Keywords

1 Fundamentals of Dielectric Materials

Dielectric materials are classified as materials in which an electric polarization \(P\) appears under the application of an external electric field \(E\). The macroscopic response of \(P\) can be described by the sum of the microscopic dipole moments \(\mu\) within a unit volume \(V\), \(P=\left(\sum \mu \right)/V\). A relative electronic susceptibility \(\chi\) is a coefficient between \(P\) and \(E\), giving \(P=\chi {\epsilon }_{0}E\), where \({\epsilon }_{0}\) is the vacuum permittivity. On the other hand, the relative dielectric permittivity \(\epsilon\) is a coefficient between the electronic displacement \(D\) and \(E\), giving \(D=\epsilon {\epsilon }_{0}E={\epsilon }_{0}E+P=\left(1+\chi \right){\epsilon }_{0}E\), thereby \(\epsilon =1+\chi\). In contrast to the macroscopic coefficients \(\chi\) and \(\epsilon\), the polarizability \(\alpha\) defines a microscopic polar response of materials that relates \(\mu\) and \(E\) as \(\mu =\alpha E\). Hereafter, \(\epsilon\) is called the permittivity for simplicity. The dielectric permittivity \(\epsilon\) is generally a complex quantity, so it is expressed as \(\epsilon ={\epsilon }^{\prime}+i{\epsilon }^{{\prime}{\prime}}\). Conventionally, we use a ratio between the real and the imaginary parts of permittivity, \(\mathrm{tan}\delta ={\epsilon }^{{\prime}{\prime}}/{\epsilon }^{\prime}\), to indicate the magnitude of energy dissipation during a charge–discharge process for the dielectric materials. There are various microscopic origins for the dipole moment \(\mu\) in the materials, for instance, polar molecules, the relative displacement of oppositely charged ions or particles, ferroelectric domains, accumulated carriers around the interfaces, and so on. Since each \(\mu\) has an individual time scale, the dielectric response of materials shows rich frequency dispersion, as shown in Fig. 13.1. In general, the response of polar molecules, \({\mu }_{\mathrm{d}}\), gives a relaxation-type frequency dispersion, while a resonance-type frequency dispersion appears due to the response of ionic displacement, \({\mu }_{\mathrm{ion}}\), or the electronic one, \({\mu }_{\mathrm{e}}\). In contrast, the extrinsic components due to the domains, \({\mu }_{\mathrm{domain}}\), and the accumulated space charges, \({\mu }_{\mathrm{sc}}\), have a relatively slow response; therefore, their responses are usually observed in the low-frequency region below 106 Hz. The frequency dispersions of \({\mu }_{\mathrm{domain}}\) and \({\mu }_{\mathrm{sc}}\) are of the relaxation type. Note that the extrinsic component means that the dielectric response does not stem from an inherent nature of the material but is caused by various inhomogeneities.

Fig. 13.1
A graph plots epsilon minus 1 versus omega. It has various peaks labeled as mu domain, mu s c, mu d, mu ion, and mu e 10 power 6, 9, 12, and 15.

The schematic illustration of the frequency dispersion of dielectric permittivity and the corresponding dipole components

2 Permittivity Engineering in a Homogeneous System

There are several ways to design dielectric materials according to a range of permittivity. For ordinary dielectric materials with permittivities 100–101, the permittivity can be roughly estimated with the additivity rule of polarizability, where the total polarizability is given by the addition of individual constituent ions, as presented in Fig. 13.2 [1]. In \({\text{Ba}}^{{{2}+}} {\text{B}}_{2}^{4+} {\text{O}}_{4}^{2-}\), for instance, the total polarizability of 14.54 is estimated from the ionic polarizability of Ba2+, B4+, and O2− of 6.40, 0.05, and 2.01 in a unit of Å3, respectively. The permittivity \(\epsilon\) is then calculated with the Clausius–Mossotti relation, to be 6.14, where \(N\) is the number of dipole moments in a unit volume. This value is nearly equal to the experimental value of 6.05. The applicability of this concept is, however, limited to harmonic materials in which the dynamics of lattice vibrations is harmonic.

$$\frac{N\alpha }{3{\epsilon }_{0}}=\frac{\epsilon -1}{\epsilon +2}$$
Fig. 13.2
A periodic table with the values of the polarizability of ions.

