Sie sind auf Seite 1von 19

ISSN 0036 0244, Russian Journal of Physical Chemistry A, 2010, Vol. 84, No. 10, pp. 16761694.

Pleiades Publishing, Ltd., 2010. Original Russian Text L.M. Kustov, I.M. Sinev, 2010, published in Zhurnal Fizicheskoi Khimii, 2010, Vol. 84, No. 10, pp. 18351856.

CHEMICAL KINETICS AND CATALYSIS

Microwave Activation of Catalysts and Catalytic Processes


L. M. Kustov and I. M. Sinev
Zelinskii Institute of Organic Chemistry, Russian Academy of Sciences, Leninskii pr. 47, Moscow, 119991 Russia e mail: lmk@ioc.ac.ru
Received April 2, 2010

AbstractThe results obtained by using microwave activation in catalytic processes such as the conversion of lower alkanes, the removal of volatile organic compounds and sulfur and nitrogen containing compounds from air, and the hydrogenation of aromatic compounds and dehydrogenation of naphthenes are discusses. DOI: 10.1134/S0036024410100055

INTRODUCTION Various radiation types (gamma irradiation, elec tron and neutron beams, RF/microwave or sound range waves, and irradiation by UV and visible light) are extensively used in material preparation and for the initiation and stimulation of chemical processes. In recent years, much attention has been given to the action of microwave radiation on matter, in particular, to the influence of microwave heating on the prepara tion of various materials and catalytic reactions. Microwaves, or hyperhigh frequency electromag netic waves, are electromagnetic waves 1 mm1 m long. Their frequency range is 200 MHz300 GHz, and they are intermediate in the spectrum of electro magnetic radiation between radiofrequencies and infrared radiation. Waves 125 cm long are extensively used in radar applications, and other waves, in long haul communications. To avoid interferences to civil communications and military and naval hardware, a set of frequencies was fixed for use in industrial, research, and medical equipment [1]. The 915 MHz frequency (33.3 cm) is used in industrial devices, 2.45 GHz (12.2 cm) is used in microwaves, and, in addition, devices working at 0.433, 5.8, and 24.125 GHz are allowed in certain countries. The basic elements of microwave devices are mag netron generators invented during World War II by Randal et al. at Birmingham University for use in radi olocation. The first commercial microwave furnaces were produced by the Raytheon corporation (USA) and for the first time appeared at USA markets in the 1950s. At world market, they appeared in the 1970s 1980s simultaneously with the development of the mass production of magnetrons. The key advantage of energy transfer with micro wave fields over convective methods for heating is the transfer of energy through radiation rather than heat transfer or convection. This ensures fast penetration of energy into the volume of materials transparent to microwave fields, that is, almost instantaneous heating

(and cooling when the field is switched off) of sub stances. Ideally, the transformation of electromagnetic field energy into heat occurs simultaneously and equally over the whole object volume absorbing micro wave radiation, which results in high heating rates. The spatial temperature distribution in the object is then different from that observed under traditional convective or contact heating conditions. The most important differences related to the appearance of temperature gradients and nonequilibrium conditions are observed when a reaction medium or material (for instance, a catalyst) consists of several phases with dif ferent abilities to be heated by microwave radiation. Examples of using microwave radiation for the exe cution of chemical processes are given in [2]. One of the spheres of industrial applications of microwave fields is microwave drying technologies. This problem was for the first time studied by Levinson [3], who used a carbonaceous material as a receiver of microwaves for heating materials to be dried to the required tem perature. The existing examples are largely related to organic synthesis under the conditions of microwave irradia tion. The first research works on the use of microwave fields in organic synthesis were published in 1986 [4, 5]. The purpose of these works was to decrease reac tion duration, increase yields and selectivity, and decrease energy and reagent consumption. Starting with the beginning of the 1980s, interest in the use of microwave fields as an energy source for the activation of chemical reactions or separation of sub stances has been increasing. The first patent dates to 1982; it describes gas phase destruction of chlorine containing organic compounds on heterogeneous cat alysts (paramagnetic and ferromagnetic powders) heated by a microwave field at a 2.45 GHz frequency to reaction temperature during a short time [6]. Another patent dated to 1985 describes the catalytic transformation of methane into ethylene on nickel and iron catalysts activated by microwave field pulses

1676

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1677

about milliseconds wide [7]. According to [8], the use of microwave energy pulses provides additional energy saving. The number of the corresponding publications substantially increased starting with 1995. Although the first objects of study were deposited metallic cata lysts, currently, the efforts of researchers are largely focused on the use of perovskite like mixed oxides because of their high catalytic activity and high ability to absorb electromagnetic radiation energy. With few exceptions, so called single mode generators working at a 2.45 GHz frequency are used for heating catalysts. Multi mode systems, in which a larger number of dif ferent modes are generated, were largely used for the preparation of catalysts. As applied to catalysis, microwave technologies can be used for both catalyst preparation and prelimi nary activation and direct execution of catalytic pro cesses. In this review, we concentrate on the use of microwave activation directly in catalytic processes. Pioneers in the use of microwave heating in heter ogeneous catalysis were Wan et al. In [9], they used ferro and paramagnetic metallic particles distributed in a liquid reaction medium, which was supposed to accelerate the reaction, because the majority of organic reagents and solvents did not absorb micro wave field to a substantial extent. Another advantage of this approach becomes clear when catalytically active components absorbing microwave radiation are placed into a carrier matrix transparent to microwave field. This approach is known as microwave catalysis. Wan as a rule worked with ferromagnetic and para magnetic metallic catalysts that well absorbed micro wave energy and had high catalytic activity. Wans team discovered that the use of pulsed microwave action allowed catalyst temperature and process selectivity to be controlled more accurately. At a high absorbing activity of catalysts, a short micro wave power pulse was sufficient for heating it to reac tion temperature. During dead time (cycle part without microwave action), heat released in the vol ume of a catalyst was indeed scattered in the matrix surrounding it and transparent to microwave field (reagents or carrier). Because of a lower temperature of the system as a whole, side reactions could be sup pressed. Wans team used a single mode microwave unit with a maximum output power of 3 kW. Apart from working magnetron power, dead time between pulses was regulated. Therefore, the selectivity of product formation and process temperature were controlled. These studies showed that metals set fire to micro wave plasmas at decreased temperatures and pres sures, which could successfully be used in catalytic reactions [10]. Short plasma discharges could, how ever, appear also when systems containing metals or pure oxide catalysts were used at reduced pressures,
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

which caused catalyst coking and, as a consequence, a decrease in catalyst activity in certain reactions. Apart from metallic powders, Wan et al. studied other commercially available materials containing nickel and copper, wires, nets, and disperse metallic catalysts distributed in nonconducting materials. They studied reactions of various types, including the hydrogenation and hydrocracking of hydrocarbons [11, 12], methane decomposition [13], the oxidation of hydrocarbons, the reduction of sulfur and nitrogen oxides [14], the synthesis of acetylene [1518], the catalytic decomposition of halogenated hydrocarbons [19], olefins [20], oil sand from Alberta province (Canada), and bitumens [21], and the synthesis of hydrogen cyanide [22]. According to Perry [23, 24] and Thomas [25, 26], substantial temperature gradients between catalyti cally and microwave active materials (for instance, platinum) and environment transparent to microwave fields could not be obtained. This was theoretically substantiated by Perry for a simplified model based on equilibrium heat exchange between a catalyst particle and gas. On the other hand, it is assumed that catalyt ically active system regions can be heated selectively when pulsed energy supply is used. At the same time, Mingos and Zhang [27] reported a temperature differ ence of ~100200 K between a catalytically active center and its nearest environment in the endothermic decomposition of H2S on MoS2/ Al2O3 at constant heating by a microwave field. Roussy et al. [28] used high dispersity platinum absorbing microwave radiation on Al2O3 transparent to microwave field for the isomerization of hexane and 2 methylpentane and hydrogenolysis of methylcyclo pentane. They also studied the isomerization of 2 methyl 2 pentene on pure oxide catalysts catalyzed by acids and the oxidative condensation of methane on transition metal oxides. They found that the use of microwave fields in situ could lead to the results iden tical to or quite different from those obtained using thermal activation. Apart from presenting their exper imental data, Roussy makes an important method ological conclusion. The temperature dependences of conversion extensively used in the literature con cerned with catalysis are not informative because of strong differences in the methods of heating and diffi culties in determining the true catalysis temperature under microwave field conditions. To obtain informa tion that can more easily be interpreted, it was sug gested to compare conversionselectivity depen dences rather than temperature curves. PHYSICAL BASICS OF MICROWAVE FIELD INTERACTION WITH SUBSTANCES In the gas phase at low pressures, the lifetimes of states obtained in the excitation of a particular rota tional mode are long, and microwave spectrum lines
Vol. 84 No. 10 2010

1678

KUSTOV, SINEV

are then fairly narrow. It follows that the spectrum of a free molecule in the microwave range is a set of narrow lines over the frequency range 360 GHz. At pressures higher than 101 torr, the frequency of collisions increases. Accordingly, the lifetime of an excited state decreases. As a consequence, according to the Heisen berg uncertainty principle, spectral lines broaden. The mean velocity (energy) of the motion of molecules and, therefore, gas temperature increase when rota tional excitation relaxes and its energy is transferred to translational degrees of freedom. This explains the mechanism of observed gas heating under the action of microwave radiation. In liquids and, especially, solids, substance mole cules cannot rotate independently. For this reason, other specific mechanisms of heating by microwave fields operate for substances in condensed states. One of the possible mechanisms is dielectric polarization, which is a relaxation process. Thermal energy release occurs in the relaxation of fixed charges (for instance, dipoles or charged point defects) from the polarized into stationary state. The second mechanism involves currents of free charges excited in solids and contrib uting to heating because of Ohmic loss; this mecha nism is characteristic of solids with substantial con ductivity (metals and semiconductors). The third mechanism that should also be taken into account is loss of vortex currents excited by magnetic fields. We can confidently rule out the possibility that microwave fields can initiate chemical reactions by exciting and, especially, breaking chemical bonds, as in photochemical processes under the action of elec tromagnetic radiation quanta at far shorter waves. The energy of microwave photons is ~1 J/mol, which is incomparable with activation energies of the majority of chemical processes. For instance, let us compare the energies of molecular motions and chemical bonds (I, Brownian movement; II, H bonds; III, covalent bonds; and IV, ionic bonds) with the energy of micro wave photons (V), kJ/mol,
I 1.64 II 3.842 III 480 IV 730 V 1 103

wave heating mechanisms. Microwave heating depends on the frequency and power of the field applied, but, since it is not based on quantum effects, it can be considered in terms of classical electrody namics. The theory of microwave heating was devel oped by several authors, among which Debye (1928), Cole and Cole (1941), and Hill (1962) should be men tioned. The rate of heating of liquids and solids under the action of the electric electromagnetic field component is determined by the equation [29]
2 T/ t = const '' fE ef / C p ,

