Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Physics

Download as pdf or txt
Download as pdf or txt
You are on page 1of 57

Chapter 5

Physics of Microwave Heating

Abstract Microwave heating generates heat by absorption and loss of energy.


Accordingly, how does a substance turn microwave into heat? How does one
measure the efficiency of microwave heating? These topics require to be under-
stood. As such, this chapter provides some commentaries on these topics.
Furthermore, a method for measuring the heating efficiency of microwaves,
impedance matching, is also explained. There are books out there that presume that
microwave heating refers to heating in which dielectric heating proceeds. This is
not entirely correct, as there are several types of microwave heating phenomena.
This chapter also describes the type of microwave heating through illustrations and
some equations. Furthermore, actual examples and chemical reactions are given;
such topics as the penetration depth of microwaves, the skin effect, the frequency
effect, simulation of electromagnetic waves, and electrothermal and transmission
modes are discussed. The coffee break shows the proper usage of microwave ovens.

 
Keywords Dielectric properties Dielectric constant Dielectric loss Tan d 
 
Permeability Transmission/reflection line method Open-ended coaxial probe
  
method Free-space method Resonant method Impedance Dipole rotation 
 
Conduction loss heating Dielectric loss heating Magnetic loss heating
 
Impedance Dipole rotation Conduction loss heating Penetration depth
  
Skin depth Industrial Scientific and medical (ISM) bands 915-MHz frequency
 
2.45-GHz frequency 5.80-GHz frequency Frequency effect
 
Diels–Alder reaction Synthesis of ionic liquids Nanoparticle synthesis
 
Electromagnetic simulations Thermodynamics simulations Transmission modes

5.1 Dielectric Properties

Microwaves are electromagnetic waves that mimic an alternating current (electric


and magnetic fields). Interactions between microwaves and materials depend upon
the physical properties of the materials; that is, how the alternating electric and
magnetic fields interact with materials. When considering microwave heating, it is

© Springer Nature Singapore Pte Ltd. 2018 87


S. Horikoshi et al., Microwave Chemical and Materials Processing,
https://doi.org/10.1007/978-981-10-6466-1_5
88 5 Physics of Microwave Heating

necessary to understand the electrical properties of the material since the alternating
electric field interacts with the electrons of the material. Therefore, the electrical
nature of the substance is related to its interaction with microwaves. Classification
of classes of materials on the basis of their electrical conductivity is reported in
Table 5.1. Materials that possess high electrical conductivity are conductors and
examples are metallic in nature. Materials that display lesser electrical conductivity
are the semiconductors (e.g., silicon). Materials that do not exhibit electrical con-
ductivity are referred to as dielectric or insulating materials; an example is quartz,
one of many insulating materials.
A material is classified as dielectric if it has the ability to store energy upon
application of an external electric field. If a direct current (DC) voltage source is
placed across a parallel plate capacitor (Fig. 5.1), more charge is stored when a
dielectric material is located between the electrodes. The dielectric material
increases the storage capacity of the capacitor by neutralizing the charges at the
electrodes, which ordinarily would contribute to the external field. The capacitance
with the dielectric material is related to its dielectric constant.
In Fig. 5.1, C and C0 are, respectively, the capacitance with and without the
dielectric; j′ = er′ is the real dielectric constant or permittivity; A and t denote,
respectively, the area of the capacitor plates and the distance between them
(Fig. 5.1). The capacitance of the dielectric material is related to the dielectric
constant as indicated in the equations reported in Fig. 5.1. If an AC sinusoidal
voltage source V is placed across the same capacitor (Fig. 5.2), the resulting current
will be made up of a charging current Ic and a loss current Il that is related to the
dielectric constant. The losses in the material can be represented as a conductance
(G) in parallel with a capacitor (C).

Table 5.1 Electrical resistivity of materials (the reciprocal of conductivity) and carrier density at
room temperature
Materials Electrical resistivity q/X cm Carrier density/cm−3
−6
Metal 10 >1022
Semiconductor 10−6–109 1013–1017
Dielectric (insulating) 1014 <1013

Fig. 5.1 Parallel plate


capacitor, DC case.
Reproduced from Ref. [1]
5.1 Dielectric Properties 89

When an electric field interacts with an insulator, there is no current flow in the
material. Electrons and ions in the material move to opposite sides under the electric
field; this displacement of electrons and ions generates an electric dipole, a phe-
nomenon referred to as dielectric polarization. The electric flux density inside the
material, induced as a result of the dielectric polarization, is expressed by the
product of the external electric field and the dielectric constant.
The complex dielectric constant j consists of a real part j′ that represents the
storage and an imaginary part j″ that represents the loss. The following notations
are often used interchangeably to denote the complex dielectric constant (Eq. 5.1):

j ¼ j ¼ er ¼ er ð5:1Þ

From the point of view of electromagnetic theory, the definition of electric


displacement (electric flux density: Df) is given by Eq. 5.2:

Df ¼ eE ð5:2Þ

From Eq. 5.2, the dielectric constant is defined as the ratio between the electric
flux density generated in the material and the electric field when the electric field
interacts with the material.
Under microwave irradiation, the external electric field in Eq. 5.2 changes with
time. Although an electromagnetic wave is represented by trigonometric functions,
for mathematical convenience it can be expressed by an exponential function and
complex numbers (Euler’s formula: ejh = cos h + j sin h). As described above,
under an alternating electromagnetic field, the dielectric constant becomes a com-
plex number (Eq. 5.3):

~ D jh D
D
e¼ ¼ e ¼ ðcos h þ j sin hÞ ¼ e0  je00 ð5:3Þ
 E
E E

In Eq. 5.3, e′ is the dielectric constant (the real part of the dielectric constant
when an alternating electromagnetic field is applied to the material) and e″ is the
dielectric loss (the imaginary part of the dielectric constant when an alternating
electromagnetic field is applied to the material). The dielectric constant, e′, repre-
sents the degree of polarization in the material under an applied electric field: the
effect of electrical energy. If polarization is zero, the value of e′ is zero. The

Fig. 5.2 Parallel plate capacitor, AC case. Reproduced from Ref. [1]
90 5 Physics of Microwave Heating

dielectric loss, e″, refers to the degree of energy loss in the material under an applied
electric field, the loss being converted to heat. If the time difference is zero, the
value of e″ is also zero. The dielectric constant (j) is equivalent to the relative
permittivity (er) or the absolute permittivity (e) relative to the permittivity of free
space (eo: eo  1/36p  10−9 s). The real part of permittivity (er′) is a measure of
how much energy from an external electric field is stored in a material. The
imaginary part of permittivity (er″) is called the loss factor and is a measure of how
dissipative or lossy a material is to an external electric field; that is, how much
energy is lost (converted to heat) in the material. The imaginary part of permittivity,
er″, is always greater than zero and is usually much smaller than er′. The loss factor
includes the effects of both dielectric loss and conductivity.
The dielectric polarization is time-dependent and, if accurately measured, it
changes slightly with time. This time lag can be represented geometrically. If the
degree and the time lag of the dielectric polarization are represented by a real
component and an imaginary component, respectively, the vector diagram is then
given as illustrated in Fig. 5.3. The phase of the real and imaginary components is
shifted 90°. The vector sum forms an angle d to the real axis e′. The relative loss of
materials can be represented by tan d, which is defined by the ratio of the dielectric
loss and the dielectric constant. When complex permittivity is drawn as a simple
vector diagram, the real and imaginary components are 90° out of phase. The vector
sum forms an angle d (the “loss tangent”) with the real axis er′. The relative
lossiness of a material is the ratio of the energy lost to the energy stored.
The loss tangent, or tan d, represents the ratio of the imaginary part of the
dielectric constant to the real part. In the equation of Fig. 5.3, D denotes the
dissipation factor and Q the quality factor. The loss tangent is called tan delta (i.e.,
tan d), or loss tangent, or dissipation factor. Sometimes, the terms quality factor or
Q-factor are used with respect to an electronic microwave material; this factor is the
reciprocal of the loss tangent. For very low-loss materials, since tan d  d, the loss
tangent can be expressed in angle units as milliradians or microradians. This is
represented as the dielectric loss tangent. We can estimate the extent to which
materials are heated by microwave radiation by considering the dielectric loss and
the relative loss as indicators. Since the dielectric constant takes a different value
depending on the material, we use a value of the ratio against the permittivity of
vacuum (eo = 8.85  1012 F m−1); this value is referred to as the relative dielectric
constant (er′). Similarly, for dielectric loss, the ratio of the permittivity of a material

Fig. 5.3 Loss tangent vector diagram. Reproduced from Ref. [1]
5.1 Dielectric Properties 91

to that of a vacuum is represented as the relative dielectric loss: er″. Hence, the loss
tangent is given by the ratios expressed in Eq. 5.4:

e00 e00r
tan d ¼ ¼ 0 ð5:4Þ
e0 er

Figure 5.4 shows the relationship between the relative dielectric constant and the
dielectric loss of powdered materials and several liquids. The dielectric constants of
the powders were measured at a microwave frequency of 6 GHz using the pertur-
bation method (Fig. 5.4a) [2]. Carbon has both a large relative dielectric constant and
dielectric loss factor. In an oxide consisting of molecules of the same type, the
dielectric loss factor is largely different depending on the valence. For example, in iron
oxides, the dielectric loss factor of Fe2O3 is less than 0.01 at the same level as SiO2,
but the dielectric loss factors of FeO and Fe3O4 are greater than 1, approximately two
orders of magnitude larger. Figure 5.4b shows the dielectric constant and dielectric
loss of liquids [3]. Polar solvents possess greater dielectric loss factors than nonpolar
solvents. In particular, water and methanol have a dielectric loss greater than 10. On
the other hand, organic solvents and gasoline have smaller dielectric loss factors, and
thus microwave heating of such solvents tends to be difficult.
Note that the dielectric polarization consists of (i) electronic polarization,
(ii) atomic polarization, (iii) orientation polarization, and (iv) ionic polarization;
each of these is determined by the frequency.
In the microwave frequency band, the peak of the complex dielectric constant is
in the region of orientation polarization. Further, as inferred from Fig. 3.7 of
Chap. 3, in the microwave band, atomic polarization and electron polarization can
respond quickly to changes in the microwave’s electric field. A material may have
several dielectric mechanisms for polarization effects that contribute to its overall
permittivity (Fig. 3.7 of Chap. 3). Electric charge carriers in a dielectric material are

Fig. 5.4 Complex permittivity and loss factor of various powers and liquids in the microwave
region. (a) powder materials (measured at 6 GHz) and (b) liquids (reference at 2.45 GHz).
Reproduced from Ref. [2]. Copyright by the GCMEA Committee
92 5 Physics of Microwave Heating

so arranged that they can be displaced by an electric field. The charges become
polarized to compensate for the electric field, so that the positive and negative
charges move in opposite directions. At the microscopic level, several dielectric
mechanisms can contribute to the dielectric behavior. Dipole orientation and ionic
conduction interact strongly at microwave frequencies. Water molecules, for
example, possess permanent dipoles, which rotate to follow an alternating electric
field, making water quite lossy—this is why water-based food heats well in a
microwave oven. Atomic and electronic mechanisms are relatively weak and
usually constant over the microwave region. Each dielectric mechanism has a
characteristic cutoff frequency. As frequency increases, the slow mechanisms drop
out, in turn, leaving the faster ones to contribute to e′. The loss factor (er″) will peak
correspondingly at each critical frequency. The magnitude and cutoff frequencies of
each mechanism are unique to each material. Water has a strong dipolar effect at
low frequencies—but its dielectric constant rolls off dramatically around 22 GHz.
On the other hand, PTFE has no dipolar mechanisms and its permittivity is
remarkably constant well into the millimeter wave region.
A resonant effect is usually associated with electronic or atomic polarization,
whereas a relaxation effect is usually associated with orientation polarization, which
depends on frequency because it changes with the nature of the microwave radi-
ation. Therefore, the electric flux density of an insulator at a given time t is sepa-
rated into a term Dp(t) that can follow and a term Dd(t) that can delay a change of
the microwave electric field (Eqs. 5.5 and 5.6):

Dp ðtÞ ¼ e0 er ð1ÞEðtÞ ð5:5Þ

Zt
D d ðt Þ ¼ E ðuÞf ðt  uÞdu ð5:6Þ
0

where f(t − u) is a function of the dielectric constant. This means that the micro-
wave electric field at an earlier time u affects the dielectric constant at current time
t (that function is referred to as the after-effect function). Thus, the electric flux
density D(t) is the sum of a term Dp(t) that can follow and a term Dd(t) that can
delay a change of the microwave electric field (Eq. 5.7),

DðtÞ ¼ Dp ðtÞ þ Dd ðtÞ ð5:7Þ

If we now let e be e0e*r , where e*r is the complex relative dielectric constant
(Eqs. 5.8–5.11),

Zt
e0 er ðxÞE ðtÞ ¼ e0 er ð1ÞE ðtÞ þ EðuÞf ðt  uÞdu ð5:8Þ
0
5.1 Dielectric Properties 93
Zt
 
e0 er ðxÞ  er ð1Þ EðtÞ ¼ E ðuÞf ðt  uÞdu ð5:9Þ
0

Zt
 
e0 er ðxÞ  e r ð 1Þ ¼ ejxðutÞ f ðt  uÞdu ð5:10Þ
0

Z0 Zt
 
e0 er ðxÞ  e r ð 1Þ ¼  e jxx
f ð xÞdx ¼ ejxx f ð xÞdx ð5:11Þ
t 0

and if we let the current time be far ahead (i.e., infinity) and consider the integral
range x = t − u = ∞ (Eq. 5.12),
Z1
 
e0 er ðxÞ  er ð1Þ ¼¼ ejxx f ð xÞdx ð5:12Þ
0

and if the after-effect function is expressed by Eq. 5.13,

ex=s
f ð xÞ ¼ e0 fer ð0Þ  er ð1Þg ð5:13Þ
s

then we obtain a complex relative dielectric constant given by Eqs. 5.14 and 5.15:
Z1
 
e0 er ðxÞ  e r ð 1Þ ¼ ejxx f ð xÞdx
0
Z1  
ex=s
¼ ejxx e0 fer ð0Þ  er ð1Þg
s
0
Z1
exð1 þ jxsÞ=s
¼ e0 fer ð0Þ  er ð1Þg dx ð5:14Þ
s
0
 1
1 s xð1 þ jxsÞ=s
¼ e0 fer ð0Þ  er ð1Þg  e
s 1 þ jxs s
e0 fer ð0Þ  er ð1Þg
¼ ½ð0Þ  ð1Þ
1 þ jxs
e0 fer ð0Þ  er ð1Þg
¼
1 þ jxs

e r ð 0Þ  e r ð 1 Þ
er ðxÞ ¼ er ð1Þ þ ð5:15Þ
1 þ jxs
94 5 Physics of Microwave Heating

Equation 5.15 shows that the complex relative dielectric constant is a function of
the frequency—it is known as Debye’s relaxation equation. The relationship
between microwave heating and the complex dielectric constant is described next in
Sect. 5.2. In addition, the dependence of the dielectric constant on temperature will
be described later (see below, Sect. 5.3.4).