The polarizability of ions, which is given a unit of Å3 (https://doi.org/10.1063/1.353856)

In anharmonic materials, where some of long-wavelength lattice vibrations or the corrective fluctuations of dipole moments are unstable due to various interactions, a permittivity on the order of 102–103 is often observed. Ferroelectric materials are typical cases. Ferroelectric materials are practically defined as materials having spontaneous polarization, the direction of which can be switched by the applied electric field. From a crystallographic viewpoint, all polar materials are potentially ferroelectric. Many ferroelectric materials show a ferroelectric phase transition from a nonpolar phase to a polar phase on cooling. If a primary-order parameter of the ferroelectric phase transition is the polarization \(P\), the permittivity \(\epsilon\) divergently increases around a ferroelectric phase transition temperature. In the case of the perovskite-type oxide BaTiO3, for example, the high-temperature nonpolar phase of \(Pm{\overline{\text{3}}}m\) symmetry transforms into the low-temperature polar phase of P4mm symmetry at 123 °C [2], around which the permittivity reaches the order of 103 (Fig. 13.3) [3]. It is known that the unstable lattice vibration, which is called the “ferroelectric soft mode,” plays an important role in the ferroelectric phase transition of perovskite-type oxides [4]. The ferroelectric soft mode is a special transverse optical phonon mode at the \(\Gamma\)-point of the Brillouin zone, whose frequency decreases toward zero as the temperature approaches the ferroelectric phase transition temperature Tc due to the strong anharmonic interaction. According to the Lyddane–Sachs–Teller relation, the permittivity is inversely proportional to the square of the soft mode frequency, and the decreasing frequency of the soft mode therefore raises the permittivity of the material [5, 6]. It is known that the soft mode also enhances the permittivity of materials even in the case when the system stays paraelectric until zero Kelvin. A typical example can be seen in the rutile-type TiO2, whose permittivity reaches 270 along the crystallographic [001] direction in the low-temperature region (Fig. 13.4) [7]. The soft mode in the rutile-type TiO2 decreases its frequency to raise the permittivity as the temperature decreases but never reaches zero at finite temperatures [8]. This behavior is called “incipient ferroelectricity.” Quantum paraelectricity in SrTiO3 is a special case of incipient ferroelectricity, where the system tends to go ferroelectric at a finite temperature but is suppressed by quantum fluctuation [9]. A delicate balance between the soft mode instability and the quantum fluctuation gives rise to the extremely large permittivity of the order of 103–104 in the low-temperature region [10].

Fig. 13.3
A graph plots the dielectric constant kappa dash versus temperature in degrees Celsius. It shows values for f equal to 1 kilohertz, 1 megahertz, 100 megahertz, and 9.4 gigahertz. The values show high peaks between 10 and 150. A graph inset plots kappa dash versus temperature. It plots a high peak at nearly 130.

Temperature dependence of the dielectric permittivity of BaTiO3 ceramics measured with various frequencies. See details in Ref. [3] (https://doi.org/10.1080/00150198808008805)

Fig. 13.4
A graph plots the dielectric constant versus temperature in degrees kelvin. It shows values for c and a that are in a slightly decreasing trend. c starts above 250, and a starts above 100.

Temperature dependence of the dielectric permittivity of a rutile-type TiO2 single crystal, which is measured along the tetragonal a and c axes (https://doi.org/10.1103/PhysRev.124.1719)