(1)

where Eef is the acting electric field strength value, is the density of the substance, Cp is the isobaric heat capacity, '' is the dielectric loss coefficient, and f is the field frequency. Radiation loss is determined by the equation [30]

T = e A t C p V

()

T 4,
rev

(2)

THE EFFECTS AND MECHANISMS OF DIELECTRIC HEATING OF CONDENSED SUBSTANCES Substances in the condensed state can be heated by high frequency electromagnetic waves. The reason for heating is the ability of the electric electromagnetic field component to act on electric charges. If there are mobile charge carriers in substances (both metals and semiconductors), currents appear under the action of electric fields. If charge carriers are, however, fixed in certain regions, their movement is limited by counter forces, and dielectric polarization can appear. Both dielectric polarization and conductivity are micro

where is the emissivity of the body, is the Stefan Boltzmann constant, and /V is the ratio between the outer surface area and volume of the body. It follows that heating is determined by the dielectric loss, ther mal radiation of the body, and the strength of the elec tric component of the electromagnetic field. All these physical properties depend on temperature, which strongly complicates a complete theoretical analysis of dielectric heating. One of few examples of a detailed study of the dielectric properties of a substance [31] is a study of the temperature dependence of the dielectric constant and dielectric loss coefficient of CuO. It was found for CuO that the dielectric loss coefficient sharply increased at ~700C, which in turn caused a sharp increase in temperature. The data on heating of vari ous liquids and solids in household and special (for laboratory studies) microwave furnaces were reported in several works. Some results obtained for liquid and solid substances [32, 33] are summarized in Tables 1 and 2, respectively. It is easy to notice that nonpolar liquids (dimethyl ether, hydrocarbons, and CCl4) remain almost unheated by microwave fields under given conditions, which complicates the accomplishment of catalytic processes in these media (when these substances play the role of solvents or catalytic reaction substrates). Similarly, solids with a high conductivity (carbon and metals) are rapidly and effectively heated by micro wave radiation, whereas insulators or high energy gap semiconductors (some ionic chlorides, CaO, La2O3, and CeO2) are characterized by only weak heating. Attention should be given to materials and oxides used as components of carriers and catalysts. Whereas car bon and transition metal oxides reach high tempera tures, titanium and zirconium oxides are heated to moderate temperatures, and aluminum, magnesium,
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1679

and silicon oxides cannot be heated above 100 150. On the whole, the results presented in Tables 1 and 2 only give a qualitative idea of the rules governing heating of pure substances in microwave fields. Some of them, especially the data on heating of semicon ducting oxides and solids with a complex dependence of properties on stoichiometry, phase composition, temperature, and other parameters (transition metal oxides), require a more detailed analysis, which is beyond the scope of the present review. One of the most frequently discussed problems of heating by microwave fields is temperature measure ments, which are complicated by the presence of high intensity electromagnetic fields. Temperature sensors most frequently used in laboratories (metallic thermo couples, resistance thermometers, semiconducting elements) should be thoroughly screened and grounded to avoid sparking and obtainment of dis torted data. Such precautions are especially necessary if volatile and easily inflammable substances are heated in microwave fields. The temperature of the surface can be measured up to 3000 with the use of infrared and optical pyrometers. The use of devices that allow radiation intensity to be measured at several wavelengths simultaneously is especially effective; this allows the influence of indeterminate and/or changing optical properties of solids to be avoided (grayness coefficient, that is, discrepancy between the temper ature dependence of radiation intensity and the laws governing blackbody radiation). Let us consider particular mechanisms of micro wave field interactions with substances (microwave heating mechanisms). Dielectric Polarization One of the reasons for microwave heating is the ability of electric fields to polarize charges in sub stances and the inability of these charges to follow fast changes in field parameters. The total polarization is the sum of several separate effects, t = e + a + d + i, (3) where e is the electronic polarization, which is a con sequence of the rearrangement of electrons with respect to the atomic nucleus; a is the atomic polar ization, which appears when the relative arrangement of atomic nuclei in a molecule changes or when the charge in a molecule is distributed nonuniformly; d is the dipole polarization, which appears when the ori entation of a constant dipole in space changes under the action of the electric field; and i is the interphase polarization, which appears at the boundary between two phases with different permittivities (the Maxwell Wagner effect). The behavior of a substance in an alternating elec tromagnetic field depends on the ratio between the characteristic orientation and relaxation times and the frequency of field oscillations. The characteristic
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

Table 1. Temperatures of various liquid samples (50 ml) af ter the action of a microwave field, power 560 W, at a 2.45 GHz frequency for 1 min (Tmw) (heating from room temperature) and boiling temperatures (Tb) Liquid Water Methanol Ethanol 1 Propanol 1 Butanol 1 Pentanol 1 Hexanol 1 Chlorobutane 1 Bromobutane Acetic acid Ethyl acetate Chloroform Acetone Dimethylformamide Dimethyl ether Hexane Heptane CCl4 Tmw, C 81 65 78 97 109 106 92 76 95 110 73 49 56 131 32 25 26 28 Tb, C 100 65 78 97 117 137 158 78 101 119 77 61 56 153 35 68 98 77

Table 2. Heating of solids by microwave field; samples 25 g, heating from 20C in a household microwave furnace, pow er 1 kW, frequency 2.45 GHz ( is heating duration) Sub stance Al C Co3O3 CuCl FeCl3 MnCl2 NaCl Ni NiO SbCl3 SnCl2 SnCl4 ZnCl2
Vol. 84

Tmw, C 577 1283 1290 619 41 53 83 384 1305 224 476 49 609

, min 6 1 3 13 4 1.75 7 1 6.25 1.75 2 8 7

Sub stance CaO CeO2 CuO Fe2O3 Fe3O4 La2O3 MnO2 PbO2 Pb3O4 SnO TiO2 V2O5 WO3

Tmw, C 83 99 701 88 510 107 321 182 122 102 122 701 532

, min 30 30 0.5 30 2 30 30 7 30 30 30 9 0.5

No. 10

2010

1680

KUSTOV, SINEV

times of atomic and electronic polarization are con siderably shorter than the period of field oscillations in the microwave range, and these types of polarization therefore do not contribute to dielectric heating. Con versely, the characteristic times of dipole and inter phase polarization are comparable with the period of field oscillations and therefore make the major contri bution to substance heating by microwave fields. They should be considered separately. Dipole Polarization Dipole polarization d appears because substance molecules have dipole moments. At low frequencies, time during which the electric field changes its direc tion is longer than time required for a change in the orientation of dipoles, and polarization is in phase with electric field. Electromagnetic field provides energy necessary for a change in the orientation of dipoles. The energy transferred is very small, and the temperature of the substance almost does not change. If the frequency of the electromagnetic field is fairly high, it changes its direction more rapidly than dipoles can change their orientation. Since dipoles do not shift, field energy transforms into heat. In the microwave range of frequencies, time during which a field changes its direction is comparable with time necessary for changes in the orientation of dipoles. Dipoles rotate because of the action of electric field rotational momentum, but the resulting polariza tion lags behind field changes. When electric field is maximum and has a direct orientation, polarization can still be low. It continues to grow while field strength already decreases. Such a lag is evidence that a substance absorbs field energy and is heated. The dielectric properties of a substance are deter mined by two constants. The dielectric constant ' describes the ability of a molecule to be polarized by an electric field. It is maximum at low frequencies, when maximum field energy can be stored in a substance. The dielectric loss coefficient '' determines the effec tiveness of the transformation of electromagnetic field energy into heat by a substance. The '' value passes a maximum as the dielectric constant decreases. The dielectric loss slope ( tan = ''/') determines the degree of the transformation of electromagnetic energy into heat in a substance at given field frequency and temperature. The dielectric characteristics of any substance change fairly substantially as a function of frequency at given temperature [34]. For instance, for water (as a solvent in catalytic processes or a substance participat ing or formed in a catalytic process), the dielectric loss coefficient has a fairly large value over a broad fre quency range and passes a maximum at about 20 GHz, rather than at 2.45 GHz, as is usually believed. The depth at which field power decreases by two times is another important parameter for planning microwave

experiments or in the development of an industrial process with the participation of a microwave field. At small '' values, the depth of electromagnetic field penetration into a substance Dp can be approximately described by the equation

D p = ( ' ") ,
1/2

(4)

where is the electromagnetic field wavelength. Theoretical studies of the frequency dependences of '' and ' are based on the Debye equations [35]
2 2 'd = ' + ( '0 ') / (1 + ),

(5) (6)

'' = ( '0 ') / (1 + 2 2) , d

where ' and '0 determine the high frequency and static dielectric constant asymptotes and and are the frequency and relaxation time characterizing the time of polarization appearance and decrease. Equa tions (5) and (6) are applicable to both liquids and sol ids, although they are obtained using different models. An interesting feature of the Debye equations is the independence of the maximum ' and '' values from field frequency and relaxation time, (7) It is as a rule assumed that dipoles in liquids can have arbitrary spatial orientations and constantly change them because of thermal motion. The Debye interpretation of relaxation is based on the existence of viscosity forces in a substance. Debye used the Stokes theorem to obtain the following equation for the relax ation time of spherical dipoles: = 4r3/kT, (8) where is the viscosity of medium, r is the radius of the dipolar molecule, and k is the Boltzmann con stant. In solids, the thermal motion of atoms and ions is minimum, and a dipole has a set of equilibrium posi tions. They are separated by potential barriers, which should be overcome by the dipole that changes its ori entation. In the simplest case, there are two such posi tions, and the dipole moments of a molecule in them are anticollinear. According to the Boltzmann statistics, the number of transitions from one state into another is propor tional to (1 et/), where t is the time and is the relaxation time. We therefore obtain the following equation relating the relaxation time and dielectric constants:
'' = ( '0 ') /2 , 'max = ( '0 + ') /2. max

exp(U a kT )( '0 + 2) . n( ' + 2)

(9)

Here, 1/n is the time of a single vibration in the poten tial well and Ua is the difference between the energies of dipole states. The absorption value for such a model was obtained by Frohlich [36], who suggested that it should be equal to that obtained for liquids. It was proved that the following equation for the Onsager
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1681

model [37] was applicable to many liquids and solids [3840]: , (10) 9kT(2'0 + ') where N is the number of molecules and m is their mass. Importantly, energy absorption decreases as the temperature increases in this model. It was found in practice that most of polar liquids, for instance, aliphatic ketones with long chains and many ethers, had small dielectric loss values because of large differences between equilibrium position energies in the condensed state. In the majority of other cases, for instance, for many ethers with long chains, dielectric loss values are large at room temper ature and decrease as the temperature increases in agreement with the Onsager equation. Interphase Polarization A disperse system consisting of conducting parti cles in a nonconducting medium, for instance, a cata lyst is a nonuniform material whose dielectric proper ties exhibit pronounced frequency and, primarily, phase composition dependences. Loss in such materi als is related to the aggregation of charges at interphase boundaries and is known as the MaxwellWagner effect. However, its influence on substance heating in the microwave range has not been determined conclu sively as yet. Energy absorption according to this mechanism is concentrated at frequencies of ~109 Hz. Wagner showed that, for the simplest model of con ducting spheres in a nonconducting medium, the dielectric loss coefficient for the volume material frac tion is determined by the equation