5.2 Permeability

Permeability (µ) describes the interaction of a material with a magnetic field.


A similar analysis can be performed for permeability using an inductor with
resistance to represent core losses in a magnetic material (Fig. 5.5). If a DC current
source is placed across an inductor, the inductance with the core material can be
related to the permeability. In the equations of Fig. 5.5, L is the inductance with the
material, Lo is the free-space inductance of the coil, and µ′ is the real permeability.
If an AC sinusoidal current source is placed across the same inductor, the resulting
voltage will be made up of an induced voltage and a loss voltage that is related to
permeability. The core loss can be represented by a resistance (R) in series with an
inductor (L). The complex permeability (µ* or µ) consists of a real part (µ′) rep-
resenting the energy storage term, and an imaginary part (µ″) representing the
energy loss term. The relative permittivity µr is the permittivity relative to free
space (Eq. 5.16):
m
lr ¼ ¼ lr  jl00r ð5:16Þ
m0

Note that µo = 4p  10−7 H m−1 is the free-space permeability.


Some materials, such as iron (ferrites), cobalt, nickel, and their alloys have
appreciable magnetic properties; however, many materials are non-magnetic mak-
ing the permeability very close to the permeability of free space (µr = 1). All
materials, on the other hand, have dielectric properties, so the focus of this dis-
cussion is mostly on permittivity measurements.

Fig. 5.5 Image of an


inductor. Reproduced from
Ref. [1]
5.3 Measurement of Electric and Magnetic Permeabilities 95

5.3 Measurement of Electric and Magnetic Permeabilities

The heating efficiency of a sample exposed to microwave radiation depends upon


the electric and magnetic permeabilities of the sample. Many researchers have
reported the electric permeability and magnetic permeability of pure materials;
however, when using a mixture in a chemical reaction it is best to measure these in
the actual sample. It would not be possible to assess the microwave heating effi-
ciency without first measuring the permeability. For example, impurities (e.g., small
quantities of some ionic species) present in a polar solvent would significantly
improve the microwave heating efficiency.
There exist many methods for measuring the complex permittivity and perme-
ability of samples, each method being limited to, among others, specific microwave
frequencies, materials, and applications. Accordingly, this section introduces and
describes the four methods used to measure the permeability. Technical books may
also be consulted for more details; the research paper by Venkatesh and Raghavan
[4] is also worth consulting as it provides an overview of the various techniques.
Heat radiated to the atmosphere from a microwave-heated sample may be large
which will affect the microwave heating efficiency of the sample. Even if the
heating efficiency were estimated by calculating the electric and magnetic perme-
abilities, a judicious researcher should always determine the difference between the
microwave heating efficiency and the radiated heat to the atmosphere of an actual
sample.
Table 5.2 identifies examples of classes of materials and the dielectric properties
determined using various measurement methods. These methods are best used at
some specified frequency, and at some given degree of microwave absorption
depending upon the nature of the sample (liquid, bulk, or powder) and on electrical
characteristics.
The Transmission/Reflection line method is a popular broadband measurement
technique in which only the fundamental waveguide modes (TEM mode in coaxial
line and TE mode in waveguides) are assumed to propagate. The Open-ended
coaxial probe method is a nondestructive technique that assumes only the TEM or

Table 5.2 Comparison between the measurement methods


Measurement techniques Features Dielectric
properties
Transmission/reflection Low-loss samples, solids er, µr
line method
Open-ended coaxial Liquids, biological specimens, semi-solids, high er
probe method loss samples
Free-space method High-temperature samples, large flat solids, er, µr
gases, hot liquids
Resonant method Rod-shaped solid materials, liquids, low-loss er, µr
(Cavity) samples, small samples
96 5 Physics of Microwave Heating

TE modes propagate. The Resonant method provides high accuracies and assumes
TE or TM propagation modes. The Free-space method is for broadband applica-
tions and assumes that only TEM modes propagate.

5.3.1 The Transmission/Reflection Line Method

The transmission/reflection line method involves placing a sample in a section of


the waveguide or coaxial line followed by measuring the two complex parts of the
scattering parameters using a vector network analyzer [5]. This vector network
analyzer is an instrument that measures the network parameters of electrical net-
works. At present, network analyzers commonly measure scattering parameters
(that is, the elements of a scattering matrix, or S-matrix) because reflection and
transmission of electrical networks are easy to measure at high frequencies.
Network analyzers are used mostly at high frequencies; operating frequencies may
range from 5 Hz to 1.05 THz. Scattering parameters (i.e., S-parameters) describe
the electrical behavior of linear electrical networks when undergoing various
steady-state stimuli by electrical signals. This method is used for measurements of
high loss materials. It can be classified into small groups of the transmission line
method, an S-parameter method, and a free-space method.
The relevant scattering parameters relate closely to the complex permittivity and
permeability of the material. The conversion of S-parameters to complex dielectric
parameters is calculated using a computer program for solving the relevant equa-
tions. In many cases, the method requires initial preparation of the sample (e.g.,
machining) so that the sample fits tightly into the waveguide or the coaxial line.
Calibration in transmission line measurements uses various terminations that pro-
duce different resonant behaviors in the transmission line (see Fig. 5.6) [6]. For
good dielectric measurements, the maximum electric field is required, which can be
achieved by open circuit or other capacitive termination, whereas calibration in
coaxial line measurements can be made using a short circuit, an open circuit, or a
matched load termination. Whatever the method used, it provides for the mea-
surement of the permittivity and permeability of dielectric materials.

5.3.2 The Open-Ended Coaxial Probe Method

The open-ended coaxial probe method has been used for years as a nondestructive
testing procedure and is the most often used procedure in microwave chemistry.
With this equipment and a probe, dielectric factors of the solution or of the sample
can be measured, regardless of whether or not a researcher has the appropriate
technical knowledge of the system. Measurements are not unlike measuring the pH
of a solution with a pH meter. The coaxial probe is inexpensive and the mea-
surement method is relatively simple. It is possible to measure the permittivity of a
5.3 Measurement of Electric and Magnetic Permeabilities 97

Fig. 5.6 Image illustrating


the measurement of
permeability using the
transmission line method.
Reproduced from Ref. [6].
Copyright 2013 by
Wiley-VCH

sample and its dissipation loss with temperature changes, because sample tem-
peratures from −40 to +200 °C can be measured with the probe for low- and
high-temperature systems. For example, the three measuring probes are (i) the
high-temperature probe (Fig. 5.7a), (ii) the slim form probe (Fig. 5.7b), and (iii) the
performance probe (Fig. 5.7c) [1].
In this open-ended coaxial probe method, the probe is pressed against a solid
specimen or else is immersed into a liquid sample, followed by measuring the
reflection coefficient subsequently used to determine the permittivity. Note that for
some measurements, it may not be possible to cut or otherwise modify the sample

Fig. 5.7 Types of probes (a) high-temperature probe, (b) slim probe, and (c) performance probe;
(d) images of measurement of samples using an open coaxial probe. Reproduced from Ref. [1]
98 5 Physics of Microwave Heating

Fig. 5.8 Images of types of


coaxial probes

of a material for measurement. This is especially important in the case of per-


forming in vivo measurements on biological specimens because the material
characteristics may change. In this case, the probe may be placed in close contact
with the sample without causing any changes in the material’s characteristics.
The reflection coefficient is measured using a vector network analyzer with a
probe system that is first calibrated so that reflection coefficient measurements are
referenced to the probe’s aperture plane. This can be done in two ways, one of
which uses reference liquids for direct calibration at the open end of the probe—a
very direct and simple procedure. Uncertainties in the measurement are due to
uncertainties in the characterization of the reference liquids, and to the selection of
reference liquids as calibration standards. The reference liquid is used as a cali-
bration standard and must be a liquid with known dielectric properties: for example,
water, a saline solution, or methanol are usually selected as reference liquids.
To perform an actual measurement, the end of the coaxial probe is immersed in
the liquid or in the case of solids is pressed tightly against the surface of the solid
sample. Dissipation loss is measured for high loss solids and liquid samples. The
three kinds of maser (microwave amplification by stimulated emission of radiation)
coaxial probes are illustrated in Fig. 5.8. Measurement and cleaning after the
measurement of a solid sample are simple for the type of maser displayed in
Fig. 5.8a. The types shown in Fig. 5.8b and 5.8c are used for liquid samples for
which measurement sensitivity is much greater than for the type depicted in
Fig. 5.8a.

5.3.3 The Free-Space Method

The free-space method uses antennas to focus the microwave energy at or through a
slab or sheet of the material (Fig. 5.9) [1]. This method is a noncontact method that
can be applied to materials to be tested under high temperatures and in hostile
environments. The measurement requires the materials to be large and flat. As a
non-contact method, the free-space measuring technique does not require the
sample to be inserted in the waveguide and resonator. Time-Domain Gating can
cancel the influence of secondary reflection, which is a problem of the transmission/
reflection line method. The precision of the free-space method depends upon the
5.3 Measurement of Electric and Magnetic Permeabilities 99

Fig. 5.9 Setup of measurement of a sample using the free-space method. Reproduced from
Ref. [1]

environment. For example, when the room temperature changes during the mea-
surement, the measured value is changed; when the cable connected to a network
analyzer is moved, the measured value is changed. Hence, control of these and
other factors are required to eliminate measurement errors.

5.3.4 The Resonant Method

Resonant measurements are the most accurate procedures for obtaining the per-
mittivity and permeability of samples. This technique measures a small sample
using a microwave resonator. The dielectric loss and the magnetic permeability can
be measured by the changes in the resonant frequency and the quality factor (Q-
factor) of the cavity resonator shown in Fig. 5.10. (Sect. 5.4.2). It is particularly
well suited for measurements of low dissipation loss materials. However, it is not
possible to measure large samples because of changes they may cause in the res-
onant conditions in the cavity. There are limitations on the frequencies and loss
characteristics of the materials that can be measured with this method. Accordingly,
there are many types of resonant methods available such as re-entrant cavities, split
cylinder resonators, cavity resonators, and Fabry–Perot resonators among others.
Two types of resonant measurements are commonly used: (1) perturbation methods
are suitable for all permittivity measurements, magnetic materials, and
medium-to-high loss materials; and (2) the low-loss measurement method is
applicable for low-loss materials using larger samples. However, the perturbation
method is more popular, especially using the TM cavity geometry as those shown in
Fig. 5.10a, b. A cavity resonator method is the method with which the electric
permeability can be determined most precisely for a solid sample, more so than a
liquid. The formula of the material constant is easily obtained when the perturbation
method is used for a minute sample inserted in a cavity resonator. Thus, a cavity
resonator method can separate the magnetic permeability from the permittivity
using a field distribution in the cavity resonator and changing the location of the
sample insert.
100 5 Physics of Microwave Heating

Fig. 5.10 Measurement of slim sample using cavity resonator for (a) electric permeability setup
and (b) magnetic permeability

5.4 Adjustment of the Impedance

5.4.1 What Is Meant by Impedance in the Present Context?

It is necessary to understand what impedance is in relation to the microwave heating


efficiency of a sample located in an applicator. Impedance is easily imagined when
it is considered to be the resistance to an alternating current (AC). Simply stated, the
characteristic impedance (Z) by microwave heating is the ratio of the alternating
electric field to the alternating magnetic field (Eq. 5.17) as shown by the electric
permeability and magnetic permittivity (or conductivity) in Maxwell’s equations:
rffiffiffi
E l
Z¼ ¼ ð5:17Þ
H e

Hence, impedance is the electrical characteristic that can be determined from the
magnetic permeability and the permittivity; it can be defined for gas and vacuum.
The impedance in vacuum is referred to as the “impedance of free space”, and is ca.
120 pX (i.e., ca. 377 X). When microwaves are transmitted through a substance
different in impedance (including vacuum and gas) a fraction of the microwaves is
reflected at the boundary between each substance in the applicator, while the other
fraction is absorbed by the substance (i.e., penetrates into the substance). The
behavior for reflection, absorption, or penetration of microwaves depends on the
microwave irradiation angle relative to the surface of the material and its separation
from the source (e.g., sample and reactor); the characteristic impedance also
changes as the material also changes when heated (see Fig. 5.11).
5.4 Adjustment of the Impedance 101

Fig. 5.11 Image of reflection


and absorption (or
penetration) of microwaves at
the phase boundary of the
material different in
impedance

5.4.2 Impedance in Equipment and Sample

If the microwave radiation emitted from a generator does not easily enter into an
applicator, the applied microwave energy cannot be used for heating. Such phe-
nomenon between the sample and the air can occur in microwave heating situations.
Accordingly, if the impedance does not match, the heating efficiency is small
because a large fraction of the microwaves is reflected. Clearly, it is necessary for
the characteristic impedance of the system to match the impedance of the sample so
as to improve the microwave heating efficiency (Fig. 5.12). For example, in a
microwave-assisted organic synthesis, the formation of the desired product causes
the permittivity of the initial sample to be changed; even if impedance matching
occurred at first, it may become invalid during the process. In addition, even though
the electric permeability and magnetic permittivity may be the same for the sub-
stances, sometimes the impedance may be different. For instance, when applying

Fig. 5.12 Image illustrating the possible adjustments of impedance A and impedance B
102 5 Physics of Microwave Heating

Fig. 5.13 Image illustrating the non-adjustment (left) and adjustment (right) of the voltage
standing wave ratio (VSWR) after impedance adjustment in a microwave equipment

microwaves using a waveguide, the size and shape of the waveguide influence the
distribution of the inner electric and magnetic fields inside the waveguide, which
causes the impedance to also change significantly according to Eq. 5.17. To resolve
this problem, a tuner is installed to adjust the impedance between the generator and
the applicator (see tuner in Sect. 6.3.5). It is also possible to change the impedance
by changing the shape of the applicator (see short plunger in Sect. 6.3.5).
The impedance can also be adjusted by connecting a tuner, an iris, and a short
plunger to the equipment (Sect. 6.3.6). Generally, the adjustment of the impedance
in the microwave equipment is done after installing the sample in the microwave
equipment (Fig. 5.13). A vector network analyzer is used to provide the microwave
radiation instead of a microwave generator. First, the three-stub tuner and the short
plunger are opened, which is then followed by adjusting the tuner and the plunger
while monitoring the resonance condition. The voltage standing wave ratio
(VSWR) is determined by a vector network analyzer, and the frequency is scanned
so that a peak may become sharp at a given microwave frequency. The standing
wave ratio (SWR) expresses the ratio of the standing wave between the incident
microwave and the reflected microwave, with the VSWR being the standard
voltage.