Several concepts for designing ferroelectric materials have recently been established in perovskite type and related oxides owing to the development of theoretical and experimental techniques. Covalency between oxygen and constituent ions is one of the clues for internal distortion within coordination polyhedra [11, 12], and a comparison between CaTiO3 and CdTiO3 provides a good example [13,14,15]. These compounds crystallize in the same orthorhombic crystal structure of centrosymmetric Pnma(Pbnm) symmetry at room temperature, where oxygen octahedra rotate as a+bb in Glazer’s notation [16, 17]. CdTiO3 undergoes a ferroelectric phase transition to the polar Pna21 structure at Tc ~ 85 K [18], whereas CaTiO3 remains in the Pnma structure until 0 K. Figure 13.5 presents the temperature dependence of dielectric permittivity in (Cd1−xCax)TiO3 with the Ca concentration x = 0–0.05. In the composition for x = 0, as shown in the figure, the permittivity rapidly increases as the temperature approaches Tc. The ferroelectric transition temperature Tc decreases with increments of x, indicating suppression of ferroelectric instability due to the Ca substitution. The Ca-concentration dependence of Tc, presented in the inset of Fig. 13.5, indicates that the ferroelectricity disappears around x = 0.05. In simple terms, the covalency can be measured by the difference in electronegativity; the closer the electronegativity, the stronger the covalency. In the case of (Cd1−xCax)TiO3, Cd has stronger covalency with O than Ca, because the electronegativity of Cd and Ca are 1.69 and 1.00, respectively, whereas that of O is 3.44. Figure 13.6 shows the electronic density of states (DOS) for CdTiO3 and CaTiO3, which are obtained with first-principles calculations, where partial DOSs for the s-orbital of Cd(Ca) and p-orbital of O are denoted by red and blue hatches. Compared to the CaTiO3, CdTiO3 has larger overlap between the s- and p-orbitals, visualizing a stronger covalency of Cd–O than Ca–O. From a crystallographic viewpoint, the Pnma structure of CdTiO3 has one symmetric O–Cd–O bond along the [010] direction, in addition to two asymmetric O–Cd–O bonds along the [101] and [−101] directions (Fig. 13.7). Although the asymmetric O–Cd–O bonds generate local dipole moments, they mutually cancel out to net zero polarization within the plane normal to [010]. Along the [010] direction, in contrast, the symmetric O–Cd–O bond induces no dipole moment intrinsically in the paraelectric phase. Like the O–Cd–O bonds along the [101] and [−101] directions, the Cd atom in O–Cd–O along the [010] direction also possesses the tendency to displace toward one of the adjacent oxygen atoms due to the strong covalency of Cd–O. In a sufficiently high-temperature region, however, thermal fluctuation leads to an average position of Cd at the center of the O–Cd–O bond, thus rendering the macroscopic symmetry centrosymmetric. As the thermal fluctuation decreases on cooling, the effect of covalency becomes more dominant, and the Cd displaces to induce the dipole moment below the specific temperature Tc. This picture can also be seen from the viewpoint of the lattice vibration. Figure 13.8a presents the phonon dispersion curves of CdTiO3, which are calculated in the paraelectric phase of the Pnma structure, where a position on the vertical axis below zero indicates an imaginary frequency. At the \(\Gamma\)-point (k = 0) in the figure, the frequency of one optical phonon branch is found to be imaginary, indicating that its vibration is unstable in the Pnma structure without any support from thermal fluctuation. This mode is the ferroelectric soft mode of CdTiO3, which drives the phase transition. A displacement pattern of the ferroelectric soft mode is shown in Fig. 13.8b. As shown in the figure, Cd atoms oscillate between two adjacent oxygens, and freezing this vibration at the ferroelectric phase transition temperature Tc induces the polar displacement of Cd in the O–Cd–O bond. Note that the reason why the ferroelectric displacement of Cd atoms at k = 0 is more energetically favorable than the non-ferroelectric displacement at k ≠ 0 is beyond the present scope. It is probable that a further concept is required to deduce the energetics of an O–Cd–O–Cd–O extended bond in an intuitive way.

Fig. 13.5
A graph plots the dielectric constant versus temperature in degrees kelvin. It shows values for x equal to 0, 0.01, 0.03, and 0.05. It shows high peaks between 0 and 150. It has a graph inset that plots temperature versus calcium concentration x. The value for x c equals 0.047.

