It should be noted that there are many diverse properties that need to be studied in the dielectric fillers embedded in a polymer matrix, such as the type of interparticle interactions, the long-range dipole–dipole interactions, the matrix–particle interactions, and clustering, as well as the adsorption of multi-contact chains on the surface of the filler. Below, the microwave absorption only in non-ferromagnetic materials is discussed, characterized by the attenuation coefficient measured in decibels divided by centimeters, as in Expression (15).
2.3.1. Mechanisms of Energy Absorption
The mechanisms of energy absorption in most of the existing dielectrics are based, first, on the losses produced by electrical conductivity or by relatively slow electrical polarization (
Figure 3, curves 1′ and 1″, below 1 GHz). Second, above a frequency of 100 GHz, dielectric losses are due to the lattice vibration mechanism of polarization (
Figure 3, curves 2′ and 2″). Thus, in ordinary dielectrics, their microwave losses are rather low, and it is impractical to use them in absorbing microwave composites.
However, in some materials, in the microwave range (
Figure 3, darkened region, curve 3″), rather high losses are observed, which cannot be explained by simple polarization mechanisms. Experimentally and theoretically, it has been proven that at the gigahertz frequency range in dielectrics with high permittivity, major losses appear due to their internal polarity only when the dielectric has an internal polar phase [
4,
5,
6].
One of the sources of the occurrence of such losses and polar dielectrics could be the anharmonicity in the polarization mechanisms. The main reason of anharmonism in crystalline dielectrics is the asymmetry in the distribution of electronic density along atomic connections. This is caused by the difference in the electronegativity of atoms, which can sometimes be large. Atoms with greater electronegativity attract electrons with higher force, so their charge becomes more negative. On the contrary, atoms with less electronegativity increase their positive charge. Combined, these atoms form non-harmonic polar bonds, leading to high energy losses in materials.
There are three main factors that contribute to microwave losses in polar dielectrics: strong anharmonicity of the lattice vibration [
4,
9], a domain structure in ferroelectrics, and a disordered structure in relaxor ferroelectrics. Domain vibrations provide maximum absorption in the region of 1–10 GHz, while disordered structures generate high losses in the full microwave range [
4,
9]. There are different ways to receive polar structures, but this naturally occurs in ferroelectrics [
5] or can be achieved through the implementation of multilayer or otherwise multicomponent structures [
10,
11,
12,
13].
Two typical cases of interdependent frequency variations of permittivity and effective conductivity in dielectrics and semiconductors are shown in
Figure 4; here, it is assumed that the intrinsic conductivity (seen at constant voltage) is so low that it can be neglected. The increase in effective
σ(
ω) is usually caused by a delay in the polarization mechanism. This effect is due to the close physical connection between the processes of polarization and conduction, which, in principle, can only be completely separated at direct voltage.
The relaxation dispersion of permittivity (
Figure 4) consists of a
ε′(
ω) gradual decrease from the initial value
ε(0) to the end value
ε(∞), when the dielectric contribution Δ
ε =
ε(0) −
ε(∞) of the relaxing polarization mechanism is completely delayed. That is, a gradual increase in the effective conductivity from almost zero to a constant value
σ′eff ≈
ε0Δ
ε/
τ is seen (
τ is the relaxation time). This dependence of conductivity in dielectrics and wide-gap semiconductors is noticed in a very broad frequency range (from 10
–5 to 10
8 Hz); that is, such dependences are typical for the completely different structures and chemical compositions of a material. This common property of effective
σ(
ω) dependence can be described by the power law
σ ~
ωn, where 0.7 < n < 1. This law is fulfilled when charged particles move in the local area in the structure of dielectrics or semiconductors under alternating electrical field influences.
The resonant dispersion of permittivity (
Figure 4) is characterized by the fact that at first, the derivative d
ε′/d
ω is positive; then, at the resonance point, it changes sign to a negative value, but after antiresonance ends, the derivative d
ε′/d
ω again becomes positive. Therefore, dielectric permittivity passes through the maximum and minimum. In the region of the resonance dispersion of permittivity, effective conductivity is characterized by a sharp maximum
σ′max =
ε0Δ
εω0/
Γ, located exactly at the resonance frequency of the oscillator describing this dispersion; here,
ω0 is the oscillator frequency and
Γ is the relative damping factor, while Δ
ε =
ε(0) −
ε(∞) is the dielectric contribution of the oscillator, which describes the dispersion of permittivity.
