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Fig. 15. SEM image of spent ion-exchange resin catalyst (Amberlyst 15)
subjected to microwave heating.

Shaabani et al. [95] carried out selective oxidations of


alkylarenes over KMnO4/montmorillonite K10, under sonic
and microwave irradiation. As can be seen from the results
presented in Table 5, the utilization of ultrasound in this
reaction decreases the time required to obtain good yields by
a factor of about 10. Moreover, the use of microwave
irradiation decreases the time required by another order of
magnitude.
Varma and Dahiya [96] studied the oxidation of alcohols in
solvent free conditions using clayfen (montmorillonite K10
clay-supported iron(III) nitrate) under microwave exposure.
Clayfen-mediated solvent free microwave thermolysis was
found to be a convenient, selective and environmentally
friendly process in comparison to the conventional solution
phase or heterogeneously catalyzed reaction pathways.
Chemat et al. [67] carried out esterification reaction
between stearic acid and n-butanol over Fe2(SO4)3
montmorillonite under microwave and conventional heating
conditions. The yield and the rate of heterogeneously
catalyzed esterification increased with microwave heating,
in comparison to the conventional heating under similar
reaction conditions (temperature, concentration, pressure).
Villemin and Alloum [97] investigated dry condensation
of 1,4-diacetylpiperazine-2,5-dione with aldehyde over KF/
Al2O3 under microwave exposure. 1,4-Diacetylpiperazine2,5-dione and benzaldehyde were absorbed on the solid base
KF/Al2O3 and irradiated for 15 min. The reaction took place
at room temperature as well, when 1,4-diacetylpiperazine2,5-dione and benzaldehyde were dissolved in DMF, reacted
with KF/Al2O3. In this case, the obtained yield was poorer
than in dry condensation under microwaves.
As a summary, significant acceleration of a wide range of
chemical reactions under microwave irradiation in comparison to the conventional heating has been reported in this
review. However, enhancement in several cases can be due to
higher reaction temperature in the dielectric field, although
main part of the reported studies were carried out in
microwave reactors equipped with efficient temperature
control, which enables exact temperature monitoring during
the reaction. Feasible explanation for observed increase of
activity and selectivity is selective heating of the catalyst

particles leading to the creation of hot-spots, moreover, when


the reaction mechanism occurs via the more polar transition
state, dielectric heating can be favored. Rapid superheating of
the solvents improves activity and selectivity as well. A very
important advantage of microwave irradiation is the
possibility to carry out many chemical syntheses rapidly
and with good yields in solutions as well as in the absence of a
solvent. This leads to enhancement of selectivity and gives
raise to the inherently greener chemical production.
Microwave irradiation can also be utilized in catalyst
preparation in order to enhance crystallization process,
improve catalyst activity, significantly shorten reaction time
and enable dry synthesis to achieve very high batch yield
[98]. However, conventional thermal methods are not
replaced by microwave heating and might in some cases
be preferable. Most of the attention was paid to the field of
zeolite synthesis, the comprehensive review with detailed
description of the area was published by Cundy [99].
Prasad et al. [100] prepared alumina- and silica-supported
palladium catalyst by conventional and microwave heating.
The catalysts were characterized and tested for benzene
hydrogenation activity. Catalysts containing 10 wt.% Pd were
prepared by impregnating commercial alumina and silica with
aqueous solutions of palladium nitrate. The catalysts were
evaporated to near dryness and dried in oven at 120 8C for
12 h. A portion of catalyst was subsequently calcinated in air
for 5 h and the remaining portion was irradiated in microwave
oven for 5 min. In the case of Pd/Al2O3 catalyst, microwave
heating degreases the total surface area (MH = 117 m2/g,
CH = 152 m2/g), but a slight increase of total surface area by
microwave heating was observed with Pd/SiO2 catalyst
(MH = 499 m2/g, CH = 462 m2/g). The particle size was
significantly increased by microwave heating (Pd/Al2O3:
, CH = 47 A
and Pd/SiO2: MH = 160 A
,
MH = 229 A