Temperature dependence of the dielectric permittivity in (Cd1−xCax)TiO3 with the Ca concentration x = 0–0.05. The inset presents a variation of the ferroelectric phase transition temperature Tc as a function of the Ca content x

Fig. 13.6
A graph plots the density of states versus energy. It has values for O p, C d s, and total for the graph at the left and O p, C a s, and total for the graph at the right. The values for C d s and C a s remain nearly on the x axis between 0 and 1. The peaks for O p and total rise above 20.

Electronic density of states (DOS) for CdTiO3 and CaTiO3, where partial DOSs for the s-orbital of Cd(Ca) and p-orbital of O are denoted by positive and negative slope lines

Fig. 13.7
A schematic illustration of C d T i O 3. It consists of 12 octahedra interconnected by yellow and small red spheres. Six octahedra belong to the paraelectric phase, and the remaining six belong to the ferroelectric phase.

Schematic illustrations for the O–Cd–O bonds in CdTiO3 at the paraelectric and ferroelectric phases. The large yellow spheres, small red spheres, and octahedra in the figure denote the Cd atoms, O atoms, and TiO6 units, respectively

Fig. 13.8
Two illustrations. a is a graph plots frequency versus wave vector. It shows various horizontal, slightly fluctuating lines. b is a schematic of C d T i O 3 that shows 12 octahedra connected with yellow and small red spheres.

a The phonon dispersion for the paraelectric phase of CdTiO3, which is obtained by first-principles calculation at 0 K. b A displacement pattern of the ferroelectric soft mode in CdTiO3, where the large yellow spheres, small red spheres, and octahedra denote Cd atoms, O atoms, and TiO6 units, respectively. The arrows in the figure schematically indicate the amplitude of the Cd oscillation in the soft mode

3 Permittivity Engineering in an Inhomogeneous System

Designing dielectric materials with a permittivity greater than 104 often requires the extra contribution from the inhomogeneity in materials in order to increase the dielectric response. Typical examples are polar nano regions (PNRs) in relaxors [19,20,21,22,23,24]. Relaxors are categorized as materials that show a broad and dispersive peak in the temperature dependence of dielectric permittivity. The temperature at which the permittivity reaches the cusp, Tm, strongly depends on a frequency of AC voltage used in the dielectric measurement, where Tm shifts to the higher temperature region as the frequency elevates (Fig. 13.9). When we estimate a relaxation time τm from a specific frequency fm (= 1/τm), at which the imaginary part of the permittivity culminates, as a function of temperature, the relation between τm and the temperature follows the Vogel–Fulcher law, \({\tau }_{\mathrm{m}}={\tau }_{0}\mathrm{exp}\left({E}_{a}/{k}_{\mathrm{B}}\left({T}_{\mathrm{m}}-{T}_{\mathrm{f}}\right)\right)\), indicating growing clusters in the system [25, 26]. Many studies have clarified the polar clusters of several tens of nanometers, which are called PNRs, as they grow on cooling, imparting an extrinsic polarization response to the applied electric field to improve dielectric permittivity over a wide temperature range around Tm [19,20,21,22,23,24,25,26,27]. The PNRs often arise because of a disturbance in the long-range polar ordering due to the ferroelectric phase transition resulting from atomic-scale heterogeneity in the system.

Fig. 13.9
A graph plots 10 superscript negative 3 epsilon dash versus temperature. It shows values for P M N, 10 superscript negative 2, 1, 10 superscript 2, 10 superscript 4, and 10 superscript 5. They show two different sets of peaks between 250 and 300.

Temperature dependence of real and imaginary parts of dielectric permittivity in Pb(Mg1/3Nb2/3)O3 (PMN), which is measured at various frequencies (https://doi.org/10.1007/s10853-005-5915-7)