2.3.2. Dielectrics with High Microwave Absorption
In fact, in many dielectrics, possible relaxation polarization, accompanied by energy absorption, is located mostly at radio frequencies and, as a rule, does not reach the microwave range. In contrast, in ionic dielectrics possessing the resonance dispersion of permittivity, strong absorption of electromagnetic waves is usually located in the terahertz range, i.e., far above the microwave range. Therefore, at first glance, in typical dielectrics, it is hard to expect the strong microwave absorption that is needed for shielding and absorbing applications. However, among dielectrics, there is an important class of materials in which structural instability is observed due to spontaneous changes in their symmetry—these are ferroelectrics. In order–disorder-type ferroelectrics, the frequency of the inherent structure relaxation increases from radio frequencies up to the microwave range, where strong absorption is observed. In ferroelectrics of the displacement type, the natural ionic–lattice dispersion of permittivity falls down from the usual terahertz range and becomes close to the microwave range, with corresponding high absorption. However, the most remarkable in this sense are relaxor ferroelectrics, in which a very wide frequency region of gigantic absorption and dispersion captures the entire microwave range.
Order–disorder-type ferroelectrics demonstrate microwave absorption due to the relaxation of their polarization; examples of well-known ferroelectrics include Rochelle salt (RS) and triglycine sulphate (TGS) (
Figure 5). A very broad maximum absorption is observed in RS near a frequency of 5 GHz; as a result, for example, at a frequency of 10 GHz, the attenuation of a signal passing through an RS sample equals 60 dB/cm. Similarly, in TGS crystals, the absorption maximum is located near a frequency of 100 GHz, where attenuation equals ~200 dB/cm and remains very high, up to submillimeter waves. It should be noted that in RS and TGS, as well as in other order–disorder-type uniaxial ferroelectric crystals, microwave absorption appears only along the polar (ferroelectric) axis, while in other main crystallographic directions, absorption is as low as in ordinary ionic crystals (their microwave attenuation rarely exceeds 0.01 dB/cm).
Displacement-type ferroelectrics are another class of ferroelectrics that also exhibit microwave dielectric dispersion, which may have some prospect for use as fillers in absorbing and shielding composites. The mechanism of microwave absorption in polycrystalline ferroelectrics that possess phase transition of the displacement type is due to their multidomain structure. Even at a large distance down from the phase transition point, when permittivity is already smaller (for example, in lead titanate, PbTiO
3, possessing
Tc = 500 °C), microwave dispersion is observed in the gigahertz range, as shown in
Figure 6a. In this case, the absorption maximum is observed near 3 GHz, when the attenuation of a signal passing through this ferroelectric equals 16 dB/cm. Note that in widely used electronic piezoelectric ceramics PZT = PbZrTiO
3, the same absorption maximum remains in the gigahertz range, but it reaches the magnitude of
ε″ = 120, with an attenuation value near 100 dB/cm at 10 GHz.
Barium titanate (BaTiO
3) and its solid solutions are also promising fillers for flexible microwave-absorbing composites (
Figure 6b). The broad maximum of microwave absorption in ferroelectric BaTiO
3 ceramics can be seen at a frequency of 4 GHz and leads to specific attenuation of 80 dB/cm at a frequency of 10 GHz. For comparison, it should be noted that in paraelectric SrTiO
3, which has approximately the same permittivity but tan
δ = 0.02, the specific attenuation is only 3 dB/cm.
Diffuse phase transition ferroelectrics, for example (as shown in
Figure 6b, solid solution (Ba, Sn)TiO
3), demonstrate an absorption maximum at a lower frequency (i.e., 0.8 GHz), but their attenuation is much higher (i.e., 250 dB/cm at 10 GHz). Other ferroelectrics with diffused phase transition, namely solid solutions (Ba, Sr, Ca)TiO
3 and Pb(Zr, Ti)O
3, have similar properties at microwave range. All of these materials are characterized by a broad
ε(
T) maximum, associated with a random distribution of same-valence cations in the corresponding sublattices. The heterogeneity of a composition in the microregion is accompanied by fluctuation of the Curie temperature, which leads to a diffused
ε′(
T) maximum. The nature of dielectric absorption for the mentioned compositions is similar to ferroelectric BaTiO
3, and their microwave absorption lies within the range 100–300 dB/cm [
14]. In the “electrically soft” ferroelectric solid solutions of (Ba, Sn)TiO
3 (
Figure 6b), the effect of domain absorption is enhanced by the partially disordered crystalline structure. The domain walls in the structure are much wider and occupy a relatively larger volume. In this case, microwave absorption increases and the frequency of its maximum decreases, which may be of interest for selecting necessary fillers for composites designed to be used at a certain frequency range.
Relaxor ferroelectrics seem to be the best candidates among the large-loss dielectrics for fillers in microwave-absorbing composites. In the entire microwave range, these materials have highly absorbing properties. The crystal lattice of relaxor ferroelectrics is characterized by cations of different valences, which randomly occupy similar structural sites. Note that relaxor ferroelectrics might have two different types of structural disordering. For example, lead magnesium niobate (PMN = PbMg
1/3Nb
2/3O
3) has (B
+2–B
+5)-type compositional disordering (the frequency dependence of its dielectric constant and the loss factor are shown in
Figure 7), while potassium-lithium tantalate crystal (KLT = K
1–xLi
xTaO
3) has a strongly disordered structure only in the lithium ions that are located in non-central positions, with various associations between them [
12].