CH = 135 A). The preparation of samples by microwave


heating leads to enhanced crystallite size and increased
hydrogenation activity.
Bond et al. [101] studied Ni/Al2O3 catalyst, utilizing
microwave heating for catalyst drying. Ni/Al2O3 was
prepared by impregnation to obtain catalysts with 5 and
10% loading of the active metal. The catalysts precursors
were prepared by wetting the pellets in aqueous solutions of
nickel nitrate of the required concentrations to provide
desired metal loading. The catalysts were dried by
microwave or conventional heating. Drying times of catalyst
precursors were reduced by using microwave irradiation
compared to conventional heating. In the case of 10% Ni/
Al2O3 catalyst the pellet strength was investigated. The
results confirmed that the microwave dried samples are
significantly stronger than those dried conventionally. The
microwave dried catalyst also showed much more uniform
metal distribution on the support.
Synthesis of TS-1 zeolite under microwave irradiation
has been studied by Uguina et al. [102]. In this work, TS-1
zeolite was synthesized by utilizing both conventional and
microwave heating. The use of microwave irradiation allows

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

the synthesis of highly crystalline TS-1 zeolite in a very


short time (2030 min), compared to conventional heating
(around 15 h). Synthesis of TS-1 zeolite under microwave
irradiation leads to larger average crystal sizes than under
conventional heating, but they posses more uniform shapes
and sizes.
Somani et al. [103] synthesized ZSM-5 zeolite using two
different techniques: first by conventional hydrothermal
method and second by exposing the hydrogel to the
microwave irradiation followed by conventional method.
The crystallinity of ZSM-5 was tested by using X-ray
diffractometer. Zeolite synthesized under microwave irradiation took half time (MH = 18 h) to crystallization of
100% crystalinity compared to conventional method
(CH = 36 h). The scanning electron micrograph showed
that size of the crystals prepared under microwave
irradiation were larger (MH = 56 mm) compared to
conventional method (CH = 34 mm).
2.9. Microwave equipment
Microwave equipment can be divided into two categories: (a) multi-mode and (b) single-mode cavities.
Considering an empty metallic volume, the electric field
repartition into that volume is very heterogeneous, if the
dimensions of that volume are too big compared to the
wavelength. This is the case for multi-mode applicators
(such as a domestic microwave oven). The repartition is well
mastered and stable if the applicator dimensions are close to
the single-mode structure. The use of wave-guides emitting
the fundamental mode at fixed frequency, allows to master
and, above all control, the power transmission when the aim
is to study the influence of microwave electromagnetic field

19

on the behavior of chemical reactions. This approach makes


it feasible to extrapolate the results to industrial scale.
2.9.1. Multi-mode cavities
Domestic microwave oven is a typical example of a
multi-mode device. Such devices comprise a large microwave cavity with reflective walls that are necessary to
prevent leakage of radiation and to increase the efficiency of
the oven. It is obvious that domestic microwave ovens are
not intended for chemical applications. For instance, they do
not allow application of high pressures since explosions can
occur [104]. Moreover, they are not compatible with
corrosive and inflammable compounds; this is being another
disadvantage of them. Domestic microwave ovens operate
on duty cycles, with the intervals between zero and full
power reaching several seconds (e.g. 600 W oven 30 s duty
cycle). Large duty cycles are undesirable in chemical
applications since samples may cool dramatically between
switching steps [61]. However, self-modified microwave
ovens were utilized in chemical applications already a
decade ago since commercially produced equipment was not
easily available.
2.9.2. Single-mode cavities
Single-mode units have a smaller cavity. These types of
equipment have greater energy efficiency and more even
temperature distribution in the reaction vessel than multimode units. Single-mode resonant heater allows a sample to
be placed at a position of much higher electric field strength
than can be obtained in a multi-mode oven [61].
Single-mode equipment (Figs. 16 and 17) needs to have a
built-in protection of the magnetron and working with small
samples does not, therefore, cause problems [105]. Single-

Fig. 16. Single-mode MW loop reactor.