Pb(Mg1/3Nb2/3)O3 (PMN) is a typical relaxor material with a perovskite-type crystal structure that has been widely studied to date. In PMN, Mg–Nb one-to-one complexes with negative charge, which are called chemically ordered regions (CORs), serve as a source of a quenched random electric field in the matrix [28,29,30]. PMN has intrinsic ferroelectric instability, as a previous Raman scattering study demonstrates. Figure 13.10 shows the temperature dependence of Raman spectra that were observed in the PMN single crystal [31]. As shown in the figure, the low-frequency component gradually grows as the temperature approaches Tc ~ 200 K. Note that, according to Fu et al. [24], Tc in PMN in this context is defined as the temperature at which the permittivity takes a maximum under the application of an electric field sufficiently high to induce long-range polar ordering, although the relaxors do not have a definite transition temperature under a zero electric field. The temperature dependence of the frequency for the low-frequency component, which is estimated by spectral analyses, is plotted in Fig. 13.11, along with a typical temperature dependence of the soft mode observed in CdTiO3. In the figure, the blue, red, and black circles denote the harmonic frequency for the soft mode \({\omega }_{\mathrm{s}0}\), damping constant for the soft mode’s vibration \({\gamma }_{\mathrm{s}}\), and the soft mode frequency \({\omega }_{\mathrm{s}}\) \((=\sqrt{{\omega }_{\mathrm{s}0}^{2}-{\gamma }_{\mathrm{s}}^{2}})\), respectively. The figure shows that, in both cases for PMN and CdTiO3, the soft mode frequency \({\omega }_{\mathrm{s}}\) decreases as the temperature approaches Tc, indicating the typical soft mode-driven phase transition. In contrast to the qualitative similarity in the softening behavior, the temperature scale is markedly extended in PMN compared to CdTiO3. The difference in the soft mode dynamics is also seen in the stability of the oscillation, which can be evaluated by a relation between \({\omega }_{\mathrm{s}}\) and \({\gamma }_{\mathrm{s}}\); the oscillation of the soft mode is stable in an underdamped region of \({\omega }_{\mathrm{s}0}> {\gamma }_{\mathrm{s}}\), whereas it is unstable in an overdamped region of \({\omega }_{\mathrm{s}0}< {\gamma }_{\mathrm{s}}.\) As shown in Fig. 13.11, a wide overdamped region spreads before the transition temperature, suggesting that the spatial inhomogeneity fractionates a coherent length of the soft mode to disturb the long-range polar ordering. Such a localized soft mode is a potential mechanism for the formation of PNR and the relaxor phenomena [24, 31, 32].

Fig. 13.10
A graph plots Raman intensity versus wavenumber per centimeter.

Temperature dependence of Raman spectra observed in the PMN single crystal (https://doi.org/10.1002/jrs.2746)

Fig. 13.11
Two scatterplots plot the wavenumber per centimeter versus temperature in Kelvin. They are labeled P M N and C d T i O 3 on the left and right, respectively.

a Temperature dependence of the harmonic frequency for the soft mode \({\omega }_{\mathrm{s}0}\) (blue), damping constant \({\gamma }_{\mathrm{s}}\) (red), and the soft mode frequency \({\omega }_{\mathrm{s}}\) \((=\sqrt{{\omega }_{\mathrm{s}0}^{2}-{\gamma }_{\mathrm{s}}^{2}})\) (black) for a PMN and b CdTiO3, where the mode frequency is plotted in terms of cm−1 according to convention (https://doi.org/10.1002/jrs.2746)

