We introduce an equivalent-circuit element based on the theory of interface pinning in random systems to analyze the contribution of domain wall motion below the coercive field to the impedance of a ferroelectric, as a function of amplitude E0 and frequency f of an applied ac electric field. We demonstrate our model on a bulk PbZrxTi1xO3 (PZT) reference sample and then investigate capacitor stacks, containing ferroelectric 0.5(Ba0.7Ca0.3)TiO30.5Ba(Zr0.2Ti0.8)O3 (BCZT) thin films, epitaxially grown by pulsed laser deposition on Nb-doped SrTiO3 single crystal substrates and covered with Au electrodes. Impedance spectra from f=10 Hz to 1 MHz were collected at different E0. Deconvolution of the spectra is achieved by fitting the measured impedance with an equivalent-circuit model of the capacitor stacks, and we extract for E0=2.5 kV/cm, a frequency-dependent permittivity of εr(f)=458+7.3ln(1Hz/2πf) for the BCZT films from the obtained fit parameters. From an extended Rayleigh analysis, we obtain a coupling strength of 0.187 cm/kV between dielectric nonlinearity and dielectric dispersion in the BCZT films and identify different domain-wall-motion regimes. Finally, we construct a schematic diagram of the different domain-wall-motion regimes and discuss the corresponding domain-wall dynamics. Our approach can be utilized to replace purely phenomenological constant phase elements (CPEs) in modeling the impedance response of ferroelectrics and extracting material properties.

Domain wall motion in ferroelectric materials does not only influence their ferroelectric but also their dielectric and piezoelectric properties.1 In general, the complex dynamics of domain wall displacement in ferroelectrics results in a dependence of their dielectric (εij) and piezoelectric (dijk) tensor components on the amplitude E0 and frequency f of an applied ac electric field.1,2

Regarding perovskite-type ferroelectrics, field and frequency dependent dielectric and piezoelectric response has been mostly investigated in ceramics and thin films based on PbZrxTi1xO3 (PZT).1–4 It was found that the field dependence of permittivity and piezoelectric coefficients at subswitching fields can often be described by the linear Rayleigh law, which results in a nonlinear and hysteretic dielectric polarization and piezoelectric strain. Based on this, the linear field dependence of permittivity is also referred to as dielectric nonlinearity.4 Moreover, the coupling of dielectric nonlinearity and frequency dispersion has been studied.2 However, a unified theory which describes the field and frequency behavior of material properties, e.g., the complex relative permittivity εr(E0,f) of ferroelectrics, with the same set of equations is not available. Thus, the measurement of εr(E0,f) is crucial since the permittivity is directly related to the microstructure of the material and essential for device modeling.1,5 This is especially true for thin film samples of new and, thus, less investigated lead-free ferroelectrics, such as K0.5Na0.5NbO3 (KNN) and 0.5(Ba0.7Ca0.3)TiO30.5Ba(Zr0.2Ti0.8)O3 (BCZT).6,7 In terms of device applications, the motion of ferroelectric domain walls at subcoercive fields is the key mechanism for the realization of ferroelectric memristors based on ferroelectric tunnel junctions (FTJs).8,9

Recently, a new method for the analysis of the field dependence of complex relative permittivity εr(E0) due to the motion of ferroelectric domain walls at subcoercive fields (i.e., Rayleigh analysis) was introduced.10 This method is based on impedance spectroscopy and equivalent-circuit fitting, which allows one to discriminate different contributions to the permittivity, originating from electronically different components (electrodes, electrode/ferroelectric interfaces, and ferroelectric) of the analyzed thin film devices. This is important, because the interfaces to the electrodes play a more crucial role in thin films as compared to thick ceramic samples,10,11 which can lead to extrinsic dielectric relaxation processes,12 masking the contributions from the ferroelectric. The superiority of the Rayleigh analysis based on impedance spectroscopy over the conventional approach (i.e., without equivalent-circuit fitting) was demonstrated in Ref. 10 for the measurement of εr(E0) for a ferroelectric Si:HfO2 thin film. However, the frequency dispersion εr(f) is not accessible with the method described in Ref. 10 because the utilized circuit-element is not based on a physical theory which describes the pinning of ferroelectric domain walls. Note that also the widely utilized constant phase elements (CPEs) are phenomenological circuit elements and not based on a physical model, which makes it difficult to extract physically meaningful parameters from equivalent-circuit fits with CPEs.

In thin film capacitor stacks, with a ferroelectric placed between two metallic electrodes, the dielectric properties can exhibit strong frequency dispersion depending on the resistivity of both, the ferroelectric film and the electrode layers,13 thus masking the true dielectric dispersion of the ferroelectric caused by pinning of domain walls on randomly distributed defects occurring in the ferroelectric layer inside the heterostructure.14,15 Thus, a circuit-element is needed which model the ferroelectric impedance response and hence allows one to extract material properties from equivalent-circuit fitting.

In this work, we introduce an equivalent-circuit element with impedance ZDW which is based on the theory of interface pinning in random systems14 and models the frequency-dependent impedance response of a ferroelectric including the domain-wall-motion induced contribution. Moreover, we report on a method to extract the domain-wall-motion induced amplitude- and frequency-dependent dielectric response εr(E0,f) in thin film ferroelectrics, embedded in a thin film capacitor stack, from impedance spectroscopy measurements at different E0 and equivalent-circuit fits with the introduced element ZDW. Finally, we present an extended Rayleigh analysis which allows one to quantify the coupling strength between dielectric nonlinearity and frequency dispersion and the identification of different domain-wall-motion regimes. Our approach will be demonstrated by applying it to thin film capacitor stacks containing the lead-free ferroelectric BCZT and, as a reference sample, to a commercial soft bulk PZT ceramic.