'0 ' =

4Nm 2'0( ' + 2)2

loss primarily depends on conductivity. For high con ductivity materials (catalysts), there are such conduc tivity values that conductivity loss is much larger than dipole relaxation effects if these values are exceeded. For alumina at room temperature, conductivity loss becomes noticeable at low frequencies only. Loss in the microwave range largely occurs because of dipole relaxation. The conductivity of a material increases as the temperature grows, and the fraction of conductivity loss in the microwave range increases and becomes comparable with polarization effects. An increase in alumina conductivity as the temperature grows is related to thermal activation of electrons, which pass from the 2p oxygen valence zone to the 3s3p conduction zone. In addition, material electronic conductivity as a rule increases in the presence of defects, which create additional levels in the forbidden band. The concentration of defects can also substan tially increase at high temperatures. THE USE OF MICROWAVE EFFECTS IN CATALYSIS Heterogeneous catalysis is a basis of many science intensive technologies used in such industrially impor tant processes as the utilization and reprocessing of wastes of many chemical and petrochemical pro cesses, purification of thermal station and car exhaust gases, waste free syntheses of valuable chemical prod ucts, oil and gas processing, and the production of high octane fuels. The activation of heterogeneous catalysts both at the stage of their preparation and dur ing use is of the greatest importance for increasing the activity and selectivity of catalytic processes. Tradi tional methods for the activation of heterogeneous catalysts, such as thermal treatment and oxidation or reduction at high temperatures in a medium of various gases, often do not provide the necessary catalyst activity, selectivity, or stability, because such treat ments generate a wide spectrum of active centers of various natures and strengths, including active centers than catalyze undesirable side reactions. In addition, currently, ever increasing use in industry is found by catalysts that differ from systems used earlier by more complex compositions, the organization of the struc ture and active centers at the molecular level, and polyfunctionality of action. This makes the problem of the development of new nontraditional methods of their activation and regeneration still more pressing. If microwave activation is used in catalytic pro cesses, it should be taken into account that the objects of study are nonuniform multiphase systems including catalysts and gases in which volumetric structural and property changes occur under the action of electro magnetic radiation. The main idea of microwave catalysis is to exert volumetric controllable electro magnetic action on the catalystreagents system. This action should change the state of the system and increase the effectiveness of the work of catalysts, the
Vol. 84 No. 10 2010

'' = i

where is the conductivity of the conducting phase and ' is its dielectric constant. Correctness of Eq. (11) was proved experimentally [41]. Conductivity Loss As mentioned above, conductivity and related Ohmic loss determine one of the microwave heating mechanisms that can operate in catalytic systems. When a microwave field acts on a medium containing mobile charge carriers, conductivity currents appear in the system. Taking into account loss related to medium resistance, the complex permittivity of a sub stance takes the form

9 f (1.8'10 )(1 + ),
max 10 2 2

(11)

( '0 ') j . (12) (1 + jt ) 0 The real part of this value is determined by the Debye theory. The loss part determines not only relaxation but also Ohmic loss. The contribution of conductivity * = ' + i
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

1682 , % 2 80 60 40 20 1

KUSTOV, SINEV

A solution to the Arrhenius equation with the temper ature of the superheated region increased by 70 K gave an almost twofold increase in the concentration of products compared with uniform temperature distri bution over the catalyst during the first seconds of heating (figure). The calculations were performed using the k = 1.45 1013 mol1 s1 preexponential fac tor and E = 80 kJ/mol activation energy. The next example of thermal microwave effects is the formation of so called hot points. Traditionally, regions with a temperature higher than that of the environment are called hot points. They appear as a result of stronger interaction with a microwave field and low effectiveness heat exchange. Hot points are difficult to observe in heterogeneous catalysts of gas phase reactions, because these zones can have atomic sizes in such typical cases as metallic catalysts or oxy gen defects in oxides. Because of poor resolution of infrared pyrometers, direct temperature measure ments cannot be performed, and the real temperature of hot points can only be determined by indirect, as a rule, kinetic methods. Perry et al. [23, 24] discuss the presence of substan tial differences between microwave and thermal acti vation for the example of the oxidation of CO on Pt/Al2O3 and Pd/Al2O3. They studied the kinetics of the reaction, specifically, Arrhenius curves (the depen dences of the rate of formation of 2 on inverse tem perature), to find that, for both catalysts, the rate of the reaction at a given temperature was higher under microwave activation conditions compared with ther mal activation. At the same time, the form of the dependence of the reaction rate on reagent concentra tions and activation energy did not depend on the method of heating. Note that the authors of [23, 24] used a fairly doubtful procedure for temperature mea surements, they introduced a thermocouple into the volume of the catalyst after switching microwave field off. However, they performed a more detailed analysis with corrections in temperature to draw the conclu sion that there was no difference between microwave and thermal activation. Moreover, Perry advanced the hypothesis that isolated microparticles of materials (including metals) absorbing microwave fields on a substrate transparent to electromagnetic waves could not be heated to temperatures substantially higher than the mean temperature of the environment. This hypothesis was theoretically substantiated by Perry [24] by calculations of heat release in spherical metal lic nanoparticles and subsequent heat transfer into the gas phase under stationary conditions. The differences between the temperatures of the surface of the metal and the surface of the carrier calculated for 1 and 100 nm particles were of 1.1 1010 and 1.6 1010 K, respectively. Mingoos and Zhang studied the catalytic decom position of H2S on 30% MoS2/ Al2O3 to find [27] that, under the action of a microwave field, conversion
Vol. 84 No. 10 2010

12

16 , s

Time dependence of product yield () under the condi tions of (1) uniform temperature distribution and (2) the presence of superheated regions on the catalyst of naph thalene sulfonation [45].

selectivity of raw material conversion into valuable products, and catalyst stability. One of the main direc tions toward solving this problem is a decrease in the temperature and other reaction parameters by process execution under mild conditions in an electromag netic microwave field with the retention of or increase in process efficiency. Numerous data show that, under microwave acti vation conditions, higher reaction rates are obtained than in thermally activated processes, contact time decreases, and the yields of final products increase. These advantages are often ascribed to so called microwave effects without performing a detailed analysis of the corresponding reaction conditions. It is likely that such a microwave effect was for the first time observed by Sun in his work on the hydrolysis of adenosinetriphosphate [42]. The results obtained by Sun and the conclusions drawn by him were later cor rected by Jahngen, who found that an increase in the reaction rate by 1215 times under microwave field action conditions was likely caused by temperature gradient [43]. Since then, it is accepted to separate thermal and nonthermal (electromagnetic) microwave effects. Mingos and Baghurst discussed the theoretical basics of dielectric heating in microwave fields and gave numerous examples of its use [44]. THERMAL EFFECTS A typical thermal microwave effect was obtained for the example of the catalytic sulfonation of naphtha lene [45]. The regioselectivity of the introduction of sulfo groups can be controlled by the power of micro wave action and the resulting rate of heating. The experimental data closely agree with a theoretical analysis based on the hypothesis of the existence of superheated regions on the surface of the catalyst [46].

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1683

was higher that predicted by analyzing thermody namic equilibrium at a mean measured system tem perature and higher than that obtained in experiments with usual thermal activation. X ray patterns showed that the catalyst after microwave activation contained the Al2O3 phase, which could be formed from Al2O3 only at temperatures above 1000, whereas the mean measured temperature in experiments did not exceed 800. On the other hand, initially amor phous molybdenum disulfide with 150170 nm parti cles transformed into hexagonal crystals under micro wave field action. Because the melting temperature of MoS2 is 1185, this substantiates the formation of hot points including both active phase and carrier par ticles in the catalyst. NONTHERMAL EFFECTS As distinct from thermal microwave effects, non thermal effects appear much rarer. Parmon et al. used a specially designed unit to study the epoxidation of ethylene with separating thermal from nonthermal effects [47]. The catalyst was heated not only by a microwave field but also by pumped hot air. To decrease convective heat loss, the reactor and part of air introduction line situated in a microwave furnace were placed into an evacuated shell. The specific action of a microwave field was only observed for the Ag/ Al2O3 catalyst subjected to reducing treatment before measurements. In an oxi dative atmosphere, no differences were observed between experiments performed with the use of a microwave field and in its absence. This effect is dis cussed taking into account another observation, according to which the formation of defects in silver volume and on the surface is decelerated in a reducing atmosphere. According to Roussy, structural changes or electron behavior effects are one of the possible reasons for an increase in selectivity in the isomeriza tion of 2 methylpentane on Pt/Al2O3 [48, 49]. Another nonthermal effect was considered for the BaBiO3 x catalyst of the oxidative condensation of methane [50]. An increase in selectivity with respect to C2 products when microwave activation was used was explained by a decrease in the rate of carbanion CH3 oxidation because of a decrease in the concentration of oxygen adsorbed by the catalyst. Different behaviors of catalysts activated by microwave field and heating were substantiated by measurements of the real and imaginary complex permittivity components under microwave field conditions (single mode high power microwave field heating catalysts) and in a classic res onator (broad band conductivity measurements [51] with a small microwave power supply after sample thermal treatment). When the catalyst reached a tem perature higher than 500, the imaginary permittiv ity component increased in oxygen and air. Con versely, dielectric loss decreased in an inert atmo
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

sphere (helium). The authors explain these effects by deceleration of redox processes under microwave field action. It is stated that metallic behavior corresponds to a maximum degree of BaBiO3 x system reduction (x = 0.49), and the conclusion is drawn that electro magnetic field decreases the rate of oxygen absorption at arbitrary oxygen pressure. PROMISING DIRECTIONS OF STUDY OF THE USE OF MICROWAVE FIELDS IN THE CATALYSIS OF GAS PHASE PROCESSES Among gas phase catalytic reactions that most fre quently were subjected to stimulation by microwave fields, the following processes should be mentioned: oxidative condensation of methane; partial oxidation of hydrocarbons; purification of gases from noxious impurities, including purification of car exhausts; hydrogenation of aromatic compounds; dehydrogenation of naphthene compounds. In the first two reactions, which are of great impor tance for applications, the key effectiveness factor is the attainment of high selectivity with respect to the desired products at equal degrees of reagent conver sion. The catalytic neutralization (purification) of var ious gases and gas exhausts is also a rich field for stud ies and developments. High heating rates combined with easy power control when a microwave field is used broaden the possibility of a decrease in the content of various contaminants in gases. In addition, in the last case, the use of microwave heating performed almost instantaneously solves the cold start problem, which usually arises with inertial thermal heating. The Oxidative Condensation of Methane The oxidative condensation of methane, the main desired products of which are ethylene and ethane, was for the first time described in the early 1980s. Since then, this reaction has been the object of serious scientific interest. Currently, the practical importance of this process again increases, because the use of nat ural and casing head oil gases as raw materials for chemical industry is an acute problem under the con ditions of constantly decreasing oil reserves and a sharp increase in oil cost. Numerous studies of this process showed that there is a kinetics limit of the yield of the desired products (2025%). This limit is determined by the ratio between methane and 2 hydrocarbon reactivities. At comparatively low conversions (1015%), a 8085% selectivity can be reached. Selectivity decreases con siderably as the conversion of methane increases. Free radical reactions substantially influence the process, because the primary reaction product, ethane, is formed in the recombination of methyl radicals gener
Vol. 84 No. 10 2010