20

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

Fig. 17. Side-view of the microwave cavity (heating zone) and the surrounding appliances, iris and short-circuit (a reflector).

mode equipment has a variable microwave power input,


which means that the resulting temperature can be measured
accurately and the pre-set temperature can be achieved. Such
approach requires a recording unit to read the signal from the
temperature measuring device. It should be noted that in
fact, temperature measuring within microwave field is quite
problematic and hence the temperature has to be measured
by special technique, like infrared pyrometry or fiber-optic
devices. The level of power emitted by generator coupled to
a load, must be determined from simple thermodynamic
dependences as follows [106]: suppose that for a certain
mass, m (kg), of material, with specific heat capacity, cp (kJ/
kg K), an increase of temperature DT is desired within a
certain time interval t (s). The necessary theoretical power is
obtained from the following equation:
Ptheor

mcp DT
t

(7)

where Ptheor denotes the theoretical power. The material to


be treated must be located into the vessel within the cavity.
The vessel must be microwave transparent (Teflon, quartz)
chemically inert and tolerate temperatures reached. Moreover, high pressures operating conditions induced by microwave irradiation are very attractive for some chemical
applications, therefore the reaction vessel should be able
to accommodate high pressures. For temperatures up to
180 8C, Teflon is a convenient choice. However, if higher
temperatures are required, a quartz vessel is preferred.
The single-mode microwave generator (Fig. 17) consists
of a boxed high voltage power supply and a control unit with
a separate microwave head. Different kinds of generators,
with various nominal effects and frequencies, i.e. 2450 and
915 MHz, can be used.
Although microwave radiation is electromagnetic but not
radioactive radiation, the harmfulness of radio frequency to
living organism needs to be accounted for proper safety

measures. First of all, good connections in between all flanges


and wave-guides are essential to avoid any leakage of
electromagnetic irradiation. It is also important to ensure that
cavity always contains matter; i.e. in case of loop configuration (Fig. 16) the control of flow is crucial since otherwise the
microwave reflected power sensor might be burnt due to the
excessive reflected power or meltdown or explosion might
happen in the cavity due to the overheating. Therefore, it is
highly recommended to install an emergency stop switch that
enables the operator to switch off the microwave radiation
source rapidly. Also, a pre-set flow rate alarm should shut off
the microwaves in case the flow rate falls below the set-point.
In addition, the safety control can easily be built in case the
real temperature in the cavity is measured by means of optic
fibers, etc., with a high temperature limit alarm/switch
controlling the microwave source.
Single-mode devices work well for either high-loss or
moderately high-loss type of dielectric materials. For
materials that are very microwave-transparent, such as
hydrocarbons, the applicator must be resounded, while using
a variable sliding short-circuit position and iris tuning. Multimode techniques are coupling with very high-loss materials
which means that if the loss is moderate or low, multi-mode
methods are less effective and single-mode methods are
preferred [107]. This method allows the microwave heating
of relatively low-loss materials such as polymers or some
ceramics, using low powers (<2.5 kW) [61].
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. Several
examples and detail description of commercial laboratory
microwave reactors are reported in [108].
However, generally the commercial equipment utilizes
magnetically stirred vessels which is not always sufficient;
in case of many reactions, especially heterogeneously

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

catalyzed ones, the mixing power is insufficient for removal


of mass transfer limitations. Therefore, mechanical mixing
should be incorporated in such devices.