A barrier layer capacitor (BLC) effect is the other concept leading to permittivity greater than 104, although it is not the true permittivity but an apparent one. In the perovskite-related paraelectric CaCu3Ti4O12 (CCTO), a permittivity on the order of 104 is observed over a wide frequency range below 106 Hz (Fig. 13.12) [33, 34]. In marked contrast to the case of ferroelectric materials that show a high permittivity only around Tc, the permittivity here shows little temperature dependence over a wide temperature region, depending on the frequency (Fig. 13.13) [35]; in the case of the frequency of 1.37 kHz, for instance, the high permittivity survives until approximately 100 K, as presented in Fig. 13.13. Upon further cooling, the permittivity steeply decreases to the order of 101. This kind of dielectric response is known to stem from electric inhomogeneity in the material due to, for instance, a core–shell structure of grains and/or several kinds of dislocations, where one part of the material is semiconducting and the other is insulating [36,37,38]. Figure 13.14 shows (a) the topography and (b) the current image of the CCTO ceramics sample, as observed by atomic force microscopy (AFM). The contrast in panel (b) indicates the magnitude of the observed current under application of bias voltage, where the darker color denotes the larger current. As shown in magnified images for the grain and the grain boundary, which are respectively presented in panels (c) and (d), there is an inhomogeneous distribution of electric conductivity in the sample. The spatial scale of the inhomogeneity is found to be on the order of a nanometer, as indicated in the panel. The line profile of the observed electric current along the line from E to F in the panel (b) is shown in the panel (e). Panel (f) presents the results of local IV measurements performed in selected areas in the inside of the grain and around the grain boundary, as labeled in panel (b). As shown in the figure, an insulating property is found in areas A and B, whereas relatively large electric conduction is detected in areas C and D. When an electric field is applied to a system with such an electric inhomogeneity, the carriers are blocked by the insulating area after traveling within the conductive area to induce huge dielectric displacement. This behavior can be modeled by the system with two different components aligned in series, as presented in Fig. 13.15, where they have mutually different electric properties mimicking the conducting and the insulating areas in the material. Figure 13.15 also presents the temperature dependence of the dielectric permittivity and the loss of two CCTO samples with the solid curves, where #A and #B are the as-synthesized and the post-annealed samples, respectively (See Ref. [37] for details). The simulated curves with the equivalent circuit model are plotted with broken lines in the figure, where the activation energies used in the simulations are experimentally determined with an Arrhenius plot in the inset. As presented in the figure, the simulated curves are in good agreement with experiment and confirm that the huge permittivity in CCTO stems from the BLC effect, which is discussed in the Sect. 13.3. This kind of dielectric phenomenon is also called the Maxwell–Wagner effect [39].

Fig. 13.12
Two graphs plot epsilon dash and epsilon double dash versus frequency in hertz. It gives values for 4, 10, 20, 40, and 80 hours.

(Top) Real and (bottom) imaginary parts of the dielectric permittivity of CCTO at room temperature as a function of frequency. The data are measured in the samples with different sintering times (https://doi.org/10.1063/1.2077864)

Fig. 13.13
A graph plots the epsilon dash and sigma dash versus temperature in Kelvin. The graph is labeled as C a C u 3 T i O 12.

Temperature dependence of (top) permittivity and (bottom) resistivity of CCTO measured with various frequencies (https://doi.org/10.1103/PhysRevB.70.172102)

Fig. 13.14
Six illustrations labeled from a to f. The first four are A F M images. e exhibits line profile representation of current measured between points E and F. f plots current voltage properties. The graph indicates measurements taken at points A, B, C, and D.

a The surface topology and bd the current image of CCTO ceramics observed with atomic force microscopy (AFM). e The line profile of current measured between the points E and F in panel b. The IV properties measured at the points A, B, C, and D in panel b (https://doi.org/10.1021/cm0710507)

Fig. 13.15
A graph plots log epsilon dash comma loss versus temperature in Kelvin. It gives values for experimentation and fit. It also gives values for f equal to 100 kilohertz.

The two-components model describing the dielectric properties of CCTO. See text for details (https://doi.org/10.1021/cm0710507)

4 Toward Functional Local Structures in Dielectric Materials

As introduced in the preceding section, local structures often play a role in enhancing the permittivity. Considering the additivity rule of polarizability, the creation of a unique local structure with extremely large polarizability in a homogeneous medium will probably raise the permittivity markedly. This idea provides a way for designing “hyperordered structures (HOSs)” in dielectric materials. Electron-pinned defect-dipoles (EPDDs), which were reported in 2013, would be categorized as a local structure of this type [40].