5.5 Microwave Heating Mechanism

5.5.1 Phenomena of Dipole Rotation on Application


of Microwaves

A molecule is formed when atoms combine by sharing one or more of their elec-
trons. This rearrangement of electrons may cause an imbalance in charge distri-
bution, thereby creating a permanent electric dipole moment. These moments are
oriented in a random manner in the absence of an electric field so that no polar-
ization exists. When microwave energy is applied to this state, the electric field
5.5 Microwave Heating Mechanism 103

Fig. 5.14 Image of dipole


rotation for a polar molecule
in an electric field (E-field)

influences the electric dipole (Fig. 5.14). The microwaves’ electric field (E-field)
creates a torque (N) on the electric dipole, and the dipole will consequently rotate to
align itself with the electric field, thus causing orientation polarization to occur [7].
As the field changes direction, the torque also changes direction, so that the ori-
entation polarization changes with the changing direction of the microwave electric
field. There is a time delay between the frequency of the microwave alternating
electric field and the electric dipoles continual rotational motion, as it tries to keep
up with the field’s changing polarity. At the general frequency of 2.45-GHz, the
microwaves vibrate 2,450,000,000 times per second, such that a molecular cluster
cannot follow this vibration through the power chain with the surrounding mole-
cules. This delay changes to heat energy (random kinetic energy) as loss of the
electromagnetic wave energy. The friction accompanying the orientation of the
dipole will contribute to dielectric losses. The dipole rotation causes a variation in
both er′ (the relative dielectric constant) and er″ (the relative dielectric loss) at the
relaxation frequency, which usually occurs in the microwave region. Water is an
example of a substance that exhibits a strong orientation polarization.
Electronic polarization occurs in neutral atoms when an electric field displaces
the nucleus with respect to the electrons that surround it. Atomic polarization occurs
when adjacent positive and negative ions stretch under an applied electric field. For
many dry solids, these are the dominant polarization mechanisms at microwave
frequencies, although the actual resonance occurs at a much higher frequency. In
the infrared and visible light regions, the inertia of the orbiting electrons must be
taken into account. Atoms can be modeled as oscillators with a damping effect
similar to a mechanical spring and mass system. The amplitude of the oscillations
will be small for any frequency other than the resonant frequency. Far below
resonance, the electronic and atomic mechanisms contribute only a small constant
amount to er′ and are almost lossless. The resonant frequency is identified by a
resonant response in er′ and a peak of maximum absorption in er″. Above resonance,
the contribution from these mechanisms disappears.
A solid with a partial electric charge (dielectric) is also an insulator. The slight
distortion of the atomic positions (lattice points) in a lattice, and the strain of the lattice
of a crystalline solid cannot follow the time changes of the microwave electric field.
Microwave heating of a solid occurs via these phenomena [8]. On the other hand,
heating of solid substances possessing a magnetic dipole moment occurs by the
microwaves’ magnetic field component. The heating process is the same phenomenon
as by the microwaves’ electric field [9]. Generation of heat by magnetic loss heating is
expected only in magnetic (solid) materials. Joule heating progresses by the
104 5 Physics of Microwave Heating

interaction of an electric or a magnetic field with solid particles of activated carbon and
with those that possess conductivity-like metallic properties [10].

5.5.2 Relationship of Microwave Heating Behavior


with the Materials’ Physical Properties

The microwave heating behavior depends on the nature of the material and the
microwave irradiation conditions. The electrical conductivity (r), permittivity (e),
and permeability (l) are three physical properties of materials with respect to
microwaves and depend upon the physical properties of the materials (dipole
moment, mass, etc.) as well as the state of matter (intermolecular force, boundary
conditions, etc.). The intensity of the electric and magnetic fields and the frequency
are examples of microwave conditions.
Corresponding to the above three physical properties, there are three types of
heating phenomena caused by microwave irradiation: (i) conduction loss heating,
(ii) dielectric heating, and (iii) magnetic loss heating. Accordingly, the thermal
energy P produced per unit volume originating from microwave radiation is given
by Eq. 5.18 [11]:

1
P ¼ rjE j2 þ pf e0 e00r jEj2 þ pf l0 l00r jH j2 ð5:18Þ
2

where |E| and |H| denote the strength of the microwaves’ electric and magnetic
fields, respectively; f is the frequency of the microwaves; eo is the dielectric constant
in vacuum; er″ is the relative dielectric loss factor; lo is the magnetic permeability
of vacuum; and lr″ is the relative magnetic loss. The first term in Eq. 5.18
expresses conduction loss heating; the second term denotes dielectric loss heating,
whereas magnetic loss heating is given by the third term.

5.5.3 Conduction Loss Heating (Eddy Current Loss


and Joule Loss)

If a current is generated in the materials, Joule loss occurs. In this case, the rela-
tionship between the electric field E, the current density i, and the conductivity r is
given by Eq. 5.19:

i ¼ rE ¼ rE0 sin xt ð5:19Þ

Defining the microwave frequency as f, the energy loss w in the material during
one cycle of microwave (1/f = 2p/x) is expressed by Eq. 5.20:
5.5 Microwave Heating Mechanism 105

Z
2p=x Z
2p=x Z
2p=x

w¼ iEdt ¼ rE dt ¼
2
rE02 sin2 xtdt
0 0 0
ð5:20Þ
Z
2p=x  2p=x
1  cos 2xt 1 1 p
¼ rE02 dt ¼ rE02 t  sin 2xt ¼ rE02
2 2 4x 0 x
0

Since the energy loss is repeated f times (=x/2p) per second, the energy W that
the material receives from the electric field per unit time and unit volume is given
by Eq. 5.21:

x p 2 1 2
W¼ rE ¼ rE ð5:21Þ
2p x 0 2 0

Clearly, Eq. 5.21 does not include x, which relates to the microwave frequency.
The hydroxide ion is a typical ionic species with both ionic and dipolar char-
acteristics. The conductive loss effect can become larger than dipolar relaxation in
solutions that contain large quantities of ionic salts. Conduction losses tend to be
slight at ambient temperatures in the case of solids but can change substantially
with increase in temperature. A typical example is alumina (Al2O3) whose dielectric
loss is negligibly small (*10−3) at ambient temperature but can reach fusion levels,
i.e., high dielectric loss, in very short times (minutes) in a microwave cavity. This
effect originates from a strong increase of conduction losses associated with the
thermal activation of electrons, which migrate from the oxygen’s 2p valence band
to the aluminum’s 3s3p conduction band. Moreover, conduction losses in solids are
usually enhanced by defects in materials. Defects sharply reduce the energy needed
to generate electrons and holes in the conduction and valence bands, respectively.
In the particular case of carbon black powder, conduction losses are rather high, so
much so that this material is often used as an additive to induce losses, i.e., heat,
within solids whose dielectric losses are too small.
Microwave heating of a metallic sample concerns the second and third terms of
Eq. 5.18. More information and examples of metal heating are described in
Sect. 9.3.

5.5.4 Dielectric Heating and Magnetic Loss Heating—An


Introduction

When high-frequency electric and magnetic fields are applied to a material, the
charge distribution within the dipole moment changes (liquid sample), and the
material may then shift (solid sample) and/or deflection of spin may occur as
portrayed in Fig. 5.15.
106 5 Physics of Microwave Heating

Fig. 5.15 Images of (a) dipole moment changes and (b) deflection of spin

Loss behavior depends upon the change of the alternating microwave electro-
magnetic field that is very fast compared to direct current, in which case the change
of charge distribution (the dipole moment and/or the deflection of the spin) can
(i) follow, or (ii) follow with delays, or (iii) cannot follow at all. For the first two
cases, the response coefficient is just the dielectric constant and the magnetic per-
meability. If the charge distribution and the deflection of the spin follow without
delay in an alternating electromagnetic field, the materials receive the energy from
the electromagnetic field and store it, i.e., they gain energy. When the electro-
magnetic field is no longer applied, the energy is released, i.e., it propagates through
the material; propagation of electromagnetic waves is illustrated in Fig. 5.16a). On
the other hand, if the charge distribution and the deflection of spin were delayed,
i.e., too slow, in an alternating electromagnetic field, the energy relaxes in the
substance and becomes a loss (Fig. 5.16b). This is because under applied micro-
waves the dielectric constant has a real part (gain) and an imaginary part (loss). The
next section describes these images by mathematical formulas.

Fig. 5.16 Images of (a) no


absorption of microwaves and
transmission of microwaves
through the sample;
(b) absorption of microwaves
and generation of heat
5.5 Microwave Heating Mechanism 107

5.5.5 Dielectric Heating—Energy Loss in a Microwave


Field

The electric flux density in materials changes according to the change of the external
field. The change of the electric flux density is divided into two parts: (a) follows with
no delay and (b) follows with delay; the change is described by Eq. 5.22:

Zt
DðtÞ ¼ Dp ðtÞ þ Dd ðtÞ ¼ e0 er ð1ÞE ðtÞ þ EðuÞf ðt  uÞdu ð5:22Þ
0

When the electric field vibration at time t is given by E0 ejxt, the electric flux
density that follows without delay (a) is expressed by Dp ejxt, while the electric flux
density (b) that follows the phase delay d is expressed as Dd ej(xt−d) lead from
Eqs. 5.23 through 5.25.

e0 er ðxÞE0 ejxt ¼ Dp ejxt þ Dd ejðxtdÞ ð5:23Þ

e0 er ðxÞE0 ¼ Dp þ Dd ejd ð5:24Þ


 
e0 e0r ðxÞ  ie00r ðxÞ E0 ¼ Dp þ Dd cos d  jDp sin d ð5:25Þ

If the electric flux density in the dielectric substance changes only by dD when
the energy change dU per unit volume, the energy loss w in the dielectric materials
during one cycle (1/f = 2p/x) can then be described by Eq. 5.26:

Z
2p=x Z
2p=x Z
2p=x
dD
w¼ dU ¼ EdD ¼ E dt ð5:26Þ
dt
0 0 0

When E is expressed as a cosine function, D then consists of the parts of the


follow-up with delay and without delay (Eqs. 5.27 and 5.28):

E ¼ E0 cos xt ð5:27Þ

D ¼ Dp cos xt þ Dd cosðxt  dÞ ð5:28Þ

Rearranging D using the addition theorem {cos(x − y) = cos  cos + sin 


sin} followed by differentiation with respect to t yields Eq. 5.29:

@D  
¼ x Dp cos xt þ Dd cos d sin xt þ xDd sin d cos xt
@t ð5:29Þ
¼ xe0 e0r ðxÞE0 sin xt þ xe0 e00r ðxÞE0 cos xt
108 5 Physics of Microwave Heating

and the energy loss w then becomes as given by Eq. 5.30:

Z
2p=x
dD
w¼ E dt
dt
0

Z
2p=x
 
¼ E0 cos xt xe0 e0r ðxÞE0 sin xt þ xe0 e00r ðxÞE0 cos xt dt ð5:30Þ
0
 2 2p=x hxti2p=x
sin xt sin 2xt
¼ e0 E02 e0r ðxÞ þ e00r ðxÞ þ e0 e00r E02
2 4 0 2 0
¼ pe0 e00r E02

Since the energy loss is repeated f times (=x/2p) per second, the energy loss
P that the material receives from the electric field per unit time and unit volume is
expressed by Eq. 5.31:

1
P ¼ pf e0 e00r E02 ¼ xe0 e00r E02 ð5:31Þ
2

The dipolar polarization mechanism only concerns polar compounds, e.g., water,
methanol, and ethanol, which possesses a permanent dipole moment [12]. With
regard to the frequency of the applied electric field, the dipoles can behave in three
ways: (i) at low frequencies, the dipoles keep in phase with the electric field and the
temperature of the irradiated medium hardly changes; (ii) at high frequencies, the
response time of the dipoles is too long to follow very rapid oscillations of the
electric field and no energy is absorbed; and (iii) at the microwave frequencies, the
dipoles’ rotation lags behind the changes of the electric field and results in the
heating of the irradiated material.
The relaxation time s is a measure of the mobility of the molecules (dipoles) that
exist in a material. It is the time required for a displaced system aligned in an
electric field to return to 1/e of its random equilibrium value (or the time required
for dipoles to become oriented in an electric field). Liquid and solid materials have
molecules that are in a condensed state with limited freedom to move when an
electric field is applied. Constant collisions cause internal friction so that the
molecules turn slowly and exponentially approach the final state of orientation
polarization with a relaxation time constant s (Eq. 5.32):