1684

KUSTOV, SINEV

ated as a result of methane activation with the partici pation of catalyst surface active centers [52]. According to several authors, the use of microwave activation can cause a decrease in the temperature of the gas phase and catalyst volume. As a consequence, it has an advantage over convective heating activation because of suppression of deep oxidation reactions. The first studies in this direction were performed by Bond [53] with the use of sodium aluminate as a cata lyst. On the whole, selectivities with respect to 2 hydrocarbons obtained under thermal and microwave heating conditions did not differ strongly. But reaction temperatures differed by more than 400 K. Work [53], pioneering for microwave catalysis, was not free of certain shortcomings. Apart from different methods for temperature measurements under different activa tion conditions, the authors discuss the results giving no process characteristics other than selectivities with respect to 2 hydrocarbons and deep oxidation prod ucts (CO and 2). Apart from a substantial decrease in reaction temperature, the composition of reaction products changed, selectivity with respect to 2 decreased, and selectivity with respect to CO increased. Under microwave activation conditions, the ethylene/ethane ratio increased, whereas ethane was the main product under thermal activation condi tions. Roussy also studied several mixed oxides and com pared their activity in a microwave field with that under traditional heating conditions. These oxides included La2O3, La2Zr2O7, SmLiO2, (SmLiO2)0.8(CaOMgO)0.2, Li/MgO, and BaBiO3 x [54]. No difference between thermal and microwave heatings was obtained for La2Zr2O7, whereas La2O3 showed higher selectivities with respect to the desired products at equal conver sions under thermal heating conditions. SmLiO2 and (SmLiO2)0.8(CaOMgO)0.2 showed opposite depen dences of selectivity on conversion: at low conver sions, selectivity with respect to 2 hydrocarbons tended to 100% under microwave activation condi tions and to zero when traditional heating was used. At high methane conversions (3040%), selectivities became close to each other (about 5060%). The authors of [54] explain the observed effects by different ratios between the rates of homogeneous and heterogeneous stages leading to the formation of the desired and side products under different heating con ditions. The kinetic patterns of the process were ana lyzed using a simplified scheme,
CH4
k1

C2H6
k5

k3

C2H4

k2

k4

COx

The k4 and k5 values were substantially higher under traditional heating conditions, which corre sponded to the observed substantial difference in selectivity with respect to the desired products. It was assumed that the reason was the occurrence of deep

oxidation of 2 hydrocarbons predominantly in the gas phase, and the measured temperature of the gas phase was substantially lower under microwave activation conditions. It was also assumed that active centers responsible for the formation of 3 radicals were specifically and selectively activated by a microwave field in samples characterized by high dielectric loss values (SmLiO2 and (SmLiO2)0.8(CaOMgO)0.2). The influence of the temperature of the gas phase was studied in detail for the Li/MgO and BaBiO3x catalysts, for which an important role was played by different mechanisms of the formation of methyl rad icals, which dimerized to yield ethane [55]. For BaBiO3x, the existence of surface oxygen ions O2(s) is supposed. These ions can detach a proton from the methane molecule with the formation of the CH3( s ) surface ion. Carbanions interact with molecular oxy gen to produce methyl radicals [56]. The second mechanism was established by [57] for the Li/MgO catalyst. It was based on the existence of O ions on the surface of the catalyst. These ions could detach a hydrogen atom from methane to produce methyl rad icals, which dimerized in the gas phase to yield ethane. It was assumed that the formation of undesirable prod ucts (carbon oxides) largely occurred heterogeneously and homogeneously when the reaction was governed by the first and second mechanisms, respectively. The influence of the gas phase oxidation of methyl radicals was proved for the Li/MgO catalyst by diluting it with a material (silica) inert with respect to microwave fields. Microwave activation of the diluted sample gave selectivity with respect to 2 products comparable with that obtained using traditional heating. To explain this result, the authors advanced the hypothe sis according to which microwave activation increased selectivity with respect to 2 hydrocarbons as a result of a decrease in the rate of the gas phase oxidation of methyl radicals, because gas surrounding the catalyst remained comparatively cold. For the BaBiO3x cata lyst, dilution with an inert substance did not influence selectivities with respect to 2 hydrocarbons of differ ently activated reactions, which, according to [57], could be evidence of a lower contribution of gas phase reactions on this catalyst. Catalysts with perovskite structures were studied by Chen et al. [5860]. Depending on their composition and structure, perovskites can have very different elec tromagnetic properties; they can be insulators, metal lic conductors, superconductors, materials with giant magnetic resistance (these materials change their elec tric resistance by orders of magnitude in the presence of a magnetic field), and dielectrics [61]. Studies were per formed for oxides classified by the authors as oxygen conducting materials, namely, SrCe0.95Yb0.05O3, BaCe0.93La0.07O3, and Li2SO4/BaCe0.93La0.07O3. It was assumed that the detachment of hydrogen from the methane molecule played a key role in the oxidative condensation of methane. In addition, the perovskites
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1685

Table 3. Influence of temperature and method of heating (I, microwave and II, thermal) on the oxidative condensation of methane on proton conducting perovskite like catalysts Catalyst SrCe0.95Yb0.05O3 SrCe0.95Yb0.05O3 BaCe0.93La0.07O3 BaCe0.93La0.07O3 Li/SrCe0.95Yb0.05O3 Li/BaCe0.95Yb0.05O3 Li/SrCe0.93Yb0.07O3 T, C 580 775 590 825 590 620 590 Conversion of 4, % 20 20 25 25 14 18 18 Selectivity, % Method C2H2 1 0 2 0 0 0 0 C2H4 23 29 30 38 27 28 34 C2H6 30 32 29 26 49 45 42 CO 9 4 6 1 1 1 2 CO2 37 35 33 35 23 26 22 I II I II I I I

Table 4. Influence of the method of heating on the oxidative condensation of methane on oxygen conducting mixed oxide catalysts Method of heating Thermal Microwave Microwave , 840 525 580 Conversion of 4, % 20 20 30 Selectivity, %
+ C2

20 11 12

2 18 17 23

24/26 2.8 1.3 1.8

62 72 65

specified were easily heated by a microwave field to reaction temperatures. A decrease in reaction temper ature by ~200 K was observed, the other conditions being equal (see Table 3). The distribution of products for microwave and thermally activated processes changed insignificantly. The oxidative condensation of methane in the presence of mixed oxide catalysts (Bi2O3)1x(WO3)x (x = 0.2, 0.3, and 0.4) was studied in [63]. According to the authors (see Table 4), equal conversions were attained when microwave activation was performed at a lower temperature compared with the thermal pro cess. In addition, selectivity with respect to 2 hydro carbons increased under microwave conditions, mainly because of a decrease in the contribution of deep ethane oxidation to CO and 2. Note that the ethane : ethylene ratios were substantially different for two methods of activation of the oxidative methane condensation process. The results were discussed on the assumption of interrelation between the mobility of lattice oxygen and catalytic properties [64] on the one hand and the influence of microwave fields on the transport of ions [65, 66] on the other. On the whole, the results obtained in studies of the influence of microwave activation on the oxidative condensation of methane are contradictory. On the one hand, effects described in some publications are fairly substantial. On the other hand, some of the results contradict the general rules governing the pro cess (see above), such as the existence of the kinetics limit independent of the particular mechanism and
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

localization of its separate stages. A decrease in the temperature required to reach equal conversions by 200300 K under microwave field conditions with the retention of and even increase in selectivity with respect to 2 hydrocarbons is also questionable. According to the current concepts of the heteroge neous homogeneous character of oxidative methane condensation (e.g., see [6769]), a decrease in tem perature below 650 should cause a sharp decrease in selectivity in the presence of molecular oxygen in the gas phase. This is caused by a shift to the right of equi librium
i i CH3 + O 2 CH3O 2 and subsequent formation of oxygen containing prod ucts (up to carbon oxides) through CH3O2 radicals. Methyl radicals are then removed from the process, and the formation of 2 hydrocarbons is blocked.

It is hardly probable that the use of microwave fields changes the mechanism of methane activation (see above the conclusion of the impossibility of chemical bond dissociation and/or electronic excita tion of molecules under the action of microwave radi ation). Therefore, the probability of the mechanism with the generation of primary methyl radicals remains fairly high. Taking into account difficulties of temperature measurements (especially, separate mea surements for a catalyst and a free gas in a granular layer), it should be stated that the results cited above need independent verification before drawing conclu sions about the special and effective action of a micro
Vol. 84 No. 10 2010

1686

KUSTOV, SINEV

Table 5. Microwave activated partial oxidation of C1C6 hydrocarbons Hydrocarbon Methane Methane Propane Propylene n Hexane Cyclohexane Catalyst Ni 1404* Ni (microstructured) Ni (1 m) CuO uO V2O5 Reaction conditions H2O (vapor) Pulsed microwave activation, H2O (vapor) H2O (vapor) H2O (vapor) H2O (vapor) H2O2 Products Methanol, acetone, dimethyl ether Acetone, C2, C3, methanol, ethanol Methanol, butanols, propanols Propanol, ethanol, acetone, propylene oxide Methanol, propanol, hexanone Cyclohexanol, cyclohexanone

* Commercial sample of the Ni/NiO composition.

wave field on the catalysts of oxidative methane con densation and the corresponding process. Partial Oxidation and Other Hydrocarbon Conversion Processes The catalytic partial oxidation of organic com pounds of various classes is a basic reaction of many fundamental organic synthesis and petrochemistry processes. In the past decades, such processes with the participation of hydrocarbons, largely saturated hydro carbons, have attracted special attention because of rise in the cost of oil and the necessity of the involve ment of other raw material resources in the production of valuable chemical products and half finished prod ucts, in the first place, of natural and casing head gases. These processes lie close to various variants of the transformation of hydrocarbons and other organic substances into synthesis gas and/or hydrogen oxida
Table 6. Influence of temperature and method of heating (I, microwave and II, thermal) on the partial oxidation of methane according to [71, 72] Method T, K Conversion of 4,% Selectivity, % 2 21 1 34 7 26 <1 23 8 43 <1