It has been reported that, when Ti of the rutile-type TiO2 is slightly substituted by the same amount of Nb and In, the permittivity of TiO2 markedly increases to the order of 104 over a wide frequency range from 101 to 106 Hz (Fig. 13.16). As shown in the figure, only 0.05% of co-dopants are found to lift the permittivity to approximately 2 × 104 with tanδ smaller than 10–1. As the concentration of co-dopants increases to 10%, furthermore, the permittivity reaches 6 × 104, keeping the value of tanδ sufficiently small. An origin of this remarkable effect was initially attributed to the special complex defect pair, called EPDD. Figure 13.17 presents the schematic illustration of EPDD, which is deduced by first-principles calculations. As shown in the figure, the co-substitution of Nb and In introduces two types of defect complexes, \({\text{In}}_{2}^{3 + } {\text{V}}_{{\text{O}}}^{ \cdot \cdot } {\text{Ti}}^{{{3} + }}\) and \({\text{Nb}}_{2}^{5 + } {\text{Ti}}^{{{3} + }} {\text{A}}_{{{\text{Ti}}}}\) (ATi = Ti3+, Ti4+, or In3+), and their overlap forms the highly polarizable extended defect complex, EPDD. Such an extended local structure can be regarded as HOS, which enables marked enhancement of the permittivity in dielectric materials. Several follow-up studies, however, pointed out that the BLC effect contributes to the observed high permittivity [40,41,42,43,44,45]. Presented in Fig. 13.18 is a current–voltage plot for 1% Nb + In co-doped TiO2 (NITO-1%) ceramics, which is measured in several characteristic areas of the sample. Panel (a) shows an optical microscopic image of the NITO-1% ceramic sample on which micro-electrodes are attached (bright areas). The current–voltage property was measured between the electrodes of 1–2, 1–3, and 3–4. There is no grain boundary between the electrodes 1–2, whereas 1–3 and 3–4 run across the grain boundaries. Between the electrodes 1–2, as plotted in panel (b), large current flows show a semiconductive property of the grain core. Between the electrodes 1–3 and 3–4, in contrast, the current is found to be very small under the application of an electric field, indicating a highly insulating property of the grain boundaries. This result clarifies the BLC effect due to the core–shell structure of grains in NITO-1%. It should be noted that Nb5+ and In3+ are nominally chosen to avoid generation of Ti3+, which seriously deteriorates the insulating property of TiO2. The observed semiconducting property is probably due to the spatial inhomogeneity of the Nb-to-In ratio, which induces partial generation of Ti3+ in the sample.

Fig. 13.16
Two graphs plot permittivity and tan delta versus frequency. It shows varied horizontal lines.

(Top) Dielectric permittivity and (bottom) tanδ of Nb + In co-doped TiO2 with several co-dopant concentrations, which are potted as a function of frequency (https://doi.org/10.1038/nmat3691)

Fig. 13.17
A 3 D structure of electron pinned defect dipole.

The structure of the electron-pinned defect-dipole (EPDD), which is obtained by first-principles calculations. (https://doi.org/10.1038/nmat3691)

Fig. 13.18
An S E M image that appears like a mosaic. A graph plots current versus voltage. It is divided into four quadrants. It gives values for 1 to 2, 1 to 3, and 3 to 4 electrodes.

a The scanning electron microscopy image of Nb + In co-doped TiO2 ceramics, where the bright spots are electrodes. b IV properties measured with the electrodes of 1–2, 1–3, and 3–4 (https://doi.org/10.1038/srep08295)