1 1
s¼ ¼ ð5:32Þ
xc 2pfc

When the field is switched off, the sequence is reversed and random distribution
is restored with the same time constant. The relaxation frequency fc is inversely
related to the relaxation time.
5.5 Microwave Heating Mechanism 109

At frequencies below relaxation, the alternating electric field is slow enough that
the dipoles can keep pace with the field variations. Because the polarization is able
to develop fully, the loss (er″) is directly proportional to the frequency (Fig. 5.17).
As the frequency increases, er″ continues to increase but the storage (er′) begins to
decrease due to the phase lag between the dipole alignment and the electric field.
Above the relaxation frequency, both er″ and er′ drop off as the electric field is too
fast to influence the dipole rotation, and the orientation polarization disappears.
Debye relation: Materials that exhibit a single relaxation time constant can be
modeled by the Debye relation, which appears as a characteristic response in
permittivity as a function of frequency (Fig. 5.17). The parameter er′ is constant
above and below the relaxation with the transition occurring near the relaxation
frequency (22 GHz). Additionally, er″ is small above and below relaxation and
peaks in the transition region at the relaxation frequency. In calculating the above
curves, the static (DC) value of the dielectric constant is es = 76.47, the optical
(infinite frequency) value of the dielectric constant is e∞ = 4.9, and the relaxation
time s = 7.2 ps.
Cole–Cole diagram: The complex permittivity may also be shown on a Cole–
Cole diagram by plotting the imaginary part (er″) on the vertical axis and the real
part (er′) on the horizontal axis with the frequency as the independent parameter
(Fig. 5.18).
A Cole–Cole diagram is, to some extent, similar to the Smith chart. A material
that has a single relaxation frequency, as exhibited by the Debye relation, will
appear as a semicircle with its center lying on the horizontal er″ = 0 axis and the
peak of the loss factor occurring at 1/s. A material with multiple relaxation fre-
quencies will be a semicircle (symmetric distribution) or an arc (nonsymmetrical
distribution) with its center lying below the horizontal er″ = 0. The curve in
Fig. 5.10 is a half-circle with its center on the x-axis; the radius of the curve is given
by Eq. 5.33:
es  e1
ð5:33Þ
2

Fig. 5.17 Debye relaxation of water at 30 °C. Reproduced from Ref. [1]
110 5 Physics of Microwave Heating

Fig. 5.18 Cole–Cole diagram of Fig. 5.10. Reproduced from Ref. [1]

The maximum imaginary part of the dielectric constant e′rmax will be equal to the
radius. Note that the frequency moves counterclockwise on the curve of Fig. 5.10.

5.5.6 Magnetic Loss

Magnetic heating is the phase delay of the vibration of the magnetic dipole moment.
Thus, a magnetic loss is obtained by substituting H for E and m for e (Sect. 8.4.3).
The magnetic loss per unit time and volume, Pl , is given by Eq. 5.34:

1
Pl ¼ xl0 l00r H02 ð5:34Þ
2

In metal oxides such as ferrites and other magnetic materials, magnetic losses
occur in the microwave region. These losses are different from hysteresis or Eddy
current losses because they are induced by the domain wall and electron spin
resonance. Such materials are typically placed at positions of magnetic field
maxima for optimal absorption of the microwave energy. Transition metal oxides
(e.g., iron, nickel, and cobalt oxides) possess high magnetic losses and thus can be
used as added loss impurities or additives to induce losses within those solids for
which dielectric loss is too small. For instance, induction heating by the micro-
waves’ magnetic field is expected to occur in catalyzed reactions involving solids,
an example being the rapid induction heating of magnetite (Fe3O4) but not hematite
(Fe2O3) as the latter is not a magnetic material [13–15]. More information and
examples of heating of solid-state magnetic materials are provided in Sect. 8.4.3.
5.6 Penetration Depth and Skin Depth 111

5.6 Penetration Depth and Skin Depth

5.6.1 What Are They?

The next topic we need to consider was a hot topic of discussion at International
Conferences in the latter part of the first decade of the twenty-first century because
experiments of microwave-assisted syntheses gave different results in various
laboratories owing to the use of different equipments from different manufacturers.
This likely occurred because of different penetration depths of the microwaves as
the reactor size and shape in commercial microwave chemical equipment may have
been different depending on the manufacturers, which may have caused variations
in the heat distribution generated for a similar sample size. Even though the
dielectric loss of a solvent may be high, the heating efficiency may sometimes be
low because of the shallow penetration depth of the microwaves. Therefore, if a
large size reactor is used, the microwaves may not heat the center of the sample as
the microwave electric field (E) is attenuated as it traverses the medium and
therefore the power is dissipated; this is a function of E2 so that the power is
reduced to an even greater extent (Fig. 5.19).
The penetration depth or skin depth is a measure of how deeply light or any
other electromagnetic radiation can penetrate a material. It is defined as the depth at
which the intensity of the radiation inside the material falls from its original value at
(or more properly, just beneath) the surface by a specific amount. When performing
microwave chemistry, the penetration depth is an important factor. Unfortunately,
many researchers do not appear to appreciate or understand this important factor.
The reason may be due to a lack of a suitable theory and a correspondingly
appropriate database. Accordingly, we now describe and explain the relevance/
significance of the penetration depth and the skin depth.

Fig. 5.19 Propagation of a plane wave into a lossy medium (lossy medium means a medium that
absorbs microwave energy well and heats). Reproduced from Ref. [8]. Copyright 1983 by Peter
Peregrinus Ltd.
112 5 Physics of Microwave Heating

5.6.2 Penetration Depth

The penetration depth Dp (in cgs units) is the depth at which microwaves penetrate
into a material and the power flux has fallen to 1/e (=36.8%) of its surface value;
That is, it denotes the depth at which the power density of the microwaves is
reduced to 1/e of its initial value. Dp can be estimated from Eq. 5.35 [8],
2 312
k 6 2 7
Dp ¼ 4 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e00  ffi 5 ð5:35Þ
4p e0 1 þ e0  1

where k is the wavelength of the radiation, k(2.45 GHz) = 12.24 cm. Since the
dielectric constant and the dielectric loss are components of this equation, we
expect the penetration depth to change with increase of temperature. For example,
even if we used a 10-cm-diameter reactor, the microwaves would not reach the
center of the sample as the penetration depth at 25 °C is only ca. 1.8 cm for water.
However, at 50 °C, the penetration depth is 3.1 cm while at 90 °C it is 5.4 cm, so
that at the latter temperature the microwaves will reach the center of the water
sample. On the other hand, if ions were added to water, the penetration depth then
would become remarkably shallow. For instance, when NaCl (0.25 M) is added to
pure water, the penetration depth changes to 0.5 cm at room temperature, while the
penetration depth in pure water is 1.8 cm [10]. Therefore, in the case of solutions
containing ions, microwave heating may generate hotspots at the surface of the
reactor and it would then be necessary to carry out vigorous stirring to avoid these
hotspots. Also, because microwave energy is constantly being lost as the micro-
waves enter the water or other material, the greatest intensity of energy is at the
surface and constantly diminishes as it enters the material, so that mixing/stirring
will always be a necessary component.
It is important to examine the heating efficiency and the penetration depth of
microwaves into a substance so as to achieve optimal microwave heating. One must
consider the reactor size with the penetration depth of the microwaves into the
sample for uniform microwave heating to occur. For example, when the tempera-
ture at the solution center reached 75 °C with the microwave energy directed to
only one side of the container of ethylene glycol in a quartz container of diameter
3.0 cm, the temperature of the solution at the container walls was 62 °C due to heat
loss; the resulting temperature gradient was 13 °C (see Fig. 5.20). However, when
the size of the reactor was changed to 1.5 cm and the solution was stirred (magnetic
stirrer), the temperature gradient between the solution center and the container walls
dropped to about 2 °C.
The penetration depths of microwaves into some common organic solvents and
water at 25 °C and at 50 °C are reported in Fig. 5.21 [16]. The penetration depths
Dp of microwaves into polar solvents at 25 °C are generally within 10 cm, for
instance, the penetration depths of 2.45-GHz microwaves in ethylene glycol,
5.6 Penetration Depth and Skin Depth 113

Fig. 5.20 Temperature gradient of ethylene glycol in a quartz reactor (3 and 1.5 cm diameter)
with stirring

Fig. 5.21 Penetration depths of the 2.45-GHz microwaves for some common solvents.
Figure drawn from the data of Ref. [16]. Copyright 2012 by Wiley-VCH Verlag GmbH
114 5 Physics of Microwave Heating

methanol, DMSO, and ethanol are less than 1 cm, so that if these solvents were
contained in a large size reactor and subjected to microwave irradiation, the tem-
perature at the center of the samples would be lower than at the reactor walls. As the
temperature of a solvent rises, however, the microwaves can penetrate deeper into
the sample. Therefore, in order to reach the appropriate reaction temperatures for a
given reactor design, it is necessary to calculate the penetration depth of the
microwaves over a range of temperatures. By contrast, the penetration depths of
microwaves into nonpolar solvents are remarkably greater, as for example, the
penetration depth of microwaves into xylene is 28 cm at 25 °C, while at a tem-
perature of 50 °C the penetration depth is several meters.
In the case of microwave heating of solid samples, the penetration depth is a
very important factor because no stirring is possible. Therefore, the sample size will
be dictated by the balance between the microwave heating efficiency, the loss of
heat to the atmosphere, and the penetration depth. When scaling up is considered in
microwave chemistry, the reproducibility obtained at the laboratory scale will no
longer be possible because of the restriction imposed by penetration depth when
going to larger reactor size. In such a case, a flow-type reactor is often used to carry
out microwave chemistry, preferably using a small reactor so that the penetration
depth is no longer a limiting factor that might affect efficiency. On the other hand,
when synthesizing a polymeric substance, the viscosity of the sample becomes a
limiting factor for the flow-type reactor. A strong stirrer is usually installed in the
batch reactor for polymer syntheses.
The distance for absorption of microwaves to reach 50% of the initial microwave
power is given by the half power density depth D (Fig. 5.22) that can be estimated
from Eq. 5.36:

3:32  107
D¼ pffiffiffiffi ð5:36Þ
f e00

where f is the frequency of the microwaves and e″ is the dielectric loss component
of the substance.

Fig. 5.22 Image of the half


power density depth of the
microwave
5.6 Penetration Depth and Skin Depth 115

5.6.3 Penetration Depths of 915-MHz, 2.45-GHz,


and 5.80-GHz Microwaves

Figure 5.23 illustrates the penetration depths of microwaves at different frequencies


into some selected solvents. Penetration of the 5.8-GHz microwaves was shallower
for all solvents compared to the depth of penetration by the 2.45-GHz microwaves.
For instance, in the case of xylene, the 5.80-GHz microwaves can penetrate to a
depth of ca. 1.24 cm, whereas the depth of penetration is about 28.32 cm for the
2.45-GHz microwaves. In particular, the largest difference in penetration depth
between the 5.80-GHz and 2.45-GHz microwaves was observed for the nonpolar
solvents. Note that, although the 2.45-GHz microwaves penetrate deeper into a
solvent/solution, they have a comparatively low energy conversion, whereas the
5.80-GHz microwaves with their shallower penetration depth display a much
greater energy conversion within the shallower depth. Accordingly, the energy
efficiency for heating thin portions of the solvents/solutions is greater when using
5.80-GHz microwaves. Hence, the penetration depth of the microwaves becomes
one of the principal factors that lead to fast-heating and superheating. In the case of
5.80-GHz irradiation, hexane displayed the deepest penetration depth (2.25 cm).
On the other hand, in the case of 2.45-GHz microwaves, the penetration depth into
ethyl acetate was the shallowest of the nonpolar solvents (11.05 cm).

Fig. 5.23 Penetration depths of microwaves at a frequency of 5.80 GHz, 2.45 GHz, and
915 MHz in water and in 22 pure organic solvents (100 mL) at ambient temperature. Figure was
drawn from data reported in Ref. [16]. Copyright 2012 by Wiley-VCH Verlag GmbH
116 5 Physics of Microwave Heating

The heating efficiency of a solvent with a shallow penetration depth such as


DMSO (0.25 cm), methanol (0.28 cm), ethylene glycol (0.29 cm), DMF (0.46 cm),
and water (3.4 mm) is not high under 5.80-GHz conditions. Accordingly, when
5.80-GHz microwaves are used with a large reactor, the microwave heating effi-
ciency is not high. Thus, a change in penetration depth of the sample involved in a
reaction should be reflected in the design of the microwave reactor.
At the 915-MHz frequency, the characteristic dielectric loss factors (er″) when
compared to 2.45-GHz were greater for ethylene glycol, ethanol, 1-propanol,
2-propanol, and acetic acid anhydride. Thus, the optimal frequency for the heating
of alcohol (but not methanol) is 915 MHz. The penetration depth at the 915-MHz
frequency for all solvents, other than alcohols, is very deep. The largest cylindrical
reactor that can be accommodated in a 915-MHz structure is ca. 24.8 cm, so that it
is suitable for heating polar solvents.