I I II II I I II II I I

723 1073 873 1073 673 973 873 1073 723 1073

10% Co/ZrO2 48 79 100 99 63 66 94 93 10% Ni/La2O3 58 74 100 100 40 77 91 92 10% Co/La2O3 34 57 100 100

tive, vapor, and carbon dioxide conversions and their various combinations. The first experiments with the use of microwave fields in such processes were described by Wan [13]. That work contained the primary data on vapor con version with various catalysts, reaction conditions, and initial hydrocarbons (see Table 5). Water vapor was added into the reaction mixture by water vaporiza tion or adding water containing salts, such as CuSO4 2H2O [70], to catalysts. According to the authors, microwave fields initiated the dissociation of adsorbed water to the and radicals. Although the yields of oxygen containing compounds were fairly low (~5%) under microwave methane activation condi tions, the authors indicate the potential possibility of using water as a cheap and ecologically pure oxidizer in such reactions. Bi and Dai studied the partial oxidation of methane to CO and H2 on 10% Co/ZrO2, 10% Ni/La2O3, and 10% Co/La2O3 [71, 72]. They explain the difference of 50250 K between the temperatures of reaching simi lar conversion values under microwave and thermal activation conditions by the formation of hot points on oxygen defects. A comparative review of the results obtained is given in Table 6. Liu and Iin [73, 74] describe the partial oxidation of xylene and toluene on vanadium catalysts depos ited on TiO2 and SiO2. Reactions were performed in a modified household microwave furnace. At this stage, only these authors comprehensively described experi ments in a multimode furnace. They, however, used a procedure in all the experiments that could not but raise natural questions. It is not quite clear why a mer cury thermometer is used and what is then the accu racy of temperature measurements under microwave field conditions. It is also clear from [74] that a peri odic microwave activation mode was used, but process parameters (power, duration, and pulse ratio) were not reported. One way or another, the authors report about an increase in the conversion of xylene under micro wave activation conditions compared with thermal heating. Selectivity with respect to phthalic anhydride decreases as the temperature increases. These observa tions are explained by stronger interaction of V2O5
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1687

particles, which are the active phase of the catalyst, with microwave field compared with the substrate. Their temperature is therefore higher than the mean catalyst temperature. Microwave field prevents active phase particles from segregation observed under ther mal heating conditions. The distribution of the V2O5 active phase over the surface of SiO2 becomes more uniform, and dispersity increases. As a result, the number of catalytically active V=O centers on the sur face of the catalyst grows. When the temperature of the active phase exceeds 370, deep oxidation processes begin to prevail, which decreases selectivity with respect to phthalic anhydride. According to the authors, this is caused by the superheating of V2O5 par ticles and the formation of electric discharges on the surface of the catalyst. The Removal of Sulfur and Nitrogen Compounds The removal of SO2 and NOx from air is one more heterogeneous catalysis area that has been extensively explored during the past decades. It spite of certain successes achieved, the problem is still far from com plete solution, especially because of a great number of various objects in which these contaminants are formed and conditions of their removal. Strictly, a sub stantial part of works concerned with the removal of sulfur and nitrogen compounds from gas mixtures cannot be treated as exclusively catalytic, because complex technological approaches are used in them. This is a combination of the adsorption, catalysis, and stoichiometric reduction of oxides; the stimulating action of microwave fields is used at various stages of contaminated gas and/or solid catalyst processing. Currently, the technology of the selective catalytic reduction of nitrogen oxides (SCR DeNOx) holds the leading position in the removal of nitrogen oxides from waste gases of power stations. Nitrogen oxides are reduced by ammonia or urea at temperatures of from 300 to 400 on MoO3/WO3/V2O5 mixed oxides deposited on TiO2. This is a highly effective process, but it requires heavy material expenditures because of corrosion activity of ammonia and the occurrence of side reactions. For instance, ammonia at high temper atures is oxidized to NO. In addition, in the presence of sulfur containing compounds in a fuel, ammonium sulfide and sulfur oxides are formed. Their removal is a nontrivial problem. However, during the past decade, technologies were developed for abandoning the use of ammonia as a reducing agent in favor of less expensive and more friendly ecologically hydrocar bons. Currently, there are no catalysts stable toward moisture contained in all waste gases. The first works on the use of microwave activation in these processes were published by Wan [13]. Exper iments with the removal of sulfur and nitrogen oxides were performed on catalysts similar to those described above for partial oxidation. Gas mixtures consisting of 9% SO2 or 25% NO diluted with an inert gas (nitrogen
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

or helium, respectively) were used. With the Ni 1404 catalyst, 99% of SO2 was removed during 3 h. Similar results were obtained for all the other catalysts except commercial Ni 1600 nickel oxide catalyst, which was rapidly poisoned by sulfur compounds. The other cat alysts were successfully regenerated in a flow of nitro gen in a microwave field and could be repeatedly used up to 10 times without a noticeable decrease in activ ity. When NO was decomposed on nickel catalysts under microwave activation conditions, a more than 99% conversion of initial nitrogen oxide was observed. Catalysts remained active for 5 h, however, after 4 h of work, N2O was observed at the exit of the reactor in addition to N2 and O2; the content of N2O increased with time. Note that the authors of [13] give no com ments concerning the results obtained in microwave activation of the removal of sulfur and nitrogen oxides. Moreover, they give no suggestions concerning the mechanism of microwave field action on these pro cesses. In addition, data are incomplete in the work cited: the results are not compared with the thermal activation data, and such important parameters as the temperature of catalysts and microwave activation conditions are not given. Starting with the mid 1990s, Cha has been study ing the destruction of air contaminants in microwave fields [75, 76]. He obtained several patents in this field [7779]. As distinct from Wan, Cha used various car bon adsorbents with different structures. Equal adsor bent samples were saturated with a mixture of 1000 ppm NO and 610 ppm NO2 in a humid air (water vapor content 0.95%). All the samples were then treated by a microwave field (power 480 W, frequency 2.45 GHz). During treatment, desorbed substances were recorded. After the desorption stage (30 min), weight loss and specific surface area were measured for all the adsorbents. According to the authors, the regeneration of carbon adsorbents saturated with nitrogen oxides activated by a microwave field increased both sorption capacity and the rate of adsorption for all the samples except GAC 830 (Atochem) commercial activated carbon. The pres ence of moisture in the initial mixture contributed to the transformation of NO into NO2 and HNO3, which were then adsorbed. According to the authors, the rea son for the improvement of the sorption properties of the samples was a higher ability of the centers of NO2 and HNO3 adsorption to absorb microwave field com pared with the surrounding volume (it is, however, not clear from the paper why the authors arrive at such a conclusion). These centers then participated in the reduction of nitrogen oxides. Although such a method for the removal of nitro gen oxides is not completely catalytic, it has several obvious advantages. When carbon adsorbents are used, the presence of reducing agents (such as tradi tionally used very toxic ammonia) is unnecessary, and the reduction of nitrogen oxides proceeds at a high rate under these conditions. In addition, because of the
Vol. 84 No. 10 2010

1688

KUSTOV, SINEV

formation of HNO3, carbon sorbents are additionally activated, which can be a source of a valuable side product (activated carbon). The authors also note that this method is applicable to the purification of gas exhausts from comparatively small sources, such as household diesel generators and heaters, when the transformation of nitrogen oxides and nitric acid into valuable products is impossible. The first studies of the reduction of nitrogen oxides in a microwave field with methane in the presence of oxygen on zeolite catalysts were described by Chang [80]. The author used Co NaZSM 5 and Co HZSN 5 and the H forms of the same zeolites. He compared microwave and convective heating regimes. The experimental data show that the reduction of NO pro ceeds much more actively and at lower temperatures under microwave activation conditions. For instance, the H ZSM 5 and Co HZSM 5 catalysts were almost inactive under convective heating conditions at tem peratures below 400. The Co NaZSM 5 catalyst catalyzes the deep oxidation of methane without the involvement of NO into the process at temperatures below 420. At higher temperatures, the reduction of NO begins, but the oxidation of methane with oxygen proceeds more actively, as previously. Conversely, the action of microwave field at the same temperatures results in high degrees of the transformation of NO on all the three catalysts. On both cobalt catalysts, the conversion of NO exceeds 68%, whereas the H form is less active in the reduction of NO. These data lead the author of [80] to conclude that a high degree of the absorption of microwave energy by catalysts changes reaction paths and results in the predominant activa tion of the reduction of nitrogen oxide by methane. Among possible mechanisms of microwave activation of NO, the author of [80] mentions the formation of nitrogen dioxide, which is known to more actively react with organic substances. Another possible mech anism of the destruction of the NO bond is interac tion with the methyl radical formed in the homolytic abstraction of hydrogen from the methane molecule by surface oxygen. Zhang studied other catalysts of the same type, In/HZSM 5 and In Fe2O3/HZSM 5 (1 : 8 : 20 by weight) [81]. Under thermal heating conditions, the two catalysts show equal activities, whereas, under microwave activation conditions, the In/HZSM 5 catalyst is absolutely inactive. The catalyst containing iron oxide was active in a microwave field; moreover, nitrogen oxide conversion reached 100% if a water free mixture was used. According to the authors, iron oxide well absorbs microwave field energy, whereas the In/HZSM 5 catalyst almost does not absorb electro magnetic waves, which explains the effect observed. Since Fe2O3 is a good chemisorbent of nitrogen oxides, the latter are activated more easily. Note that the observed temperature of catalyst volume necessary for the complete conversion of NO decreased by 200 K in microwave compared with convective heating experi

ments. Mingos et al. [82] observed the formation of superheated regions 0.091.0 mm in diameter under the action of microwave radiation on MoS2/ Al2O3 during the endothermic decomposition of H2S. The catalyst volume temperature at which equal process rate was achieved decreased by 150200 K compared with convective heating. One of the possible methods for creating adsorp tion catalytic systems for the protection of the envi ronment, in particular, for the removal of volatile organic compounds, is the addition of components strongly interacting with electromagnetic fields to adsorbents and catalysts already studied under thermal activation conditions [83]. Several patents [84] and publications [85] are concerned with the development of catalytic systems for the purification of car exhausts under cold start conditions. The influence of micro wave radiation on the catalytic purification of car exhaust gases was studied with varying radiation power from 0 to 200 W [86]. The catalyst was the reduced Pt PdRh/Al2O3 system on a ceramic monolith. Micro wave and thermal heating conditions were compared. Microwave radiation allowed the temperature of reac tion initiation to be decreased substantially even in the presence of water vapor and catalyst activity at low temperatures to be increased. The mechanism of the effect was not studied and explained, but the sugges tion was made that adsorbed particles were activated by microwaves directly or through catalyst particles. Similar data were published in [74]. It was shown in [24, 86] that temperature gradients within a catalyst granule were not very large, and that the temperatures of active particles and substrate were equal, for instance, in the oxidation of CO on Pd/Al2O3, even when the substrate itself did not absorb microwave energy. According to the authors, this con clusion is valid for a wide range of metal particle sizes. Studies performed led the authors to conclude that, at reliable temperature measurements for the oxidation of CO, the reaction proceeded identically and did not depend on the source of heat (microwave or usual heating). Thomas [26] did not agree with this conclu sion. According to Thomas, substantial temperature gradients between metal particles and substrates can be obtained by varying particle size and microwave radiation frequency. An important conclusion that fol lows from the discussion on this point in the literature is the necessity to take great care with communica tions in which the authors assert that they were able to decrease reaction temperature with increasing or retaining activity or increase activity at a constant temperature. Other Catalytic Processes with Microwave Activation Other directions in catalysis related to the use of microwaves were discussed in reviews [87, 88]. In [89] and several other publications by the same authors, microwave activation of the PdFe/Al2O3 catalyst of
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1689