Despite the negative experimental results for the EPDD concept, dielectric measurements in the low-temperature region, in which the semiconducting property of co-doped TiO2 is suppressed, suggests that the concept of EPDD itself works, although the enhancement of permittivity is not as large as initially reported. Figure 13.19 presents the (a) real and (b) imaginary parts of the dielectric permittivity of NITO-0.5% single crystal along the crystallographic c-direction, which are measured as a function of temperature and frequency. At 30 K, NITO-0.5% shows a permittivity on the order of 104 over the frequency range from 103 to 106 Hz, indicating that the sample remains semiconducting in this temperature region. As the temperature decreases, the real part of the permittivity steeply falls, as in the case of CCTO. The decrease in permittivity occurs from the higher frequency and accompanies a peak-like anomaly in the imaginary part of the permittivity, indicating a typical dielectric relaxation for the BLC effect. Figure 13.19c shows the Arrhenius plot for the observed dielectric relaxation, in which the peak frequency of the imaginary part of the permittivity is plotted as a function of temperature inverse on a single logarithmic scale. The experimental results are found to follow the Arrhenius low with an activation energy of 7.6 meV for the carrier excitation. At the lowest temperature region of approximately 2 K, where the thermally excited carriers are sufficiently suppressed to recover the fine insulation of the sample, the real part of the permittivity decreases to the order of 102. The dielectric permittivity of NITO-0.5%, which is measured along the [001] and [110] directions at 2 K, is presented in Fig. 13.20 with the results of the pristine TiO2. As presented in the figure, NITO-0.5% shows a permittivity of approximately 900 over a wide frequency range, from 102 to 106 along [001]. This value is more than three times greater than that for the pristine TiO2, indicating the intrinsic permittivity enhancement due to the co-doping of Nb and In, although its magnitude is not as large as the order of 104 that was originally reported in Ref. [40]. Along the direction of [110], on the other hand, the permittivity of NITO-0.5% reaches 250, whereas it is 130 for the pristine TiO2. The enhancement of permittivity is found to be more remarkable along the [001] direction than along [110]. This anisotropy suggests the anisotropic polarization property of the co-doping-induced EPDD in TiO2.

Fig. 13.19
Three graphs depicts N I T O 0.5 percent 001. a and b plot frequency versus temperature. c plots relaxation frequency versus 1 by T per Kelvin. The value of E a equals 7.6 m e V.

a The real and b imaginary parts of dielectric permittivity of Nb + In co-doped TiO2, which are potted as a function of temperature and frequency. c The Arrhenius plot for the dielectric relaxation, which is observed in the low-temperature region

Fig. 13.20
A graph plots dielectric permittivity versus frequency. The graph is labeled as N I T O 0.5 percent 001.

The frequency dispersion of the dielectric permittivity of Nb + In co-doped TiO2, which is measured along the [001] and [110] directions at 2 K. The broken lines in the figure denote the dielectric permittivity of the pristine TiO2 at 4 K

Although the expected configuration of EPDD was reported based on first-principles calculations, its real structure has remained an open question. The EPDDs form through the weak doping of heterovalent ions in the host matrix. Their structure is expected to depend strongly not only on the choice of the dopants and the crystal structure and physical properties of the host materials, but also on the synthesis conditions and concentration ratio of dopants. In Nb + In co-doped TiO2, for instance, the amounts of Nb5+ and In3+ are nominally chosen to be the same to realize the average valence of 4+ that is the same as Ti4+ in TiO2. If this condition is perfectly satisfied in the system, the co-doped sample is expected to remain insulating and no BLC effect will appear. The observed huge permittivity thus indicates a partial nonstoichiometry and/or segregation of the dopants. In this situation, the structure of EPDD would be different from that calculated under the assumption of homogeneity and perfect stoichiometry. Nevertheless, EPDD is undoubtedly one of the fascinating concepts contributing to the development of novel dielectric materials with high permittivity in terms of HOS engineering.

5 Summary

This section provides an overview of permittivity engineering in dielectric materials. There are several concepts according to the range of permittivity: the additivity rule of polarizability for permittivities of 100–101, the ferroelectric phase transition for permittivities of 102–103, and the special inhomogeneities for permittivities greater than 104, including the relaxors and the coexistence of multiple components with different electric properties that cause the BLC effect. The new concept of local structure engineering for permittivity boosting, EPDD, was also introduced in reference to recent studies. If we envisage the HOSs as extended local structures to enhance the permittivity, which are induced by synergy between the local and the bulk properties, the PNR in the relaxors can come under this category. Active explorations and investigations of EPDD with state-of-art experimental techniques and large-scale calculations are now expanding the frontier of functional HOSs in dielectric materials. Although not discussed in this section, modified structures around twin and anti-phase boundaries would provide a rich prospect for HOSs with exotic properties. Not only in ground states, field-induced local structures such as photo-excited defect states are also potential targets for the development of HOSs of excellent functionalities that are useful for practical applications.