5.6.4 Skin Depth

Everyone understands that light, an electromagnetic wave, is reflected at a metal


surface. Microwaves, which are also electromagnetic waves, are mostly reflected
when used to irradiate highly conducting substances such as metals that possess
limited impedance. The skin effect refers to the tendency of an alternating electric
current to become distributed within a conductor such that the current density is
largest near the surface of the conductor but decreases at greater depths in the
conductor [17]. The electric current flows mainly at the skin of the conductor,
between the outer surface and a level called the skin depth. The skin effect causes
the effective resistance of the conductor to increase at higher frequencies where the
skin depth is smaller, thus reducing the effective cross section of the conductor. The
skin effect is due to opposing Eddy currents induced by the changing magnetic field
resulting from the alternating current [17].
The skin depth, n, can be calculated from Eq. 5.37 in which q is the bulk
resistivity of the sample (X  10−8 m), lr is the relative magnetic permeability, l0
is the magnetic permeability constant (4p  10−7), and f is the frequency of the
microwaves. The skin depths of various metals are listed in Table 5.3.
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
q
n¼ ð5:37Þ
p  f  lr  l0

In the case of bulk metal, even if the surface were exothermic when heated with
microwaves (most are reflected), the resulting heat would be scattered and lost by
the heat conduction mechanism. Therefore, the temperature of the metal sample
does not rise. However, it is possible to heat metallic microparticles using micro-
waves. For example, irradiating a sample consisting of 1-lm copper powder par-
ticles (skin depth, ca. 1.3 lm) with 2.45-GHz microwaves causes microwave
5.6 Penetration Depth and Skin Depth 117

Table 5.3 Skin depths of various metals


Material Chemical formula Bulk resistivity at Relative Skin
20 °C/X  10−8 m permeability/l l−1
0 depth/lm
Aluminum Al 2.65 1 1.66
Beryllium Be 3.3 1 1.85
Brass 70%Cu + 30%Zn 7 1 2.69
Bronze 89%Cu + 11%Sn 15 1 3.94
Chromium Cr 13.2 1 3.69
Copper Cu 1.69 1 1.32
Gold Au 2.2 1 1.51
Graphite C 783.7 1 28.5
Inconel alloy 72%Ni + 16% 103 1 10.3
600 Cr + 8%Fe
Lead Pb 20.6 1 4.61
Magnesium Mg 4.2 1 2.08
Mercury Hg 95.9 1 9.96
Molybdenum Mo 5.7 1 2.43
Palladium Pd 10.8 1 3.34
Platinum Pt 10.58 1 3.31
Rhodium Rh 4.7 1 2.20
Silver Ag 1.63 1 1.30
Tin (pure) Sn 12.6 1 3.61
Titanium Ti 54 1 7.47
Tungsten W 5.4 1 2.36

heating of the copper sample, a phenomenon reported by Roy and coworkers and
who developed this microwave metallic heating method [13] (see Sect. 10.1.3).

5.7 Frequency Effect

5.7.1 Is It Possible to Use Various Frequencies?

As reported elsewhere in this book, microwaves are electromagnetic waves that


span the frequency range between 0.3 and 300 GHz. Accordingly, we do not need
to adhere to the commonly used microwaves at the 2.45-GHz frequency. However,
when examining frequency effects, two barriers have obstructed the researcher’s
curiosity. First, microwave devices that can generate microwave frequencies other
than 0.915, 2.45, and 5.80 GHz are very expensive compared to those devices
operating at the common frequencies. Currently, chemical reaction equipment that
118 5 Physics of Microwave Heating

Fig. 5.24 Global distribution map of frequency bands (GHz) used for industrial, medical, and
scientific (ISM) purposes [16]. Copyright 2012 by Wiley-VCH Verlag GmbH

can generate different frequencies is available at costs similar to those available at


the 2.45-GHz frequency. The second barrier has been with the industrial, scientific,
and medical (ISM) bands that have been reserved internationally for the use of radio
frequency energies employed for industrial, scientific, and medical purposes other
than communications [18]. Examples of applications in these bands include radio
frequency process heating, microwave ovens, and medical diathermy machines.
The powerful emissions of these devices can create electromagnetic interference
and disrupt radio communications that use that frequency, so that these devices
have been limited only to certain frequency bands. The ISM bands are defined by
the ITU-R (International Telecommunication Union—Radio Communications
Sector). The globally used frequencies are 2.45, 5.80, and 24 GHz; individual
countries have established other frequencies. The global distribution map of fre-
quencies used is reported in Fig. 5.24 [16].

5.7.2 Historical Overview of Microwave Frequency Effects


in Chemical Reactions and Sintering

The microwave (MW) frequency response and its impact on chemical reactions
have been examined in only a few cases; that is, research reports regarding fre-
quency effects in microwave-assisted organic syntheses have tended to be rather
scarce. In the field of inorganic chemistry, Möller and Linn [19] used 5.80-GHz
microwaves to calcine ceramics, while Takizawa and coworkers [20] utilized
28-GHz microwaves for the rapid synthesis of (In0.67Fe0.33)2O3, whereas Malinger
et al. [21] used 5.5-GHz microwaves in the rapid syntheses of cryptomelane-type
manganese oxides for catalysis. In the field of organic chemistry, Gedye and Wei
5.7 Frequency Effect 119

[22] used a variable-frequency microwave oven to effect the Knoevenagel reaction


of acetophenone with ethyl cyanoacetate in the presence of piperidine (no solvent;
microwave frequency, 8.1 GHz); with anisole as the solvent, the reaction was
carried out at 12.2 GHz. Séguin et al. [23] reported a new IR reactor cell built with
a microwave cavity operating at a frequency of 5.80 GHz; this equipment allowed
efficient and time-resolved heating of the catalyst. The latter authors also showed
that the 5.80-GHz microwave frequency is the optimal frequency for catalyst
heating. Microwave-induced oligomerization of methane with nickel, iron, and
activated carbon catalysts has been examined relative to microwave frequency and
other parameters by Conde and Suib [24]; an increase in the microwave frequency
enhanced the activity of the catalysts.

5.7.3 Microwave Chemical Equipment for 0.915, 2.45,


and 5.80 GHz Frequencies

Figure 5.25 illustrates three different microwave chemical reaction systems that use
single-mode microwave irradiation applicators at 2.45-GHz, 915-MHz (i.e.,
0.915 GHz), and 5.80-GHz frequencies. Each microwave system is constructed
with a waveguide as the resonant cavity, a short plunger, a three-stub tuner (and/or
an iris), a power monitor, and an isolator. Impedance matching is accomplished by
measuring the output power and the reflected power using the power monitor. The
microwave source for each system is a semiconductor generator.
Differences in the sizes of the waveguide are responsible for variations in the
output power density of the microwave radiation. In this regard, Fig. 5.26 illustrates
the various waveguides seen in the photographs of Fig. 5.25. The cross-sectional
area of the waveguide at the common microwave frequency of 2.45 GHz is
59.645 cm2 (10.92 cm  5.46 cm). On the other hand, the cross-sectional areas of
the waveguides of the 915-MHz and 5.80-GHz apparatuses are, respectively,
306.64 cm2 (24.77 cm  12.38 cm) and 8.0 cm2 (4.0 cm  2.0 cm). For
cross-sectional area comparison, the ratios of the power density (Watt cm−1) in the
waveguides for the 2.45-GHz microwaves relative to the 915-MHz microwaves and
the 2.45-GHz microwaves relative to the 5.80-GHz microwaves are, respectively,
0.19 and 7.46. The corresponding wavelengths of each of the frequencies are
32.79 cm (915-MHz) > 12.24 cm (2.45-GHz) > 5.17 cm (5.80-GHz). The power
density is remarkably lower in the 915-MHz apparatus, which can accommodate a
reactor of a large size. On the other hand, superheating of nonpolar solvents can be
achieved with 5.80-GHz microwaves (see Sect. 4.6.3) because the higher micro-
wave frequency in that cross-sectional area of the waveguide is some threefold
smaller than that of the 2.45-GHz microwaves. Moreover, the overall structure of
the device comprises a much smaller waveguide.
120 5 Physics of Microwave Heating

Fig. 5.25 Commercial


single-mode semiconductor
microwave generator systems
operating at frequencies of
(a) 2.45 GHz
(FSU-201VP-01),
(b) 915 MHz
(FSU-301VP-01), and
(c) 5.80 GHz
(FSU-131VP-01).
Photographs provided by
courtesy of Fuji Electronic
Industrial Co. Ltd. [16].
Copyright 2012 by
Wiley-VCH Verlag GmbH

5.7.4 Frequency Effects for the Common Solvents

Frequency effects are seen even for a simple chemical structure such as that of
water. Changes in the dielectric constant and dielectric loss factors of water as a
function of the microwave frequency in the frequency range 0.1–1000 GHz are
illustrated in Fig. 5.27 [25]. The dielectric loss increases at first with an increase in
frequency reaching a maximum at ca. 10 GHz, and then decreases with further
increase in frequency. By contrast, the dielectric constant is relatively constant at
first and then decreases with increase in the microwave frequency.
The most important characteristics of the 23 solvents examined at two micro-
wave frequencies and ambient temperature are the penetration depth of the
microwaves into the bulk of the solution and the dielectric parameters of these
5.7 Frequency Effect 121

Fig. 5.26 Photograph


displaying the different sizes
and sectional areas of the
waveguides used for
915-MHz microwaves (left),
2.45-GHz microwaves
(center) and 5.80-GHz
microwaves (right).
Reproduced from [16].
Copyright 2012 by
Wiley-VCH Verlag GmbH

Fig. 5.27 Changes in the


dielectric constant (e′) and
dielectric loss (e″) of water at
25 °C as a function of
frequency. Reproduced with
permission from [25]

solvents: the dielectric constant (er′), the dissipation factors [sometimes also referred
to as the dissipation loss tangent, tan d (=er″/er′)], and the dielectric loss factors (er″)
[16]. Note that the dielectric loss factors include the leakage currents. Results show
that the dielectric constants (er′) of all the solvents at the 2.45-GHz microwave
frequency are greater than the analogous constants at 5.80 GHz (Fig. 5.28a). In
particular, such solvents as DMSO, methanol, ethylene glycol, and ethanol display
significant variations in the dielectric constants. Hence, insofar as dielectric con-
stants are concerned, there are but slight advantages in using the 5.80-GHz
microwaves. The dielectric loss (er″) at 5.80 GHz exceeds those at 2.45 GHz for
most solvents except for the alcohols (Fig. 5.28b). Water, DMSO, methanol, DMF,
and ethylene glycol display the largest er″ values at 5.80 GHz, whereas at the
2.45-GHz microwave frequency the dielectric loss factors for methanol, ethylene
glycol, ethanol, 1-propanol, 2-propanol, and acetic acid anhydride are greater than
those at 5.80 GHz. Dissipation factors (tan d) for the 5.8-GHz microwave radiation
are greater than those for 2.45-GHz microwaves for almost all solvents except
ethylene glycol, ethanol, 1-propanol, 2-propanol, and acetic acid anhydride for
which tan d are exceptionally greater at the latter microwave frequency
(Fig. 5.28c). In heating solvents other than these, the expectation from the
122 5 Physics of Microwave Heating

dissipation factors (tan d) is that for each of the solvents the 5.80 GHz would be the
optimal frequency.
Recall that the penetration depth of microwaves is the depth at which micro-
waves penetrate into the material and at which the power flux has fallen to 1/e
(=0.368%) of its surface value. Penetration of the 5.80-GHz microwaves in all
solvents was shallower compared to the depth of penetration by the 2.45-GHz
microwaves (Fig. 5.28d). For instance, in the case of xylene, the 5.80-GHz
microwaves could penetrate to a depth of ca. 1.24 cm, whereas the depth of pen-
etration is about 28.32 cm for the 2.45-GHz microwaves. The largest difference in
penetration depth between the 5.80-GHz and 2.45-GHz microwaves was observed
for the nonpolar solvents. Note that although the 2.45-GHz microwaves penetrate
deeper into a solvent/solution, they have a comparatively low energy conversion,
whereas the 5.80-GHz microwaves with their shallower penetration depth display a
much greater energy conversion within that shallow depth. Accordingly, the energy
efficiency for heating thin portions of the solvents/solutions is greater when using
5.80-GHz microwaves. Hence, the penetration depth of the microwaves is one of
the principal factors that lead to rapid heating and superheating (see Sect. 4.6.3).
For the 915-MHz frequency, the characteristic dielectric loss factors (er″) against
the 2.45-GHz frequency are greater for ethylene glycol, ethanol, 1-propanol,
2-propanol, and acetic acid anhydride. Thus, the optimal frequency for the heating
of alcohols (except for methanol) is 915 MHz. The penetration depth of micro-
waves at the 915-MHz frequency in all solvents other than alcohols is very

Fig. 5.28 Dielectric parameters at 5.80 GHz, 2.45 GHz, and 915 MHz for water and some 22
pure organic solvents at 25 ± 1 °C: (a) dielectric constant, er′; (b) dielectric loss factor, er″;
(c) dissipation factor, tan d; (d) penetration depth of the microwaves (cm). Volume of the sample
was 100 mL. Reproduced from [26]. Copyright 2012 by Elsevier B.V
5.7 Frequency Effect 123

deep. The largest cylindrical reactor that can be accommodated into a 915-MHz
waveguide is ca. 24.8 cm, so that it is suitable for heating polar solvents.
Changes in the dissipation factors (tan d) that accompany microwave heating of
various polar solvents at three different frequencies are displayed in Fig. 5.21 [26];
the temperature was measured with an optical fiber thermometer. An initial increase
in the dissipation factor with increasing temperature was observed at 5.80 GHz for
ethylene glycol, ethanol, 1-propanol, 2-propanol, and acetic anhydride, whereas a
decrease was observed for the other solvents (Fig. 5.29a); when using 2.45-GHz
microwaves, tan d increased initially with temperature only for 1-propanol and
2-propanol (Fig. 5.29b). On application of the 915-MHz frequency, a decrease of
the dissipation factor was observed with increasing temperature for all solvents
(Fig. 5.29c).