3. Conclusion
The last two decades brought along the renaissance of
sonochemical research so that ultrasound has become more
and more common laboratory tool to enhance chemical
synthesis. The use of ultrasound in chemical systems
increases rates, improves yields and selectivities, prevents
catalyst deactivation and improves the heat and mass
transfer. It is very probable that ultrasound will become an
important tool in industrial applications for enhancement of
chemical reactions in the near future.
Microwave irradiation is becoming an interesting method
to heat materials in chemical laboratories, offering a clean,
cheap, convenient, selective and instantaneous method of
heating. Microwave heating improves yields and selectivities, increases rates and allows working in solvent free
conditions. Many reactions that do not occur under classical
heating can be carried out with high yields under MW
irradiation. Since 1986, increasing amount of articles
concerning microwaves was published. However, it can
be stated, that, at present, moment microwave irradiation in
catalytic chemical reactions still remains largely
unexploited and all the potential of microwaves in chemical
catalytic applications has not yet being discovered.
Both of these methods offer new avenues for more energyefficient and environmentally benign processes of 21st
century. The future is always hard to foresee, but we have
certainly not yet seen all the potential of innovative, novel
techniques for the chemical industry that reshapes its image
and efficiency in the spirit of elegance and greener values.

References
[1] K.S. Suslick, Sonocatalysis, Handbook of Heterogeneous Catalysis,
VCH Verlagsgesellschaft mbH, Weinheim, Germany, 1997, p. 1350.
[2] L.H. Thompson, L.K. Doraiswamy, Ind. Chem. Res. 38 (1999) 1215.
[3] W.T. Richards, A.L. Loomis, J. Am. Chem. Soc. 49 (1927) 3086.
[4] E.A. Neppiras, Phys. Rep. 61 (1980) 159.
[5] K.S. Suslick, Sci. Am. (1989) 80.
[6] T.J. Mason, Ultrason. Sonochem. 7 (2000) 145.
[7] M.A. Margulis, Ultrasonics 23 (1985) 157.
[8] F. Lepoint-Mullie, D. De Pauw, T. Lepoint, Ultrason. Sonochem. 3
(1996) 73.
[9] A. Henglein, C. Kormann, Int. J. Radiat. Biol. 48 (1985) 251.
[10] P. Cintas, J.L. Luche, J. Green Chem. (1999) 115.
[11] T. Linker, J. Organomet. Chem. 661 (2002) 159.
[12] C. Guillard, P. The ron, P. Pichat, C. Pe trier, Water Res. 36 (2002)
4263.
[13] A.B. Pandit, P.R. Gogate, S. Mujumdar, Ultrason. Sonochem. 8
(2001) 227.
[14] I.Z. Shirgaonkar, A.B. Pandit, Ultraon. Sonochem. 5 (1998) 53.
[15] J. Berlan, F. Trabelsi, H. Delmas, A.M. Wilhelm, J.F. Petrignani,
Ultrason. Sonochem. 1 (1994) 97.

21

[16] N. Serpone, R. Terzian, J. Phys. Chem. 98 (1994) 2634.