hydrodechlorination was studied. It was shown that the size of metallic deposited catalyst particles changed under the action of microwaves, and that metal particles could form alloys. The possibility of increasing the activity of reduced catalysts of chlo robenzene hydrodechlorination was shown. Micro wave heating had the same effect (enlargement of par ticles) on the Pd/Nb2O5 catalyst of the same reaction [90], which increased the activity and stability of the catalytic system. Similar results were obtained for microwave irradiated Pd/Al2O3 and Pd/SiO2 catalysts of the hydrogenation of benzene. Conde et al. [91] showed that microwave radiation frequencies heavily influenced the effects related to microwaves in catalysis. They performed the oligo merization of methane on nickel catalysts. Changes in microwave frequencies changed field distribution and the set of observed reaction products. At all the fre quencies studied, an increase in radiation power increased the conversion of methane and decreased selectivity with respect to 2 hydrocarbons. At a 4.6 GHz frequency, an increase in power decreased selectivity with respect to ethane and ethylene but increased selectivity with respect to acetylene and benzene; at 2.45 GHz, acetylene did not form. An increase in frequency at a constant power increased selectivity with respect to benzene. The use of a gas diluent (helium) also changed the picture of selectivi ties under microwave conditions. The results obtained in studies of the influence of radiation frequency led the authors to conclude that the most probable reason for changes in process characteristics was a change in the real temperature of the process in the zone of transformations rather than some specific microwave activation of reaction participants or catalyst. This conclusion was also substantiated by the insufficiency of microwave energy for splitting chemical bonds in the system in question. The ability of a material to transform microwave energy into heat was shown to pass a maximum as the frequency increased. It follows that frequency variations can be used to influence the amount of heat released and the real temperature of catalyst active centers. Microwave heating was also used to perform other reactions, namely, the hydrogenation of chlorinated phenols on Pt/C [92]; the isomerization of 2 methylpentane on Pt/Al2O3 [93]; the epoxidation of ethylene on Ag/Al2O3 [94]; the hydrogenation of olefins, hydrocracking of cyclic hydrocarbons, and water shift reaction CO + 2 = 2 + 2 [11]; arylation according to Heck on palladium cata lysts deposited on Al2O3, C, MgO, and CaCO3 [95]; the oxidation of toluene to benzoic acid on V2O5/TiO2 [73];
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

and several other processes. The decomposition of 4 chlorophenol to 2, HCl, and 2 by oxidation with air on nickel oxide catalysts (pH 7, 70C, t = 5 min) was performed under microwave activation conditions in situ in [96]. Sev eral examples of successful use of microwave radiation in the hydrogenation and hydrogenolysis of alkenes and alkanes and water shift reaction are given in [97]. We thoroughly studied the dehydrogenation of sev eral naphthene and polycyclic (including heterocy clic) compounds on deposited metallic catalysts under the action of microwave heating [98]. The dehydroge nation of such compounds was suggested as one of two stages of safe keeping of hydrogen; it is based on cyclic reversible reactions of the hydrogenation of aromatic (heteroaromatic) compounds followed by the dehy drogenation of the saturated substance obtained. In such cycles, capacity no less than 77.5 wt % of the composite material, including the organic substrate and catalyst, was reached. This is much higher than the capacity of intermetallic compounds traditionally used for hydrogen storage (34 wt %) or carbon nano tubes suggested earlier (12 wt %). Oxidation reactions. Interesting data on the use of microwave activation in the oxidation (epoxidation) of styrene were reported in [99]. The CoYZrO2 sul fated catalyst was used. The optimum microwave radi ation power for it was found to be ~400 W; it decreased the time of the reaction to several minutes, while the conversion of styrene and selectivity with respect to epoxide remained high (91% at 120C). An increase in power to 800 W resulted in the formation of a new product, styrene glycol. Effective use of microwave heating in the oxidation of alkenes and alcohols under the action of hydrogen peroxide was described in [100]. The catalyst was MSM 41 type mesoporous titanium silicate modified by organic compounds. It was recycled several times without activity loss. A positive effect of microwave radiation was obtained in the oxidative dehydrogena tion of ethylbenzene to styrene on iron oxide depos ited on multilayer carbon nanotubes [101]. An increase in selectivity with respect to styrene was observed under microwave activation conditions (380450C); it was found that part of Fe2O3 trans formed into Fe0 nanocrystals. The oxidation of benzene to phenol with the use of nitrous oxide as a mild oxidizer is a very promising one stage method for the production of phenols. Microwave activation of the zeolite catalyst of this reaction (Fe ZSM 5) was used to decrease the contri bution of coking [102]. Although some increase in selectivity with respect to phenol and a decrease in the rate of catalyst deactivation were observed, on the whole, no substantial improvement of process charac teristics was obtained. The influence of microwave radiation on the formation of acrylonitrile in the ammoxidation of glycerol (a side product of the pro
Vol. 84 No. 10 2010

1690

KUSTOV, SINEV

duction of biodiesel fuel from vegetable oils) under the action of hydrogen peroxide and ammonia on VSb Nb oxide catalysts was studied in [103, 104]. Selectiv ity with respect to acrylonitrile was quite substantial under microwave heating conditions (83.8%) at a 46.8% conversion. For comparison, the reaction under thermal activation conditions gives selectivity of only ~60% at a conversion not exceeding 15%. Organic synthesis. Organic synthesis of com pounds of various classes is the most extensively devel oping area of microwave technology applications [105107]. In these studies, both homogeneous (e.g., see [108]) and heterogeneous (e.g., see [109]) catalysts were used. For instance, in work [108] concerned with the addition reactions of CCl4 and CCl3COOEt, reac tion rates increased by 320 times under microwave activation conditions compared with thermal heating. The authors of [109] studied the decomposition of trichloroethylene on a deposited Pt/Ni catalyst. They found that optimum process was characterized by a catalytic layer thickness of 13 m. A decrease or increase in thickness decreased catalyst activity or caused coke accumulation. Waste and renewable raw material processing. Sev eral examples of the use of microwave heating in bio catalysis, in the first place, in the hydrolysis of cellu lose for plant raw material processing were described [110]. Among diverse reactions of side product and waste processing, the pyrolysis of glycerol into synthe sis gas [111] and the oxidation of compounds model ing lignin on the Co(salen)/SBA 15 mesoporous cat alyst [112] can be mentioned. In the last case, the complete conversion of apocynol (model compound) under the action of hydrogen peroxide was obtained in 40 min, whereas the usual heating regime gave conver sion no higher than 60% and only in ~24 h. According to [113], a reactor with microwave activation could be used for acid hydrolysis of cellulose and cellulose wastes with phosphoric acid. A 90% yield of glucose was obtained. A review of the methods for the utilization of wastes with the use of microwave technologies was given in [114, 115]. The processes considered included repro cessing of old tires and plastic wastes and restoration of (removal of harmful contaminants from) soils and underground water. It is, however, noted that there are scaling problems with the passage from laboratory studies to microwave technologies on an industrial scale. The use of microwave methods for the produc tion of biogas with a low content of CO2 and CH4 by processing dead water precipitates was described [116]. Compared with thermal pyrolysis, the micro wave method allows a higher yield (~94%) of synthesis gas (CO + 2) with a lower content of CO2 and CH4 to be obtained. Review [117] considers the use of microwave tech nology in processing of tar sand and oil products, in particular, for the extraction of bitumens, conversion

of heavy residues, the removal of sulfur and nitrogen compounds, and heating tar sand for decreasing the viscosity of bitumens. One more example of useful conversion of wastes is the biotechnological produc tion of hydrogen under microwave processing condi tions [118]. Recently, microwave technologies have begun to be applied for liquefaction of coal [119]. PROSPECTS FOR THE USE OF MICROWAVE RADIATION IN CATALYSIS To summarize, there are good grounds in certain instances for believing that catalytic processes occur under the action of microwave fields differently than under traditional thermal activation with convective or conductive heating. The energy of microwave radia tion quanta is, however, insufficient for directly break ing chemical bonds in reacting molecules. A necessary condition for the action of a microwave field on a pro cess is therefore the possibility of interaction of a solid (a catalyst or its separate components) with this field. Among various mechanisms of the action of micro wave field on solids, several classes of catalysts and processes that offer promise for use in practice can be identified. For instance, the MaxwellWagner inter phase polarization mechanism can likely operate in mixed oxide catalysts (such as extensively studied cat alysts of partial oxidation based on vanadium and molybdenum oxides). For such catalysts, nontrivial and nonthermal effects can most probably be expected; these effects are of obvious interest for the fundamental science of catalysis. Among heterogeneous catalysts, there are also many systems representing massive conductors (met als and activated carbons) and systems containing a deposited active metallic phase. Strictly, the Maxwell Wagner mechanism can also be valid for metallic cat alysts deposited on a carrier inert with respect to microwave fields. Heat exchange between a deposited conducting phase and a carrier transparent to micro wave radiation is of great importance for heterostruc tures of the type metal + carrier transparent to electro magnetic field. If the characteristic time of thermal relaxation is larger than reaction duration, such a sys tem is of great interest for bifunctional catalysis with the separation of functions between a metal and a car rier active in the given case, for instance, a carrier pos sessing acid properties or acting as a donor of active lattice oxygen. The main special feature of heating of conductors is the propagation of electromagnetic field at a near light velocity. This can be of obvious advan tage for processes that require fast and selective cata lyst heating. In these instances, microwave field can be used to solve technological problems. Importantly, clear cut separation between thermal and nonthermal effects is necessary for revealing the real action of a microwave field on catalytic (and, on the whole, chemical) reactions. Even with compara
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES

1691

tively simple reactions, errors can arise if experimental data are analyzed without due care. Nontrivial micro wave effects are largely exclusively thermal, and the reason for a change in reaction path is the ability of various catalyst components to absorb electromag netic field better than their environment. For instance, an analysis performed by Perry for the oxidation of CO on Pt/Al2O3 and Pd/Al2O3 showed [22, 23] that there can only be insignificant differences between catalytic processes performed in a microwave field and under thermal heating condi tions. Because of inaccuracy of temperature measure ments in a strong electromagnetic field, these differ ences can be leveled when experimental conditions are improved. In addition, the formation of superheated regions can be excluded for exothermic reactions on metallic deposited catalysts under the continuous action of a microwave field. Only a small number of researchers present proofs that the observed phenomena have a nonthermal ori gin. Roussy repeatedly mentions that a comparison of activation by microwave fields and thermal heating cannot be based exclusively on measurements and comparison of reaction temperatures because of strong differences in methods for heating and difficul ties of accurate temperature measurements. Other reaction parameters (conversions, selectivities, and the type and distribution of microwave active materi als in a catalyst) also require detailed consideration. Currently, the procedure for selecting an appropri ate catalyst for a heterogeneous gas phase catalytic process that is ideally suitable for activation in micro wave fields is as previously not quite clear. The selec tion is complicated because the catalyst should, first, have substantial catalytic activity and, secondly, pos sess certain electrophysical properties (dielec tric/magnetic properties and conductivity). To obtain a maximum positive effect of microwave radiation on a catalytic system, comparatively inert components strongly interacting with electromagnetic fields are introduced into catalysts. The advantages of microwave technologies can more effectively be used if catalytically and electro magnetically active catalyst components are placed on a substrate transparent to microwave fields. This is done to form superheated regions only in the volume or on the surface of the active catalyst phase, which decreases the temperature of the catalyst as a whole and the gas phase. This offers the possibility of sup pressing side reactions and decreasing energy expendi tures for heating inert materials that do not directly participate in the catalytic process. Apart from the selection of objects of study and the formation of structures providing optimum combina tions of adsorption, catalytic, and electrophysical properties, it is necessary to develop a methodological approach to studies of the influence of microwave acti vation on the processes under study. We should rule out
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

(or minimize) the possibility of incorrect interpreta tion of experimental data. This in the first place relates to temperature measurements. It would be ideal to separately measure the temperature of various solid catalyst components (adsorbent, carrier, and active catalyst phases) and the gas phase. Such procedures do exist. In particular, it is possible in principle to deter mine the temperature of various complex system com ponents by measuring their optical spectra. However in practice, such procedures are difficult to use in the majority of cases. In addition, they require a fairly complex and expensive equipment. For this reason, the development of additional cri teria for revealing specific effects that appear under the action of microwave fields on reacting systems is an important problem. In the first place, this relates to changes in kinetic characteristics and dependences (reaction orders with respect to components and acti vation energies) that are less sensitive to errors in tem perature measurements than a comparison of activities or general transformation rates. Changes in process selectivity with respect to certain products at equal conversions and different methods of reacting system activation are most interesting and informative. It is also necessary to track changes in the state of a solid under the action of microwave fields. In certain instances, the appearance of structures that are not formed in the absence of such action can be expected. If they have (as expected) catalytic properties and/or reactivity different from those observed usually, the appearance of the corresponding nontrivial effects in chemical reactions is possible. Lastly, a special position is held by so called bifunctional catalysis, in which the occurrence of var ious stages of the process as a whole takes place on var ious catalytic system components. A typical example is transformations of hydrocarbons on metallic catalysts deposited on acid carriers. A metallic component (as a rule, a Pt family metal) activates the initial molecule with breaking bonds and/or charge transfer (dehydro genation, hydrogenolysis), whereas carbon skeleton transformations (alkylation and isomerization) occur on acid carrier centers. Simultaneous hydrogena tion/dehydrogenation, cyclization, aromatization, etc. processes can occur on both complex catalyst components in various combinations. Microwave radiation creating temperature gradients in a hetero geneous system whose components differently interact with electromagnetic fields can then cause substantial changes in the general rate and, which is especially important, distribution of reaction products. Technologically, the advantages of the use of microwave activation of heterogeneous gas phase cat alytic reactions compared with thermal activation become obvious when fast heating or cooling and pulsed energy supply are required. The mobile removal of contaminants from air is a suitable area for the use of microwave technologies, because they easily
Vol. 84 No. 10 2010

1692

KUSTOV, SINEV

provide high intensity pulses and solve the cold start problem. Apart from the search for and synthesis of suitable catalysts, the development of new microwave equipment for applications on pilot and industrial scales is necessary. Effects observed in catalysis under the action of microwaves are still not well understood and have been poorly studied, although it is clear that irradiation of microwave absorbing materials (catalysts, carriers, and reaction medium) can cause rapid volume heat ing, effectively remove moisture from solids, and mod ify surface properties. This distinguishes the action of microwaves from traditional thermal treatment, which is used in the preparation of catalysts and in reactions. Solid catalytic materials can be divided into three groups according to their character of interaction with microwave radiation. The first group includes metals whose smooth surface fully reflects microwave rays. Metals then are not heated, because there is almost no microwave radiation loss into their volume. However, if the surface is rough, microwave radiation can cause an arc discharge on it. The second group includes dielectrics transmitting microwave radiation almost unchanged through their volume. These are alumi num and silicon oxides, fused quartz, various glasses, porcelain and faience, polyethylene, polystyrene, and fluoroplastics (Teflon etc.). Lastly, the third group include dielectrics that absorb microwave radiation, which is in particular accompanied by sample heating. In practice, mixtures containing substances that weakly and strongly absorb microwave radiation are often used for microwave heating. By varying the com position of such mixtures, we can control the maxi mum temperature of mixture heating and the compo sition of reaction products. Clearly, the use of microwaves in catalysis will be developed in two directions. The first direction is related to the preparation and preliminary activation of catalysts with the use of microwaves for traditional reactors, where heating is performed by traditional (thermal) methods [120122]. The second direction is the action of microwaves on catalysts and reaction media (if they absorb microwaves) during chemical catalytic reactions [123128]. The literature data considered can be generalized as follows. The use of microwave action on heteroge neous catalysts during their preparation sometimes allows catalysts with a more uniform distribution of particles to be obtained, the preparation of catalysts to be accelerated, and uniform catalyst volume phase heating to be performed. In certain instances, we can obtain catalysts with the required dispersity by varying radiation frequency. With catalysts consisting of sev eral phases, the replacement of traditional with micro wave heating can contribute to the preferable forma tion of separate phases. When microwave activation during a catalytic reaction is used, a decrease in the temperature of reaction beginning, an increase in

activity or selectivity, and a change in the distribution of reaction products are often observed. A comparison of a process performed under traditional conditions with the same process in the presence of a microwave field is, however, possible not always because of diffi culties related to true process temperature measure ments in the latter case. Most likely, we can expect both purely thermal microwave action effects and effects related to the transformation of the active cata lyst component because of the simultaneous action on it of microwaves and reaction medium. REFERENCES
1. Microwave Power Engineering, Vol. 1: Generation, Transmission, Rectification, Ed. by E. Okress (Aca demic, New York, 1968; Mir, Moscow, 1971). 2. B. Ondrushka, Chem. Eng. Technol. 27, 2 (2004). 3. M. L. Levinson, US Patent 3585258 (1965). 4. R. N. Gedye, F. E. Smith, K. Westaway, et al., Tetrahe dron Lett. 27, 279 (1986). 5. R. J. Giguere, T. L. Brays, S. M. Duncan, et al., Tetra hedron Lett. 27, 4945 (1986). 6. J. K. S. Wan, US Patent 4345983 (1982). 7. J. K. S. Wan and J. F. Kriz, US Patent 4545879 (1985). 8. M. Nuechter, B. Ondruschka, W. Bonrath, et al., Green Chem 6, 128 (2004). 9. J. K. S. Wan, Res. Chem. Intermediat. 19, 147 (1993). 10. C. S. Cundy, Collect. Czech. Chem. Commun. 63, 1699 (1998). 11. J. K. S. Wan, K. Wolf, and R. D. Heyding, Catal. Energy. Sci., p. 561 (1984). 12. K. Wolf, H. K. Choi, and J. K. S. Wan, Austral. J. Res. 3, 53 (1986). 13. J. K. S. Wan, M. Tse, and H. Husby, J. Microwave Power Electromagn. Energy 25, 32 (1990). 14. J. K. S. Wan, M. Tse, and M. C. Depew, Res. Chem. Intermediat. 13, 221 (1990). 15. G. Bamwenda, M. C. Depew, and J. K. S. Wan, Res. Chem. Intermediat. 16, 241 (1991). 16. G. Bamwenda, E. Moore, and J. K. S. Wan, Res. Chem. Intermediat. 17, 243 (1992). 17. M. S. Ioffe, S. D. Pollington, and J. K. S. Wan, J. Catal. 151, 349 (1995). 18. G. Bamwenda, M. C. Depew, and J. K. S. Wan, Res. Chem. Intermediat. 19, 553 (1993). 19. T. R. J. Dinesen, M. Tse, M. C. Depew, and J. K. S. Wan, Res. Chem. Intermediat. 15, 113 (1991). 20. K. L. Cameron, M. C. Depew, and J. K. S. Wan, Res. Chem. Intermediat. 16, 57 (1991). 21. M. C. Depew, S. Lem, and J. K. S. Wan, Res. Chem. Intermediat. 16, 213 (1991). 22. J. K. S. Wan and T. A. Koch, Res. Chem. Intermediat. 20, 29 (1994). 23. W. L. Perry, J. D. Katz, D. Rees, et al., J. Catal. 171, 431 (1997). 24. W. L. Perry, D. W. Cooke, and J. D. Katz, Catal. Lett. 47, 1 (1997). 25. J. R. Thomas, Jr., J. Am. Ceram. Soc., p. 397 (1997).
Vol. 84 No. 10 2010

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

MICROWAVE ACTIVATION OF CATALYSTS AND CATALYTIC PROCESSES 26. J. R. Thomas, Jr., Catal. Lett. 49, 137 (1997). 27. X. Zhang, D. O. Hayard, and D. M. P. Mingos, Chem. Commun., p. 975 (1999). 28. Foundations and Industrial Application of Microwave and Radio Frequency Fields, Ed. by G. Roussy and J. A. Pierce (Wiley, New York, 1995). 29. T. T. Meek, J. Mater. Sci. Lett. 6, 638 (1987). 30. K. A. Alberty, Physical Chemistry, 7th ed. (Wiley, New York, 1987), p. 326. 31. W. R. Tinga, Electromagn. Energy Rev. 1, 1 (1988). 32. R. N. Gedye, F. E. Smith, and K. G. Westaway, Can. J. Chem. 17, 66 (1988). 33. S. L. McGill and J. W. Walkiewicz, J. Microwave Power Electromagn. Energy Symp. Summ., p. 175 (1987). 34. A. R. von Hippel, Dielectric Materials and Applications (Massachusetts Inst. Technol., Cambridge, 1954). 35. P. Debye, Polar Molecules. Chemical Catalog (New York, 1929). 36. H. Frohlich, Theory of Dielectrics (Oxford Univ., Lon don, 1958). 37. C. J. F. Bottcher, Theory of Electric Polarization (Elsevier Biomed., Amsterdam, 1952). 38. R. J. Meakins, Trans. Faraday Soc. 51, 953 (1955). 39. C. P. Smyth, Dielectric Behavior and Structure (McGraw Hill, New York, 1955). 40. R. J. Meakins, Trans. Phys. Soc. 52, 320 (1956). 41. K. W. Wagner, Arch. Elektrotech. 2, 371 (1914). 42. W. C. Sun, P. M. Guy, J. H. Rossomando, et al., J. Org. Chem. 53, 4414 (1988). 43. E. G. E. Jahngen, R. R. Lentz, P. S. Pesheck, et al., J. Org. Chem. 55, 3406 (1990). 44. D. M. P. Mingos and D. R. Baghurst, Chem. Soc. Rev. 20, 1 (1991). 45. D. A. C. Stuerga, K. Gonon, and M. Lallemant, Tet rahedron 42, 6229 (1993). 46. D. Stuerga and P. Gaillard, Tetrahedron 52, 5505 (1996). 47. A. Y. Klimov, B. S. Balzhinimaev, L. L. Makarshin, et al., Kinet. Catal. 39, 511 (1998). 48. G. Roussy, J. M. Thiebaut, and M. S. Mdejaram, Catal. Lett. 21, 133 (1993). 49. G. Roussy, L. Seyfried, F. Garin, et al., J. Catal. 148, 281 (1994). 50. G. Roussy, E. Marshal, J. M. Thiebaut, et al., Fuel Process. Tech. 12, 542 (2000). 51. G. Roussy, E. Marchale, J. M. Thiebaut, et al., Mea sure Sci. Technol. 12, 542 (2000). 52. V. S. Arutyunov and O. V. Krylov, Oxidative Methane Transformations (Nauka, Moscow, 1998) [in Russian]. 53. G. Bond, R. B. Moyes, and D. A. Whan, Catal. Today 17, 427 (1993). 54. G. Roussy, J. M. Thiebaut, M. Souiri, et al., Catal. Today 21, 349 (1994). 55. G. Roussy, E. Marchal, J. M. Thiebaut, et al., Fuel Process. Technol. 50, 261 (1997). 56. T. Ito, T. Tashiro, M. Kawasaki, et al., J. Phys. Chem. 95, 4476 (1991).
RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