Fig. 5.29 Dissipation factors (tan d) at a microwave frequency of (a) 5.80-GHz, (b) 2.45-GHz,
and (c) 915-MHz for pure solvents at various temperatures: Volume of the sample, 100 mL.
Reproduced from [26]. Copyright 2012 by Elsevier B.V
124 5 Physics of Microwave Heating

5.7.5 Rates of Temperature Increase of Common Solvents

Heating rates of common solvents determined in a quartz reactor irradiated con-


tinuously with 5-Watt microwaves from a semiconductor generator and a
single-mode resonator at the 5.80 GHz, 2.45 GHz, and 915 MHz are reported in
Fig. 5.30a–c. Samples of the solvents were located at the maximal position of the
electric field. The heating rate was not affected by the penetration depth when using
a 4-mm reactor. Ratios in heating rates between the 5.80-GHz and 2.45-GHz fre-
quencies are shown in Fig. 5.30d based upon the heating rates from Fig. 5.30a, b.
In some instances, differences were well over tenfold: n-pentane (18-fold), toluene
(15-fold), and xylene (12-fold). Clearly, the largest differences between heating
with 5.80-GHz versus 2.45-GHz microwaves were mostly displayed by the non-
polar solvents. The solvents with the highest heating rate under 2.45-GHz micro-
wave heating were ethylene glycol, ethanol, 2-propanol, and 1-propanol
(Fig. 5.30b).
The heating rates at the 915-MHz frequency were greatest for the alcohols
followed by DMF, DMSO, and water (Fig. 5.30c). On the other hand, most non-
polar solvents were not heated by the 915-MHz radiation. In this experiment, a
quartz test tube (diameter, 20 mm) containing a 10-mL sample was set at the
maximal position of the electric field inside the 915-MHz waveguide. To the extent
that the sample size was small relative to the cross-sectional area of the waveguide,
heating was not very efficient when using the 4-mm reactor; the microwave applied
power was 70 W at this frequency. The ratios of the heating rates at the 915-MHz
frequency to the 2.45-GHz frequency are reported in Fig. 5.30e.

5.8 Frequency Effects in Organic Synthesis

5.8.1 Application to a Diels–Alder Reaction

This section examines reactions using nonpolar substances (reacting substances and
solvents) and shows that some of the issues of microwave chemistry can be
resolved by taking advantage of the frequency effect. The usefulness of 5.80-GHz
microwaves in performing organic syntheses is demonstrated by the synthesis of
3,6-diphenyl-4-n-butylpyridazine (DBP) through a Diels–Alder process using two
representative solvents of different polarities: ethyl acetate and xylene [27].
The Diels–Alder synthesis of 3,6-diphenyl-4-n-butylpyridazine (DBP) from
diphenyl- tetrazine and 1-hexyne reagents in ethyl acetate and xylene solvents is
illustrated in reaction 1; the resulting chemical yields under various experimental
heating conditions after silica gel treatment of the resulting product(s) are listed in
Table 5.4.
5.8 Frequency Effects in Organic Synthesis 125

Fig. 5.30 Rates of increase of temperature for some 23 common solvents under non-stirring
conditions under irradiation with (a) 5.80-GHz and (b) 2.45-GHz microwaves at a power level of 5
Watts using single-mode resonant systems with a semiconductor generator at 5-s intervals up to
30 s; (c) 915-MHz microwave irradiation (70 W; continuous irradiation) using a single-mode
apparatus with semiconductor generator at 5 s intervals; (d) ratios of heating rates for the
5.80-GHz and 2.45-GHz frequencies; (e) ratios of heating rates for the 915-MHz and 2.45-GHz
frequencies; black bars: polar solvents; empty bars: nonpolar solvents. Volumes of samples, 1 mL.
Reproduced from [26]. Copyright 2012 by Elsevier B.V
126 5 Physics of Microwave Heating

Table 5.4 Yields of 3,6-diphenyl-4-n-butylpyridazine (DBP) under various experimental


conditions
Heating method Solvent Temperature/°C DBP yield/% after
90 min 160 min
5.8-GHz Ethyl acetate 100 12 29
Xylene 122 12 38
2.45-GHz Ethyl acetate 43 0 0
Xylene 33 0 0
Oil bath Ethyl acetate 76 <1 2
Xylene 135 5 7
Reproduced from [27]

(reaction 1)

The yield of the DBP product under continuous 5.80-GHz microwave irradiation
ranged between 12 and 29% for ethyl acetate; for xylene, the chemical yield was
12–38%. This was not unexpected for the synthesis of this pyridazine since irra-
diation of the ethyl acetate and xylene solvents by the 2.45-GHz microwaves was
comparatively inefficient. For comparison, heating the xylene solvent with an oil
bath led only to a 5–7% yield of DBP; note that the reactor was introduced into the
oil bath which had been preheated at 80 ºC because the initial rate of temperature
rise was rather slow with this method. Although the yields of the pyridazine product
under 5.80-GHz microwave radiation were not spectacular, they were nonetheless
significantly greater than conventional heating of the reacting solution, and far
better than under 2.45-GHz irradiation.
Temperature-time profiles of the ethyl acetate and xylene solvents used are
illustrated in Fig. 5.31a, b, which show that 5.80-GHz microwave heating was
constantly higher than with the 2.45-GHz microwaves. Irradiation with 2.45-GHz
microwaves for longer times had no effect upon the reaction. The oil bath heating of
ethyl acetate was limited by its boiling point of 77 °C. However, heating of ethyl
acetate with 5.80-GHz microwaves exceeded the boiling point (100 °C after
6 min). Such superheating cannot be achieved by existing conventional methods.
The high chemical yield was caused by the temperature rise under the continuous
5.80-GHz microwave irradiation. The temperature-time profiles for the reacting
substances and the solvents used in the open reactor system are shown in
Fig. 5.23c. The temperature of 3,6-diphenyl-1,2,4,5-tetrazine (tetrazine) and of
1-hexyne increased rapidly under continuous irradiation with the 5.80-GHz
5.8 Frequency Effects in Organic Synthesis 127

Fig. 5.31 Temperature-time profiles of (a) ethyl acetate, and (b) xylene under 5.80-GHz and
2.45-GHz microwave heating, and heating in an oil bath; (c) Temperature-time profiles of naked
3,6-diphenyl-1,2,4,5-tetrazine (i and iii) and 1-hexyne (ii and iv) when subjected to 5.80-GHz (i
and ii) and 2.45-GHz (iii and iv) microwave continuously irradiation (in all cases the solvents were
stirred); (d) Temperature-time profiles of pure solvent and solvent with reacting substances
(sample solution) under 5.80-GHz (I: sample solution with ethyl acetate; II: ethyl acetate; III:
sample solution with xylene; IV: xylene) and 2.45-GHz (V: each of the four samples). Reproduced
from [27]. Copyright 2008 by the American Chemical Society

microwaves. Note that the tetrazine is a solid substance at ambient temperature. The
heating of the reacting substances under otherwise identical conditions progressed
effectively with the 5.80-GHz microwaves as shown in Fig. 5.23d. Temperature
increases of the solvents were somewhat attenuated with the use of 2.45-GHz
microwaves. In the case of the 5.80-GHz microwave system, the rates of increase of
temperature of the sample solution were greater than those of the pure solvents
used. That is, the rate of heating of the reacting agents also aided the
microwave-assisted organic synthesis at the 5.80-GHz frequency.

5.8.2 Synthesis of a Room-Temperature Ionic Liquid


(RTIL)

Both 5.80-GHz and 2.45-GHz microwaves were used to carry out the novel one-pot
solvent-free synthesis of the room-temperature ionic liquid (RTIL) 1-butyl-3-
methylimidazolium tetrafluoroborate ([bmim]BF4) in a reactor containing the three
required reagents, namely 1-methylimidazole, sodium tetrafluoroborate, and
128 5 Physics of Microwave Heating

1-chlorobutane (reaction 2) [28]. The synthesis yields of [bmim]BF4 under various


experimental conditions are listed in Table 5.4. Results clearly demonstrated that
the one-step solvent-free synthesis of this RTIL was possible with microwave
heating at 5.80-GHz and 2.45-GHz frequencies and by conventional heating. The
synthetic efficiencies were significantly different as evidenced by the data in
Table 5.5. In the batch mode, the chemical yields of [bmim]BF4 (87% after a
30-min period) from the 5.80-GHz microwave heating were some three-to-fourfold
greater than the yields obtained with either the 2.45-GHz microwave heating (28%)
or the oil bath heating (21%), even though the temperature reached with 2.45-GHz
microwaves was greater than 200 °C, whereas the corresponding temperature
reached with the 5.80-GHz microwaves was less than 170 °C. The oil bath heating
temperatures reached to synthesize this RTIL were similar to those of the 5.80-GHz
microwaves.

(reaction 2)

The temperature-time profiles of the mixture of reacting substrates in the syn-


thesis of [bmim]BF4 under irradiation with 5.80-GHz and 2.45-GHz microwaves
were quite dissimilar. After 10 min into the process, upon heating with the
5.80-GHz microwaves the measured temperatures were constantly lower than with
the 2.45-GHz microwaves. Temperature increase was not monotonic at the latter
frequency. In fact, for the synthesis of [bmim]BF4 by the batch process, unusual
temperature increases were observed with the 2.45-GHz radiation after 3, 10, and
20 min. This irregular increase was likely due to the formation of the [bmim]BF4
ionic liquid. That is, the heating efficiency with the 2.45-GHz microwaves changed
because the composition of the reacting mixture changed with the formation of
[bmim]BF4. Contrary to the latter (2.45 GHz), the temperature profile observed for
the 5.80-GHz microwave radiation was somewhat noisy beyond 20 min of

Table 5.5 Chemical yields of [bmim]BF4 from the one-step solvent-free synthesis under
irradiation at various microwave frequencies and experimental conditions. Microwave power level
was constant at 30 W
Heating source Reaction time/min Temperature/°C Yield/%
5.80-GHz 30 155 87
60 169 84
2.45-GHz 30 204 28
60 219 29
Oil bath 30 168 21
60 176 25
Reproduced from [28]
5.8 Frequency Effects in Organic Synthesis 129

irradiation, but otherwise of no consequence with regard to the smooth formation of


the product. After a 10-min irradiation period, the differences in the
temperature-time profiles of the individual reacting substrates 1-chlorobutane (CB),
1-methylimidazole (MI), and the solid NaBF4 in the batch mode showed significant
differences on irradiation with 2.45-GHz microwaves (CB: 84 ºC, MI: 147 ºC, and
NaBF4: 56 ºC). Differences in the temperature profiles of the reacting materials
subjected to the 5.80-GHz microwave radiation were smaller (CB: 105 ºC, MI:
124 ºC and NaBF4: 143 ºC, after 10 min) than with the 2.45-GHz microwaves.
Closer examination of the results also revealed that the temperature rise in the
synthesis of [bmim]BF4 with the 2.45-GHz microwaves was more effective than
with the 5.80-GHz microwaves.
The temperature distributions for each of the reacting substances (heterogeneous
CB/MI/NB) were determined with a thermographic system which measures the tem-
perature at the surface of the reactor; temperatures were measured through the slit in the
corresponding 2.45-GHz and 5.80-GHz microwaves’ waveguides [16]. Comparatively,
uniform heating of the CB and MI layers was observed from continuous irradiation
with 5.80-GHz microwaves for a 60-s period. Clearly, best microwave heating with the
2.45-GHz microwaves occurred in the 60-s period at the 1-methylimidazole layer that
contained some dissolved NaBF4. In the case of 2.45-GHz microwave heating, mea-
surements of the temperature distributions by thermographic imaging clearly demon-
strated that the 1-methylimidazole is a heat conductor that converts microwave energy
into thermal energy. Note that the uniform heating distribution of the sample solution
was confirmed by thermographic measurements on heating pure water under 2.45-GHz
microwave irradiation. The temperatures of the CB, MI, and NB layers in Fig. 5.32,
estimated from the thermographic images, were 73.3 °C (CB), 77.0 °C (MI), and
60.9 °C (NB) for continuous 5.80-GHz microwave irradiation and 61.6 °C (CB),
73.1 °C (MI), and 53.5 °C (NB) for continuous 2.45-GHz microwave irradiation. The
differences in temperatures of the layers correlated with the result of differences in the
dielectric loss factors of the pure substances. Note that the surface of the reactor was
heated by thermal conductivity from the sample solution. Further microwave heating
led to boiling of the sample mixture (boiling points of pure CB and MI liquid are 78
and 198 °C, respectively). Subsequently, the macroscopic temperature distribution of
the solution became uniform as a result of convection.

5.8.3 Synthesis of Gemini Surfactants Under 915-MHz


Microwave Irradiation

Dimeric surfactants represent a new class of surfactants and thus the name Gemini
surfactants. Common Gemini cationic surfactants with different chain lengths C12–
C2–C12 have been synthesized in ethanol solvent (reaction 3) using 915-MHz
microwave irradiation (80 Watts) [29]. The synthesis of such surfactants using a
2.45-GHz multimode applicator was also performed for comparison.
130 5 Physics of Microwave Heating

Fig. 5.32 Thermographic images indicating the temperature distributions in the profiles of the
1-chlorobutane (CB) layer, the 1-methylimidazole layer containing a small quantity of dissolved
NaBF4 (MI) layer, and the undissolved NaBF4 salt (NB) layer under microwave heating (MW
power, 10 W) and under non-stirring conditions. Times of continuous microwave irradiation were
(a) 0 s, (b) 60 s at 5.80 GHz, (c) 60 s at 2.45 GHz. Reproduced from [16]

Product yields of C12–C2–C12 Gemini surfactants synthesized with 915-MHz


and 2.45-GHz microwaves after a 60-min period were 34% (915-MHz) and 28%
(2.45-GHz) (Table 5.6). The maximum quantity of the sample solution in a
2.45-GHz single-mode irradiation system was about 50 mL. However, single-mode
irradiation can be performed effectively at the 915-MHz frequency even with a
500-mL scale quantity. Note that using the common 2.45-GHz microwaves to heat
large-scale samples necessitates equipment that is equipped with a circulating
reactor whether using a single-mode applicator or a multimode applicator.