[17] P. Nebois, Z. Bouaziz, H. Fillion, L. Moeini, M.J. Aurell Piquer, J.L.
Luche, A. Rieara, A. Moyano, M.A. Pericas, Ultrason. Sonochem. 3
(1996) 7.
[18] N. Kardos, J.L. Luche, Carbohydr. Res. 332 (2001) 115.
[19] A.L. Crum, Ultrason. Sonochem. 2 (1995) 147.
[20] T.J. Mason, J.P. Lorimer, Applied Sonochemistry, WileyVCH
Verlag GmbH, Weinheim, 2002.
[21] H. Frenzel, H. Schultes, J. Phys. Chem. 27b (1934) 421.
[22] K.S. Suslick, Y. Didenko, M.M. Fang, T. Hyeon, K.J. Kolbeck, W.B.
McNamara III, M.M. Mdleleni, M. Wong, Phil. Trans. R. Soc. Lond.
A 357 (1999) 335.
[23] M.A. Margulis, I.M. Margulis, Ultrason. Sonochem. 9 (2002) 1.
[24] A.J. Walton, G. Reynolds, Adv. Phys. 33 (1984) 595.
[25] K.S. Suslick, S.J. Doktycz, S.J. Flint, Ultrasonics 28 (1990) 280.
[26] J.P. Lorimer, T.J. Mason, Chem. Soc. Rev. 16 (1987) 239.
[27] M. Gutierrez, A. Henglein, J. Phys. Chem. 94 (1990) 3625.
[28] N. Ratoarinoro, A.M. Wilhelm, J. Berlan, H. Delmas, Chem. Eng. J.
50 (1992) 27.
[29] G.O.H. Whillock, B.F. Harvey, Ultrason. Sonochem. 4 (1997) 33.
[30] A. Henglein, M. Gutierrez, J. Phys. Chem. 94 (1990) 5169.
[31] F.R. Young, J. Acoust. Soc. Am. 60 (1976) 100.
[32] A. Gedanken, Chem. Israel 9 (2002) 13.
[33] K.J. Moulton, S. Koritala, E.N. Frankel, J. Am. Oil Chem. Soc. 60
(1983) 1257.
[34] B. To ro k, G. Szo llo si, K. Bala zsik, K. Felfo ldi, I. Kun, M. Barto k,
Ultrason. Sonochem. 6 (1999) 97.
[35] J. Lindley, J.P. Lorimer, T.J. Mason, Ultrasonics 25 (1987) 4548.
[36] J.-T. Li, G.-F. Chen, W.-Z. Yang, T.-S. Li, Ultrason. Sonochem. 10
(2003) 123.
[37] D.H. Shin, B.H. Han, Bull. Korean Chem. Soc. 6 (1985) 247.
[38] B. To ro k, K. Felfo di, G. Szakonyi, M. Barto k, Ultrason. Sonochem. 4
(1997) 301.
[39] P.W. Cains, L.J. McCausland, D.M. Bates, T.J. Mason, Ultrason.
Sonochem. 1 (1994) 45.
[40] M.G. Sulman, Russ. Chem. Rev. 69 (2000) 165.
[41] T. Ando, S. Sumi, T. Kawate, J. Ichihara, H. Terukiyp, J. Chem. Soc.,
Chem. Commun. 7 (1984) 439.
[42] B.H. Han, P. Boudjouk, J. Org. Chem. 47 (1982) 5030.
[43] A. Fuentes, J.M. Marinas, J.V. Sinisterra, Tetrahedron Lett. 28 (1987)
4541.
[44] P. Boudjouk, B.-H. Han, J. Catal. 79 (1983) 489.
[45] G. Szo llo si, B. To ro k, G. Szakonyi, I. Kun, M. Barto k, Appl. Catal. A
172 (1998) 225.
[46] B.-H. Han, P. Boudjouk, Organometallics 2 (1983) 769.
[47] J. Gonzalo Rodrguez, A. Lafuente, Tetrahedron Lett. 43 (2002) 9581.
[48] G.D. Yadav, M.S.M. Mujeebur Rahuman, Ultrason. Sonochem. 10
(2003) 135.
[49] J.-P. Mikkola, B. Toukoniitty, E. Toukoniitty, J. Aumo, T. Salmi,
Ultrason. Sonochem. 11 (2004) 233.
[50] B. Toukoniitty, E. Toukoniitty, P. Ma ki-Arvela, J.-P. Mikkola, T.
Salmi, D.Yu. Murzin, P.J. Kooyman, Ultrason. Sonochem., in press.
[51] J. Lindley, Ultrasonics 30 (1992) 163.
[52] C.L. Bianchi, F. Martini, V. Ragaini, Ultrason. Sonochem. 8 (2001)
131.
[53] J.-L. Li, T. Inui, Appl. Catal. A 139 (1996) 87.
[54] U.R. Pillai, E. Sahle-Demessie, R.S. Varma, Appl. Catal. A 252
(2003) 1.
[55] G.J. Price, D.J. Norris, P.J. West, Macromolecules 25 (1992) 6447.
[56] Y. Hatate, T. Ikeura, M. Shinonome, K. Kondo, F. Nakashio, J. Chem.
Eng. Jpn. 14 (1981) 38.
[57] G.J. Price, Ultrason. Sonochem. 10 (2003) 277.
[58] K.S. Suslick, M. Fang, T. Hyeon, Sonochemistry and Sonoluminescence, Kluwer Academic Publishers, Dordrecht, The Netherlands,
1999, p. 291.
[59] M.H. Entezerai, P. Kruus, R. Otson, Ultrason. Sonochem. 4 (1997)
49.