1693

57. T. Ito, J. Wang, C. Lin, et al., J. Am. Chem. Soc. 107, 5062 (1985). 58. C. L. Chen, P. J. Hong, and S. S. Dai, J. Chem. Soc., Faraday Trans. 91, 1179 (1997). 59. C. L. Chen, P. J. Hong, S. S. Dai, et al., Chin. Chem. Lett. 6, 1003 (1995). 60. C. L. Chen, P. J. Hong, S. S. Dai, et al., React. Kinet. Catal. Lett. 61, 181 (1997). 61. A. S. Bhalla, R. Guo, and R. Roy, Mat. Res. Innovat., 3 (2000). 62. X. J. Bi, P. J. Hong, X. G. Xie, and S. S. Dai, React. Kinet. Catal. Lett. 66, 381 (1999). 63. C. L. Chen, P. J. Hong, S. S. Dai, et al., React. Kinet. Catal. Lett. 61, 175 (1997). 64. P. J. Gellings and H. J. M. Bouwmeester, Catal. Today 58, 1 (2000). 65. S. A. Freemann, J. H. Booske, R. T. Cooper, et al., Ceram. Trans. 36, 123 (1991). 66. J. D. Katz, R. D. Blake, and V. M. Kenkre, Ceram. Trans. 21, 95 (1991). 67. D. J. Driscoll and J. H. Lunsford, J. Phys. Chem. 89, 4415 (1985). 68. D. J. Driscoll, W. Martir, J. X. Wang, et al., J. Am. Chem. Soc. 107, 58 (1985). 69. T. Fang and C. T. Yeh, J. Catal. 69, 227 (1981). 70. J. K. S. Wan, M. Y. Tse, and M. C. Depew, US Patent 5215634 (1993). 71. X. J. Bi, X. G. Xie, A. H. Duan, et al., Chin. Chem. Lett. 9, 775 (1998). 72. X. J. Bi, X. G. Xie, P. J. Hong, et al., React. Kinet. Catal. Lett. 66, 381 (1999). 73. Y. Liu, Y. Lu, P. Liu, et al., Appl. Catal. A: Gen. 170, 207 (1998). 74. Y. Liu, Y. Lu, S. M. Liu, et al., Catal. Today 51, 147 (1999). 75. Y. Kong and C. Y. Cha, Energy Fuels 9, 971 (1995). 76. C. Y. Cha and Y. Kong, Carbon 33, 1141 (1995). 77. C. Y. Cha, US Patent 5246554 (1993). 78. C. Y. Cha, US Patent 5256265 (1993). 79. C. Y. Cha, US Patent 5269892 (1993). 80. Y. F. Chang, A. Sanjurjo, J. G. McCarthy, et al., Catal. Lett. 57, 187 (1999). 81. X. Wang, T. Zhang, C. Xu, et al., Chem. Commun., p. 279 (2000). 82. X. Zhang, D. O. Hayward, C. Lee, et al., Appl. Catal. B: Environmental 33, 137 (2001). 83. H. Will, PhD Thesis (Univ. of Jena, Jena, 2003). 84. T. Thomas, US Patent 5180559 (1993). 85. Z. Yang, J. Zhang, X. Cao, et al., Appl. Catal. B: Envi romental, p. 129 (2001). 86. J. Tang, T. Zhang, L. Ma, et al., J. Catal. 211, 560 (2002). 87. S. L. Suib, CATTECH 2, 75 (1998). 88. S. S. Berdonosov, D. G. Berdonosova, and I. V. Zna menskaya, Khim. Tekhnol., No. 3, 2 (2000). 89. F. J. Berry, L. E. Smart, P. S. S. Prasad, et al., Appl. Catal. A: Gen. 204, 191 (2000).
Vol. 84 No. 10 2010

1694

KUSTOV, SINEV 109. H. Takashima, M. Karches, and Y. Kanno, Appl. Surf. Sci. 254, 2023 (2008). 110. S. Zhu, Y. Wu, Z. Yu, et al., Bioresource Technol. 97, 1964 (2006). 111. Y. Fernandez, A. Arenillas, M. A. Diez, et al., J. Anal. Appl. Pyrolys. 84, 145 (2009). 112. S. K. Badamali, R. Luque, J. H. Clark, et al., Catal. Commun. 10, 1010 (2009). 113. A. Orozco, M. Ahmad, D. Rooney, et al., Proc. Safety Environm. Protect. 85, 446 (2007). 114. T. J. Appleton, R. I. Colder, S. W. Kingman, et al., Appl. Energy 81, 85 (2005). 115. D. A. Jones, T. P. Lelyveld, S. D. Mavrofidis, et al., Resourc. Conservat. Recycl. 34, 75 (2002). 116. A. Dominguez, Y. Fernandez, B. Fidalgo, et al., Chemosphere 70, 397 (2008). 117. S. Mutyala, C. Fairbridge, J. R. J. Pare, et al., Fuel Process. Technol. 91, 127 (2010). 118. L. Guo, X. M. Li, X. Bo, et al., Bioresource Technol. 99, 3651 (2008). 119. T. X. Wang, Z. M. Zong, J. W. Zhang, et al., Fuel 87, 498 (2008). 120. E. Vileno, H. Zhou, Q. Zhang, et al., J. Catal. 187, 285 (1999). 121. A. L. Astakhov, A. Yu. Stakheev, L. M. Kustov, et al., in Proc. 1st EFCATS School on Catalysis, March 27Apr. 1, 2001, Prague, p. 27. 122. A. L. Astakhov, A. Yu. Stakheev, L. M. Kustov, et al., in Proc. of the EuropaCat 5, Limerick, Ireland, Sept. 27, 2001, vol. 5, pp. 161. 123. L. M. Kustov, A. V. Dolgolaptev, A. V. Uvarov, et al., in Proc. of the Europacat 7, Sofia, Sept. 2005, P. OF 10 01. 124. J. Tang, T. Zhang, L. Ma, et al., J. Catal. 211, 560 (2002). 125. P. S. S. Prasad, N. Lingaiah, P. K. Rao, et al., Catal. Lett. 35, 345 (1995). 126. G. Bond, R. B. Moyes, S. D. Pollington, et al., Stud. Surf. Sci. Catal. 75, 1805 (1993). 127. Y. Wada, H. B. Yin, T. Kitamura, et al., Chem. Lett., 632 (2000). 128. S. Ringler, F. Garin, G. Maire, et al., in Proc. of the Europacat 3, Krakow, Poland, Aug. 31Sept. 6, 1997, p. 124.

90. R. Gopinath, K. N. Rao, P. S. S. Prasad, et al., J. Mol. Catal. A: Chem. 181, 215 (2002). 91. L. D. Conde, C. Marun, S. L. Suib, et al., J. Catal. 204, 324 (2001). 92. Y. Wada, H. Kuramoto, T. Sakata, et al., Chem. Lett., p. 607 (1999). 93. S. Ringler, F. Garin, G. Maire, et al., in Proc. of the Europacat 3, Krakow, Poland, Aug. 31Sept. 6, 1997, p. 124. 94. A. Yu. Klimov, B. S. Balzhinimaev, L. L. Makarshin, et al., in Proc. of the Europacat 3, Krakow, Poland, Aug. 31Sept. 6, 1997, p. 189. 95. A. Wali, S. M. Pillai, and S. Satish, React. Kinet. Catal. Lett. 60, 189 (1997). 96. T. L. Lai, C. C. Lee, G. L. Huang, et al., Appl. Catal. B: Environmental 78, 151 (2008). 97. J. K. S. Wan, K. Wolf, and R. D. Heyding, Stud. Surf. Sci. Catal. 19, 561 (1984). 98. J. S. Sung, A. L. Tarasov, L. M. Kustov, et al., Int. J. Hydrogen Energy 33, 4116 (2008). 99. B. Tyagi, B. Shaik, and H. C. Bajaj, Catal. Commun. 11, 114 (2009). 100. A. M. Balu, J. M. Hidalgo, J. M. Campelo, et al., J. Molec. Catal. A: Chem. 293, 17 (2008). 101. B. Nigrovski, U. Zavyalova, P. Scholz, et al., Carbon 46, 1678 (2008). 102. S. Gopalakrishnan, J. Munch, R. Herrmann, et al., Chem. Eng. J. 120, 99 (2006). 103. V. Calvino Casilda, M. O. Guerrero Perez, and M. A. Banares, Appl. Catal. B: Environmental (Accepted Manuscript,) Available online Jan. 18, 2010. 104. V. Calvino Casilda, M. O. Guerrero Perez, and M. A. Banares, Green Chem. 11, 939 (2009). 105. A. de la Hoz, A. Daz Ortiz, and A. Moreno, Chem. Soc. Rev. 34, 164 (2005). 106. M. Kidwai, Pure Appl. Chem. 73, 147 (2001). 107. Microwaves in Organic Synthesis, Ed. by A. Loupy (Wiley VCH, Weinheim, 2006); B. L. Hayes, Micro wave Synthesis: Chemistry at the Speed of Light (CEM Publ., Matthews, 2002); R. S. Varma, Microwave Technology Chemical Synthesis, Applications, in Kirk Othmer Encyclopedia of Chemical Technology (Wiley, New York, 2003). 108. F. Adamek and M. Hajek, Tetrahedron Lett. 33, 2039 (1992).

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A

Vol. 84

No. 10

2010

Das könnte Ihnen auch gefallen