Table 5.6 Chemical yields of a C12–C2–C12 Gemini surfactant in ethanol solvent under 915-MHz
and 2.45-GHz irradiation conditions for a 60-min period
Reflux time/min Output power/W C12–C2–C12 Gemini surfactant/%
915-MHz 60 70 34
2.45-GHz 60 150 28
Oil bath 3-days – 41
Reproduced from [29]
5.8 Frequency Effects in Organic Synthesis 131

5.8.4 Frequency Effects in Nanoparticle Synthesis

The synthesis of inorganic nanoparticles (e.g., metallic, metal oxides and metal
chalcogenides, among others) has made effective use of dielectric heating provided
by microwave radiation [30]. The major role of microwaves in nanoparticle syn-
thesis is rapid and uniform heating. Uniform nano- and bulk particle sizes are
typically obtained by uniform nuclear growth through microwave specific heating.
In this regard, synthetic methods that utilize conventional convective heating,
because of the need for high-temperature initiated nucleation followed by controlled
precursor addition to the reaction media, rely upon a conduction pathway to drive
the synthetic process in which the reactor acts as the intermediary to transfer the
thermal energy from the external heat source to the solvent and ultimately to the
reactants [31]. In addition, uniform nanoparticle synthesis by microwave heating
supports the possibility of scale-up [32]. On the other hand, no synthesis of
nanoparticles takes place when the reacting substrates do not absorb strongly the
microwaves. Moreover, if a microwave absorber (e.g., activated carbon) unrelated
to the reaction system were added it would become an impurity within the reaction.
Accordingly, there are limitations as to the reacting substrates that can be used.
Let us now examine the synthesis of gold nanoparticles as model nanoparticles
in nonpolar and polar media [33]. In particular, we examine the influence of the
microwave frequency on the synthesis of gold nanoparticles in the presence of
oleylamine in a nonpolar solvent of low dielectric constant (e′) that does not absorb
2.45-GHz microwave radiation and so is not dielectrically heated by these micro-
waves (reaction 4). Let us also examine the effect of microwave frequency on the
shape and size of gold nanoparticles that are produced in a highly polar medium
such as ethylene glycol, under otherwise identical temperature conditions to those
used in nonpolar media, so as to probe other possible factors that could influence
the nanoparticle synthesis.

The temperature-time profiles of a mixture of oleylamine and aurochloric acid


were first determined at the two microwave frequencies used (5.80 and 2.45 GHz).
The initial rates of temperature increase for a 6-min irradiation period were
16.5 °C min−1 and 2.79 ºC min−1 for the 5.80-GHz and 2.45-GHz microwaves,
respectively, under a continuously applied microwave power of 45 Watts. The rate
of increase of temperature was nearly fivefold faster for the 5.80-GHz microwaves
than for the 2.45-GHz microwaves, with the latter increasing gradually with irra-
diation time.
132 5 Physics of Microwave Heating

The surface plasmon resonance absorption spectra of the resulting gold


nanoparticles are displayed in Fig. 5.33 for irradiation times of 7 and 15 min for the
5.80-GHz microwaves and 15 and 30 min for the 2.45-GHz microwaves. A surface
plasmon resonance (SPR) band was seen at 528 nm on irradiation with the
5.80-GHz microwaves for an irradiation period of 7 min and 15 min in accord with
the absorption spectra of gold nanoparticles that had been reported earlier by
Mohamed and coworkers [34] and by Yu and Andriola [35]. By contrast, no
absorption bands were seen after a 15-min and 30-min heating time with the
2.45-GHz microwaves, which indicated that no gold nanocolloids formed under
these conditions. Evidently, the Au3+ ions in HAuCl4 are reduced by pure oley-
lamine acting as a reducing agent under the 5.80-GHz microwaves. That the pro-
duct displayed only one Plasmon band at 528 nm was taken as an indication of the
formation of spherical gold nanoparticles; note that oleylamine also acted as a
capping agent [34].
Transmission electron microscopy (TEM) images of the resulting Au nanopar-
ticles produced subsequent to the 5.80-GHz microwave heating for an irradiation
time of 15 min are illustrated in Fig. 5.34, which show a fairly uniform particle size
distribution of the spherical nanocolloids. A random sampling of 35 particles gave a
size distribution of 9 ± 2 nm after the 15-min heating period. By contrast, the TEM
images (not shown) from the 2.45-GHz microwave heating of the oleylamine/
aurochloric acid mixture resulted in no formation of Au nanoparticles even after an
irradiation period of 30 min, which confirm the results observed from the surface
Plasmon resonance spectra of Fig. 5.33. These observations demonstrate that the
lack of formation of gold nanoparticles under the 2.45-GHz microwaves was the
result of not reaching the appropriate reactive temperature. Generally, it is assumed
that microwave-assisted syntheses do not occur in a nonpolar solvent, as the latter is
not a microwave absorber. However, using microwaves of higher frequency (e.g.,
5.80 GHz) it is possible to synthesize gold nanoparticles, even in nonpolar solvents
for the same microwave power and solution composition.

Fig. 5.33 UV–visible


absorption spectra of the
product from irradiating the
oleylamine/aurochloric acid
solution with 5.80-GHz and
2.45-GHz microwaves
(power, 45 W) at the times
indicated. Reproduced from
[33]. Copyright 2011 by the
Royal Society of Chemistry
5.8 Frequency Effects in Organic Synthesis 133

Fig. 5.34 TEM images of the gold nanoparticles produced after the (a) 7 min and (b) the 15 min
of dielectric heating by the 5.8-GHz microwave radiation. Note the larger gold nanoparticles
produced at the longer irradiation time

Gold nanoparticle synthesis by the polyol method has also been investigated in a
polar medium such as in ethylene glycol solvent [36]. Only a small difference in
heating efficiency was observed between the 5.80-GHz and 2.45-GHz microwaves.
The corresponding measured dielectric loss factors were er″ = 9.8 (5.80 GHz) and
14.7 (2.45 GHz), whereas the depth of penetration of the microwaves was esti-
mated to be Dp = 2 mm (5.80 GHz) and 5 mm (2.45 GHz). Any temperature
fluctuation caused by the frequency effect was canceled by the continuous micro-
wave irradiation at 40-Watts power for both the 5.80-GHz and 2.45-GHz micro-
waves. The corresponding initial rates of temperature increase in the first 1 min of
irradiation were both 86 ºC min−1. Interestingly, no differences were seen in the
shape and quantity of Au nanoparticles produced when the reacting substrates were
heated with the same heating profile in a polar solvent and the same two microwave
frequencies were used [36]. Evidently, the frequency effect had no consequence in
the formation of gold nanoparticles in polar media.

5.8.5 Summary Remarks on the Frequency Effect

The frequency effect is attractive in microwave-assisted chemistry and materials


processing. When this effect can be managed, it can be applied to various appli-
cations. For instance, using a single-mode applicator, 915-MHz microwaves can
irradiate a reactor of large size because of the inherent large size of the waveguide
when compared to the waveguide when using 2.45-GHz microwaves. Furthermore,
heating large quantities of alcohols is remarkably effective at the lower 915-MHz
frequency compared with the 2.45-GHz and 5.80-GHz frequencies. At the industrial
scale, it is well known that the extent of energy conversion (more than 80%) and the
operating lifetime (more than 10,000 h) of a 915-MHz magnetron are two attractive
features of this frequency compared to those of the 2.45-GHz magnetron (70% and
134 5 Physics of Microwave Heating

8,000 h). The conditions for which 915-MHz microwaves are more suitable for
making large-sized equipment are well known. On the other hand, heating nonpolar
solvents with 5.80-GHz microwaves is particularly effective owing to the higher
densities of the electric and magnetic fields. Choosing the optimal frequency of the
microwaves for each situation can further promote existing microwave-assisted
chemistry. Moreover, the use of several microwave frequencies could also prove to
be of great help to the chemical industry in microwave organic chemistry. However,
in order to fully understand the microwave frequency effect(s), it is necessary to
perform additional studies in this area so as to identify, and otherwise clarify, some
of the factors that play an important role in microwave chemistry carried out at
different frequencies.

5.9 Electromagnetic and Thermodynamics Simulations

A chemist who is interested in experimenting with microwave-assisted chemistry


will likely inquire how microwave radiation is applied to a sample and how the
resulting heat is distributed. Typically, a sensor is used to visualize microwaves
(Sect. 6.7); however, it is difficult to imagine the distribution of microwaves in
space because the observation of microwaves with a sensor reveals the field only at
the location of the sensor. A similar problem arises when attempting to measure the
temperature distribution with an optical fiber thermometer. Consequently, a simu-
lation is often used to visualize the spatial distribution of microwaves and the
distribution of temperature. Accordingly, this section will attempt to explain the
significance of the use of an electromagnetic field simulation and a heat transfer
simulation.
An electromagnetic field simulation models the target space (and/or sample) by
dissecting the space (or sample) into what will look like a mesh, after which an
equation is used to solve every mesh component. It is then possible to stack the
results for the whole mesh to visualize the distribution of temperature in a sample.
Two methods can be used to solve the relevant integral equations: (a) by the
boundary element method (BEM) and (b) by the finite-difference time-domain
method (FDTD). The electromagnetic field simulation is made possible given
appropriate initial conditions and boundary conditions as required by Maxwell’s
equations to calculate the distribution of microwaves in space-time. The initial
condition is the location of the microwave radiation, while the boundary condition
is the location of the sample in space. There exist appropriate software packages for
computers to solve the equations. A flow diagram of the calculation methodology
will involve the following considerations:
1. It is necessary to restrict and simplify the model to avoid long calculation times
typically seen for long complicated models. In addition, one must also consider
the accuracy of the calculations.
5.9 Electromagnetic and Thermodynamics Simulations 135

2. The input data will be the values of several physical parameters. For instance,
for dielectric factors one uses the actual measurements of the factors with
changes in temperature.
3. The boundary condition will typically be the location of the designated
microwave radiation source.
4. Dissection of the space or sample into a mesh is subsequently made, and to
reduce errors the mesh size is typically made less than one-tenth (1/10) of the
wavelength of the microwave radiation.
5. Finally, the calculations are performed.
In the present instance, to illustrate some simulations carried out using the FDTD
method, liquid samples (benzene, ethylene glycol, and water) with different
microwave heating efficiencies were placed either in the multimode applicator or in
the single-mode applicator, following which the changes in the electric field were
simulated. The results of the electric field distribution and heat distribution in the
multimode applicator are reported in Fig. 5.35.
Benzene, which has the lowest microwave heating efficiency of the three sam-
ples in the multimode applicator, showed no large changes in the distribution of the
electric field, and so the temperature of the benzene hardly increased. For the
ethylene glycol solution, the electric field does not reach the center because the
penetration depth of the microwaves in ethylene glycol is shallow, which results in
significant uneven heating. On the other hand, the microwaves permeate the water
sample to the center, and as such heating occurs progressively and uniformly.
Results of the simulated electric field inside the resonator of a single-mode
applicator for all three samples (benzene, ethylene glycol, and water) within the
waveguide are shown in Fig. 5.36. In the case of benzene, no standing wave is
disordered because no microwaves are absorbed by this nonpolar sample. On the
other hand, the standing wave in the single-mode applicator is significantly disor-
dered in the case of ethylene glycol; internal heating of the ethylene glycol sample
occurs. The standing wave also changes significantly in the case of water; internal
and surface heating also occur.

5.10 Transmission Modes

It is important to remember that air is also present when placing a sample of a pure
material in an applicator; also, a boundary is generated in the reactor between the
components in microwave heating, causing impedance changes. The distribution of
an electromagnetic field in space changes as per its boundary conditions and/or with
the shape of the waveguide. The expression used to classify this state is referred to
as Transmission Modes. As electromagnetic waves, microwaves are transmitted
into free space by the crossed waves of the electric and magnetic fields normal to
the direction of propagation; these are known as a TEM (transverse electric mag-
netic wave) mode. In this section, the microwaves inside the waveguide are
136 5 Physics of Microwave Heating

Fig. 5.35 Simulations of electric field and temperature distributions in three samples of different
microwave heating efficiencies located in a multimode applicator. Equations were solved using the
FDTD methodology. Samples analyzed and shown are benzene, ethylene glycol, and water

depicted as a sine wave. An actual electromagnetic field may be illustrated by the


distribution shown in Fig. 5.37.
When an electromagnetic wave propagates down a waveguide, only one field—
either electric or magnetic—will actually be transverse to the wave’s direction of
travel. The other field will loop longitudinally to the direction of travel, but still be
perpendicular to the other field. Whichever field remains transverse to the direction
of travel determines whether the wave propagates in the TE mode (Transverse
Electric) or in the TM mode (Transverse Magnetic). An electric field of the TE
mode exists in parallel to the short side of the waveguide, and its strength is
distributed over the sine wave distribution along the long side, so that no electric
field component exists along the tube’s axial direction. On the other hand, a
magnetic field is circular and exists along the long side direction of the waveguide
and the tube’s axial direction. Thus, only a magnetic field ingredient exists in the
tube’s axial direction, and the mode, which has no electric field components, is
shown in the TE mode. By contrast, the electric and magnetic fields show a reverse
mode with a TM mode.
5.10 Transmission Modes 137

Fig. 5.36 Simulations of electric field and temperature distributions in three samples of different
microwave heating efficiencies located in a single-mode applicator. Equations were solved using
the FDTD methodology. Samples considered were benzene, ethylene glycol, and water

Fig. 5.37 Microwave distribution image of a waveguide illustrating the presence of a transverse
electric mode (TE) and a transverse magnetic mode (TM)

The distribution of the electric field and the magnetic field in the waveguide is
either in the TE or TM mode; generally, it is the TE mode. These modes change
with a change in the cross-sectional area of the waveguide. The number of electric
138 5 Physics of Microwave Heating

Fig. 5.38 Images of (a) a TE10 mode and (b) a TE20 mode in a waveguide

field intensity distribution in the long side of the waveguide and the number of the
electric field intensity distribution in the short side are added behind the TE mode.
In the TE10 mode, the first image on the left side of Fig. 5.38 displays the number of
half wavelengths (k/2) across the broad dimension of the electric field. The second
image indicates that the length of the narrow dimension is less than k/2 of the
magnetic field. Also, TE10 means that during propagation through the waveguide,
the electric field is perpendicular to the direction of propagation. It is equal to k/2
across the broad dimension, while the magnetic field is less than k/2. In the TE20
mode, the first image shows the number of full wavelengths k across the broad
dimension of the electric field, while the second image indicates that the length of
the narrow dimension is less than the k of the magnetic field. The TE20 means that
the electric field is perpendicular to the direction of propagation.
When showing a standing wave of the resonance type in a single-mode appli-
cator, other numbers may be added to the designation TE10 or TE20. For instance,
when there are two numbers at the peak top of the electric field inside the
single-mode applicator, it is expressed as TE102 (see e.g., Fig. 5.39). When the

Fig. 5.39 Simulations of


TE102 modes in a waveguide
5.10 Transmission Modes 139

regularity of the mode is recognized, then it becomes possible to visualize the


electric and magnetic fields inside the waveguide. For a single-mode applicator,
only one transmission mode exists, whereas for a multimode applicator, more than
one mode exists.