22

B. Toukoniitty et al. / Applied Catalysis A: General 279 (2005) 122

[60] K.J. Laidler, J.H. Meiser, Physical Chemistry, Houghton Mifflin,


1995.
[61] D.M.P. Mingos, D.R. Baghurst, Applications of Microwave Dielectric Heating Effects to Synthetic Problems in Chemistry, MicrowaveEnhanced Chemistry, American Chemical Society, Washington, DC,
USA, 1997.
[62] S.L. Percy, US Patent 2,473,550 (1949).
[63] R. Gedye, F. Smith, K. Westaway, H. Ali, L. Baldisera, L. Laberge, J.
Rousell, Tetrahedron Lett. 27 (1986) 279.
[64] G. Roussy, J.A. Pearce, Foundations and Industrial Applications of
Microwave and Radio Frequency Fields, Wiley, Chichester, UK,
1995.
[65] D.L. Rakhmankulov, S.Yu. Shavshukova, F.N. Latypova, V.V. Zorin,
Russ. J. Appl. Chem. 75 (2002) 1377.
[66] R.B. Moyes, G. Bond, Handbook of Heterogeneous Catalysis, VCH
Verlagsgesellschaft, Weinheim, Germany, 1997.
[67] F. Chemat, D.C. Esveld, M. Poux, J.L. DiMartino, J. Microwave
Power Electromagn. Energy 33 (1998) 88.
[68] S.L. Suib, CatTech 2 (1998) 75.
[69] K.D. Raner, C.R. Strauss, J. Org. Chem. 57 (1992) 6231.
[70] D.R. Baghurst, D.M.P. Mingos, J. Chem. Soc., Chem. Commun. 9
(1992) 674.
[71] A. Loupy, A. Petit, J. Hamelin, F. Rexier-Boullet, P. Jacquault, D.
Mathe , Synthesis (1998) 1213.
[72] L. Perreux, A. Loupy, Nonthermal Effects of Microwaves in Organic
Synthesis, Microwaves in Organic Synthesis, WileyVCH Verlag
GmbH, Weinheim, 2002.
[73] R.A. Abramowich, Org. Prep. Proc. Int. 23 (1991) 683.
[74] F. Langa, P. De La Cruz, A. De La Hoz, A. Diaz-ortiz, E. Diez-Barra,
Contemp. Org. Synth. 4 (5) (1997) 373.
[75] I. Almena, A. Diaz-Ortiz, E. Diez-Barra, A. De La Hoz, A. Loupy,
Chem. Lett. 5 (1996) 333.
[76] B. Bennett, Sci. Environ., ESOS 79 (1994) 29.
[77] A. Miklavc, Chem. Phys. Chem. 2 (2001) 552.
[78] D.A.C. Stuerga, P. Gaillard, J. Microwave Power Electromag. Energy
31 (1996) 87.
[79] W.-C. Sun, P.M. Guy, J.H. Jahngen, E.F. Rossomando, E.G.E.
Jahngen, J. Org. Chem. 53 (1988) 4414.
[80] E.G.E. Jahngen, R.R. Lentz, P.S. Pesheck, P.H. Sackett, J. Org.
Chem. 50 (1990) 3406.
[81] K.D. Raner, C.R. Strauss, J. Org. Chem. 57 (1992) 6231.
[82] R.N. Gedye, F.E. Smith, K.C. Westaway, Can. J. Chem. 66 (1988)
17.
[83] R.N. Gedye, W. Rank, K.C. Westaway, Can. J. Chem. 69 (1991)
706.