Therefore, relaxor ferroelectrics usually have an ordered crystalline structure, in which “electrical disordering” is observed in the form of built-in quasi-dipole formations. They are easily amenable to orientation in the alternating electrical field, which leads to extremely high polarizability and, consequently, their permittivity reaches value of ~105 (at low frequencies). The relaxation of clusters of different sizes (which are formed from structural quasi-dipoles) occurs over a wide range of frequencies, including microwaves. The diffuse maximum of microwave absorption covers the entire microwave range, which leads to attenuation in the order of 300 dB/cm.
Thus, highly absorbing dielectrics seem to be promising fillers for shielding and absorbent composites in the short-wavelength part of the microwave range. The properties of these composites can be calculated on the basis of experimental data obtained using dielectric spectroscopy.
2.3.3. Model
As can be seen from the material presented above, the absorbing properties of a composite are largely determined by its complex dielectric constant. Therefore, it is important to have a mathematical model that makes it possible to predict the value of the complex dielectric constant of a composite material based on the dielectric properties of its components. As for composites consisting of a dielectric matrix and a dielectric filler, this task does not cause any particular difficulties, but in the case of a metal filler, a number of problems arise. Therefore, we propose a model that makes it possible to predict the value of the complex dielectric constant of polymer–metal composites for metal volume fractions below the percolation threshold.
To describe the properties of the dielectric matrix, relaxor and oscillator models were used. The “relaxor” is a physical model that describes the interaction of dipoles (or polar complexes) with an applied electrical field whose action is opposed by thermal chaotic motion.
The description of relaxation dielectric dispersion is based on the Debye formula [
14]:
The oscillator describing the resonant dielectric dispersion is a physical object that characterizes the dynamic reaction of a system of elastically coupled electrical charges to an externally applied electrical field, which is opposed by internal elastic forces, tending to return the system to a non-polarized state [
15].
The frequency dependence of permittivity on a simple oscillator model can be described by the following expression:
where
ωTO is the transverse optical frequency,
Γ is the relative attenuation, and the difference
ε(0) −
ε(∞) characterizes the dielectric contribution of the ionic polarization mechanism. It is necessary to separate the above equation into the real and imaginary parts of complex permittivity [
16]:
The dielectric constant of a composite material can be calculated using the Lichtenecker formula:
where
i is the number of components of the compound, and
qi is the volume fraction of the
i-th component.
For the case of two components, the Lichtenecker formula has the form:
where
εd is the dielectric constant of the dispersed phase, and
εm is the dielectric constant of the matrix.
The dielectric constant of the metal dispersed phase can be calculated using the Drude–Lorentz theory:
where
ωp is the plasma frequency, and
τd is the time of the free path of electrons.
The free path of electrons can be determined from the following expression:
where
σd is the conductivity of the metal dispersed phase.
The dielectric constant of a polymer matrix can be calculated using the next formula:
where
Ar.i is the amplitude of the
i-th relaxor, and
Aos.j is the amplitude of the
j-th oscillator.
The final formula for the dielectric constant of a two-component composite material is:
The resulting model makes it possible to predict the value of the complex dielectric constant of a polymer–metal composite material in a wide frequency range.
2.3.5. Serial RLC Circuit Method of Dielectric Constant Measurement
For measurement of the dielectric constant of specimens (
ε), a serial RLC (R—resistor, L—inductor, C—capacitor) circuit was used. The measuring circuit (
Figure 11) consists of a signal source U
1, resistor R
1, inductor L
1, capacitor C
1, and registration unit V. As the signal source, we used a DDS (Direct Digital Synthesis) signal generator JOY-IT JDS 6600; for signal registration, we used a digital oscilloscope RIGOL DS 1054 Z; the inductor was chosen with nominal 150 uH, and the resistor was used with nominal 470 Ohm. As capacitor, we used self-made parallel capacitors (
Figure 12a) from 0.035-mm copper foil and samples. The foil was used as conductive plates as it has an adhesive layer that provides a good connection between copper and composites.
The capacitor is connected via small copper foil sockets (
Figure 12b). The sockets fit well to the jumper contacts and provide good contact. Furthermore, this connection method provides simple and fast switching between capacitors.
The measurement process could be divided into three parts: coarse tuning of the RLC circuit, smooth tuning of the RLC circuit, and calculating ε. During coarse tuning, the frequency at the signal source changes in large steps until the signal amplitude decreases. During smooth tuning, two frequency points are found on either side of the maximum amplitude, where the signal begins to decrease. The medium point of this segment is accepted as a resonance frequency f
0. Such a script simplifies the search for the resonant frequency by the maximum amplitude. At the last stage, the previously found f
0 is substituted into Formula (28) to calculate
ε.