5.11 Coffee Break 5: Using the Microwave Oven

What makes heating foods with microwave energy different from heating them with
conventional heating? Microwave heating can best be described as bulk heating.
Conventional non-microwave ovens heat the surface of foods, after which that heat
slowly migrates by conduction to the interior of the foodstuffs. However, in a
microwave oven, the microwave energy penetrates the food, heating both the
surface and the interior at the same time, so it becomes hot much faster than by
conventional heat conduction. Nonetheless, shape, density, and volume of the food,
along with its dielectric properties and specific heat capacity may create unusual
heat patterns inside the foods. Microwave energy penetrates only a few centimeters
into most foods before it is dissipated, causing some unusual temperature profiles.
For instance,
1. microwaving a 10-cm-diameter beef roast is likely to result in the microwave
energy not penetrating deeply enough to reach the center, which remains
uncooked. Cooking can be completed by heat conduction outside the oven after
first wrapping the roast with aluminum foil, and allowing it to stand for 10 or
15 min during which time the heat from the hot-microwaved areas is conducted
inward to complete cooking. Note that there is no more microwave energy in the
meat—that energy disappears as soon as the microwave oven is turned Off.
2. when microwaving a potato, the opposite may occur, i.e., the microwave energy
concentrates inside the potato and if heated too long may lead to smoke and even
fire. In this case, it is both the smaller size of the potato, permitting deeper
penetration, and the dielectric properties that lead to concentrated central heating.
3. a peculiar phenomenon is that the air inside a microwave oven is cold unless
some other auxiliary heating method is employed. This means that the surface of
foods is never the hottest, but it will be someplace inside the food. This kind of
center heating can create dangerous situations; for example, the interior of baby
food in small glass jars may become hot enough to injure the infant by scalding
its throat. These high interior temperatures may be hidden from the consumer
because the cold air in the oven cools the food and glass surfaces, hiding the
very hot interior. This cold surface phenomenon is also what makes it so difficult
to brown and crisp foods in a microwave oven, unless some other auxiliary heat
form, such as dry steam or hot air, is employed. Some browning and crisping
can also be achieved by means of specially designed food ingredients, or with
special packaging and cookware as will be explained below.
140 5 Physics of Microwave Heating

What are some of the problems related to formulating or designing foods for
microwave ovens? Designing microwavable food products is especially difficult
because of the large variety of microwave ovens in consumers’ homes.
Accordingly, the food products should be compatible with as many varieties of
ovens as possible despite the variations in wattage, size, and otherwise. Some of the
challenges that need to be faced are the following:
(a) Frozen foods will heat differently than refrigerated or room-temperature foods.
This is because ice and frozen foods are very poor absorbers of microwave
energy and heat very poorly, whereas nonfrozen foods are generally much
better microwave absorbers and heat well and faster. Thus, as some portion of
the food thaws, becomes unfrozen, it absorbs microwaves very well and heats
quickly while the area next to it may remain frozen; frozen foods usually heat
nonuniformly.
(b) Salt is an ingredient that may have a profound effect upon how well and
uniformly a food product heats up. Salt reduces the ability of microwaves to
penetrate into foods making it harder to heat the interior. Also, salt has a
peculiar property, a phenomenon called runaway heating, i.e., the hotter a salt
containing food becomes the faster it heats, making its heating very difficult to
control (Fig. 5.40). Most prepared foods contain 1–2% salt, but, where possi-
ble, it is best to reduce this amount to as little salt as possible.
(c) Temperature: water and high-water-containing foods such as vegetables,
meats, and fish tend to heat more slowly the hotter they become (the opposite of
the runaway salt effect noted above); this adds an element of control to prevent
overheating of foods. It is like stepping on the brake in an automobile to slow it
down (Fig. 5.41).

Fig. 5.40 Illustration of


runaway heating, also called
thermal runaway. The hotter
the material becomes, the
faster it heats! Copyright 2017
by R.F. Schiffmann
References 141

Fig. 5.41 Dielectric loss


factor versus temperature.
Note that as water and
high-water-containing foods
become hotter their loss factor
decreases, meaning that they
heat more slowly as they
become hotter. Note also that
Ham, which has a high salt
content, heats faster as it
becomes hotter. Reproduced
with permission from [37].
Copyright 1971 by the
International Microwave
Power Institute

References

1. Keysight Application Note, Basics of Measuring the Dielectric Properties of Materials (http://
literature.cdn.keysight.com/litweb/pdf/5989-2589EN.pdf#search=%27Keysight+Basics+of+
Measuring+the+Dielectric+Properties+of+Materials%27)
2. H. Fukushima, Microwave processing and its applications to the future automobile, in
Proceedings of 1st Global Congress on Microwave Energy Applications, Japan, 2008
3. C. Gabriel, S. Gabriel, E.H. Grant, E.H. Grant, B.S.J. Halstead, D.M.P. Mingos, Dielectric
parameters relevant to microwave dielectric heating. Chem. Soc. Rev. 27, 213–223 (1998)
4. M.S. Venkatesh, G.S.V. Raghavan, An overview of dielectric properties measuring techniques.
Can. Biosyst. Eng. 47, 7.15–7.30 (2005). http://www.csbe-scgab.ca/docs/journal/47/c0231.
pdf#search=%27An+overview+of+dielectric+properties+measuring+techniques%27
5. J. Baker-Jarvis, M.D. Janezic, J.H. Grosvenor, R.G. Geyer, in Transmission/reflection and
short-circuit line methods for measuring permittivity and permeability, NIST Technical Note
1355-R, U.S. Government printing office Washington, (1993)
6. N. Wagner, T. Bore, J.-C. Robinet, D. Coelho, F. Taillade, S. Delepine-Lesoille, Dielectric
relaxation behavior of Callovo-Oxfordian clay rock: A hydraulic-mechanical-electromagnetic
coupling approach. J. Geophys. Res. Solid Earth 118, 4729–4744 (2013)
7. Agilent basics of measuring the dielectric properties of materials, Application Note, June 26,
2006 (http://www3.imperial.ac.uk/pls/portallive/docs/1/11949698.PDF)
8. A.C. Metaxas, R.J. Meredith, Industrial Microwave Heating (Peter Peregrinus, London, UK,
1983)
9. Z. Peng, J.-Y. Hwang, M. Andriese, Magnetic loss in microwave heating. Appl. Phys.
Express 5, 027304-1–027304-3 (2012)
10. S. Horikoshi, T. Sumi, N. Serpone, Unusual effect of the magnetic field component of the
microwave radiation on aqueous electrolyte solutions. J. Microwave Power Electromagn.
Energy 46, 215–228 (2012)
11. S. Horikoshi, N. Serpone, Microwave frequency effect(s) in organic chemistry. Mini-Rev.
Org. Chem. 8, 299–305 (2011)
12. R. Cherbański, E. Molga, Intensification of desorption processes by use of microwaves—an
overview of possible applications and industrial perspectives. Chem. Eng. Process. 1, 48–58
(2008)
142 5 Physics of Microwave Heating

13. R. Roy, D. Agrawal, J. Cheng, S. Gedevanishvil, Full sintering of powdered-metal bodies in a


microwave field. Nature 399, 668 (1999)
14. J. Cheng, R. Roy, D. Agrawal, Experimental proof of major role of magnetic field losses in
microwave heating of metal and metallic composites. J. Mater. Sci. Lett. 20, 1561–1563
(2001)
15. J. Cheng, R. Roy, D. Agrawal, Radically different effects on materials by separated
microwave electric and magnetic fields. Mater. Res. Innov. 5, 170–177 (2002)
16. S. Horikoshi, N. Serpone, in Microwaves in Organic Synthesis, 3rd edn., ed. by A. de la Hoz,
A. Loupy (Wiley-VCH Verlag GmbH, Weinheim, Germany, 2012)
17. https://en.wikipedia.org/wiki/Skin_effect
18. Technical report by A.C. Metaxas (http://www.pueschner.com/downloads/MicrowaveHeating.
pdf#search=‘40680ISMband3390’)
19. M. Möller, H. Linn, New microwave frequency 5.8 GHz for industrial applications. Key Eng.
Mater. 264–268, 735–739 (2004)
20. H. Takizawa, K. Uheda, T. Endo, Rapid formation and growth of bixbyite-type
(In0.67Fe0.33)2O3 by 28 GHz microwave irradiation. J. Am. Ceram. Soc. 83, 2321–2323
(2000)
21. A.K. Malinger, Y.-S. Ding, S. Sithambaram, L. Espinal, S. Gomez, S.L. Suib, Microwave
frequency effects on synthesis of cryptomelane-type manganese oxide and catalytic activity of
cryptomelane precursor. J. Catal. 239, 290–298 (2006)
22. R.N. Gedye, J.B. Wei, Rate enhancement of organic reactions by microwaves at atmospheric
pressure. Can. J. Chem. 76, 525–532 (1998)
23. E. Séguin, S. Thomas, P. Bazin, G. Bond, C. Henriques, F. Thibault-Starzyk, Infrared and
microwaves at 5.8 GHz in a catalytic reactor. Phys. Chem. Chem. Phys. 11, 1697–1701
(2009)
24. L.D. Conde, S.L. Suib, Catalyst nature and frequency effects on the oligomerization of
methane via microwave heating. J. Phys. Chem. B 107, 3663–3670 (2003)
25. J.B. Hasted, Aqueous Dielectrics (Chapman and Hall, UK, 1974)
26. S. Horikoshi, S. Matsuzaki, T. Mitani, N. Serpone, Microwave frequency effects on dielectric
properties of some common solvents and on microwave-assisted syntheses: 2-Allylphenol and
the C12–C2–C12 Gemini surfactant. Rad. Phys. Chem. 81, 1885–1895 (2012)
27. S. Horikoshi, S. Iida, M. Kajitani, S. Sato, N. Serpone, Chemical reactions with a novel
5.8-GHz microwave apparatus. 1. Characterization of properties of common solvents and
application in a Diels–Alder organic synthesis. Org. Process Res. Dev. 12, 257–263 (2008)
28. S. Horikoshi, T. Hamamura, M. Kajitani, M. Yoshizawa-Fujita, N. Serpone, Green chemistry
with a novel 5.8-GHz microwave apparatus. Prompt one-pot solvent-free synthesis of a major
ionic liquid: The 1-butyl-3-methyl- imidazolium tetrafluoroborate system. Org. Process Res.
Dev. 12, 1089–1093 (2008)
29. S. Horikoshi, T. Sato, N. Serpone, Rapid synthesis of Gemini surfactants using a novel
915-MHz microwave apparatus. J. Oleo Sci. 62, 39–44 (2013)
30. I. Bilecka, M. Niederberger, Microwave chemistry for inorganic nanomaterials synthesis.
Nanoscale 2, 1358–1374 (2010)
31. J.A. Gerbec, D. Magana, A. Washington, G.F. Strouse, Microwave-enhanced reaction rates
for nanoparticle synthesis. J. Am. Chem. Soc. 127, 15791–15800 (2005)
32. S. Horikoshi, H. Abe, K. Torigoe, M. Abe, N. Serpone, Access to small size distributions of
nanoparticles by microwave-assisted synthesis. Formation of Ag nanoparticles in aqueous
carboxymethylcellulose solutions in batch and continuous-flow reactors. Nanoscale 2, 1441–
1447 (2010)
33. S. Horikoshi, H. Abe, T. Sumi, K. Torigoe, H. Sakai, N. Serpone, M. Abe, Microwave
frequency effect in the formation of Au nanocolloids in polar and non-polar solvents.
Nanoscale 3, 1697–1702 (2011)
References 143

34. M.B. Mohamed, K.M. Abouzeid, V. Abdelsayed, A.A. Aljarash, M.S. El-Shall, Growth
mechanism of anisotropic gold nanocrystals via microwave synthesis: formation of
dioleamide by gold nanocatalysis. ACS Nano 4, 2766–2772 (2010)
35. L. Yu, A. Andriola, Quantitative gold nanoparticle analysis methods: a review. Talanta 82,
869–875 (2010)
36. M. Tsuji, N. Miyamae, M. Nishio, S. Hikino, N. Ishigami, Shape selective oxidative etching
and growth of single-twin plate-like and multiple-twin decahedral and icosahedral gold
nanocrystals in the presence of au seeds under microwave heating. Bull. Chem. Soc. Jpn 80,
2024–2038 (2007)
37. N. Bengtsson, P.O. Risman, Dielectric properties of foods at 3 GHz as determined by a cavity
perturbation technique. J. Microwave Power Electromagn. Energy 6, 107–123 (1971)

You might also like