[84] U.R. Pillai, E. Sahle-Demessie, R.S. Varma, Tetrahedron Lett. 43


(2002) 2909.
[85] M. Ha jek, H. Richterova , J. Stach, in: Proceedings of the 22nd
Conference of Organic Chemists on Advances in Organic Chemistry,
Slovakia, June, 1997.
[86] R.S. Huber, G.B. Jones, J. Org. Chem. 57 (1992) 5778.
[87] D. Villemin, B. Labiad, Synth. Commun. 22 (1992) 2043.
[88] I. Calinescu, M. Radoiu, R. Calinescu, D. Martin, in: Proceedings of
the 5th International Electronic Conference on Synthetic Organic
Chemistry (ECSOC-5), September, 2001.
[89] B.K. Banik, M.S. Manhas, Z. Kaluza, K.J. Barakat, A.J. Bose,
Tetrahedron Lett. 33 (1992) 3603.
[90] J. Ipaktschi, M. Bru ck, Chem. Ber. 123 (1990) 1591.
[91] R.S. Varma, A.K. Chatterjee, M. Varma, Tetrahedron Lett. 34 (1993)
4603.
[92] M.Y. Tse, M.C. Depew, J.K.S. Wan, Res. Chem. Inermed. 13 (1990)
221.
[93] L.D. Conde, C. Maru n, S.L. Suib, J. Catal. 218 (2003) 201.
[94] B. Toukoniitty, J.-P. Mikkola, K. Era nen, T. Salmi, D. Yu. Murzin,
Catal. Today, in press.
[95] A. Shaabani, A. Bazgir, F. Teimouri, D.G. Lee, Tetrahedron Lett. 43
(2002) 5165.
[96] R.S. Varma, R. Dahiya, Tetrahedron Lett. 38 (1997) 2043.
[97] D. Villemin, A.B. Alloum, Synth. Commun. 20 (1991) 3325.
[98] Z.J. Ping, C.S. Cundy, J. Dwyer, in: Proceedings of the 12th International Zeolite Conference, 1999, pp. , p. 1591.
[99] C.S. Cundy, Collect. Czech. Chem. Commun. 63 (1998) 1699.
[100] P.S.S. Prasad, N. Lingaiah, P.K. Rao, F.J. Berry, L.E. Smart, Catal.
Lett. 35 (1995) 345.
[101] G. Bond, R.B. Moyes, S.D. Pollington, D.A. Whan, Stud. Surf. Sci.
Catal. 75 (1993) 1805.
[102] M.A. Uguina, D.P. Serrano, R. Sanz, E. Castillo, in: Proceedings of
the12th International Zeolite Conference, 1999, pp. , p. 1917.
[103] O.G. Somani, S.L. Choudhari, B.S. Rao, S.P. Mirajkar, Mater. Chem.
Phys. 82 (2003) 538.
[104] L. Gilman, W. Grooms, Anal. Chem. 60 (1988) 1624.
[105] G. Bond, R.B. Moyes, D.A. Whan, Catal. Today 17 (1993) 427.
[106] P.J. Walter, S. Chalk, H.M. Kingston, Overview of MicrowaveAssisted Sample Preparation, Microwave-Enhanced Chemistry,
American Chemical Society, Washington, DC, USA, 1997.
[107] K. Takatsu, F. Kurogi, A. Matsue, M. Kasaya, Soc. Automot. Eng.
960344 (1996) 69.
[108] D. Stuerga, M. Delmotte, WaveMaterial Interactions, Microwave
Technology and Equipment, Microwaves in Organic Synthesis,
WileyVCH Verlag GmbH, Weinheim, 2002.

You might also like