Domain wall dynamics below the coercive field EC of a ferroelectric can be considered as a superposition of center of mass motion and relaxational motion of internal modes, i.e., domain wall segments.15 From theory it is known, that the interaction of domain walls with pinning centers, that form a random energy landscape, gives rise to the so called Rayleigh behavior and the corresponding Rayleigh law.16,17 Rayleigh behavior originates from the irreversible center of mass motion of domain walls and results in a linear dependence of permittivity and piezoelectric coefficients on the field amplitude E0 above a specific threshold field ET and below the coercive field EC, which was also observed experimentally on ceramic and thin film samples.4,18–20 As a consequence of Rayleigh behavior, the dielectric polarization and piezoelectric strain responses of ferroelectrics become nonlinear in E0 and hysteretic even at subswitching fields ET<E0<EC. In the following, we will solely discuss dielectric properties and disregard piezoelectric properties.

The Rayleigh law for the real and imaginary parts of the complex relative permittivity εr=εriεr is given by Refs. 10 and 20,

(1)

Here, εr,in and εr,in are real and imaginary parts of the reversible (initial) Rayleigh parameter εr,in due to lattice and reversible domain wall contributions, α denotes the irreversible Rayleigh parameter (Rayleigh constant), and α=43πα.

In the following, we will denote a general linear field dependence of εr(E0) and εr(E0) with positive slope as “Rayleigh-like behavior.” In that sense, the Rayleigh law is a special case of Rayleigh-like behavior with the additional requirement α=43πα.

The irreversible domain wall contribution to the real part of the permittivity is given by αE0. It turns out that α is a direct and quantitative measure for the mobility of domain walls, and it is dependent on the concentration of pinning centers, the film thickness, and the density of domain walls.4 However, Rayleigh law Eq. (1) is rate-independent and no frequency dependence of material properties can be derived from it.2–4 

Theory predicts for domain wall pinning processes in random systems a logarithmic dependence of permittivity on frequency f, which is due to the relaxational contribution of internal modes,14,15

(2)

Here, Θ is related to the roughness of the domain wall and t0 denotes a microscopic time constant, which is related to the rate of transitions of domain wall segments between metastable states.14,15

In this pinned regime, the center-of-mass of the domain wall is captured in a potential valley, while domain wall segments can still jump between different metastable states with close energies. The potential valley of a random energy landscape can be described by Ref. 15,

(3)

with a constant Cn, the transverse spatial coordinate r¯, a single relaxation time τ1, and the random pinning force g. The first term in Eq. (3) describes the average (parabolic) shape of the well and the second term describes spatial fluctuations due to the residual part of the random pinning force g. This term is important because it results in a distribution of relaxation times Ψ1(τ) instead of the single relaxation time τ1.15 As a consequence, the dielectric response of a ferroelectric due to domain wall pinning in a random energy landscape is not expected to exhibit a Debye-like dielectric relaxation with a single relaxation time and, hence, one dominant relaxation frequency.

In many ferroelectrics—including relaxors—it was found that Θ1, which corresponds to a frequency-independent imaginary part of permittivity, and is valid for systems with a broad distribution Ψ1(τ) of relaxation times.2–4,21–23 In this case, the logarithmic frequency dependence of the complex relative permittivity can be well approximated by the following expression:2,3,22,23

(4)

where εr,0 and εr,0 denote static contributions to the permittivity and Δεr is the logarithmic dispersion strength. We note that in Eq. (4), we fixed the value of t0 to 1 s, which is also the common choice in the literature.4,21–24 We will use this convention throughout the remaining part of this paper.

It has been shown for systems with Rayleigh-like behavior and logarithmic frequency dispersion that the frequency dependence of εr in Eq. (4) is due to the logarithmic frequency dependence of both reversible (εr,in) and irreversible (α) Rayleigh parameters defined in Rayleigh law Eq. (1), which can then be expressed as21 

(5)

The physical interpretation of the coefficients Δεr,in and Δα in Eq. (5) will be discussed later in Sec. IV. As a consequence, the Rayleigh relation for the real part of relative permittivity εr in Eq. (1) can be written with frequency-dependent Rayleigh parameters given by Eq. (5) as

(6)

However, Rayleigh law Eq. (1) comprises also the Rayleigh relation for the imaginary part of relative permittivity εr and the relations are coupled by α=43πα. The coupling relation predicts a logarithmic frequency dependence of α which results in a frequency dispersion of εr, which is contradictory to the assumption of a frequency independent imaginary part given by Eq. (4), which implies a constant α. Thus, the logarithmic frequency dispersion due to domain wall pinning in random systems and the Rayleigh law are only compatible at a fixed frequency. Taking account of a constant α, we can implicitly define this frequency ffR by

(7)

At f=fR, the Rayleigh law is fulfilled for variable field amplitudes E0 within the Rayleigh region. The frequency fR will be called the “Rayleigh frequency” in the following.

The more general Rayleigh-like behavior exists in the frequency range where α(f) is positive. At frequencies f within this Rayleigh-like frequency range, the center of mass of domain walls can move irreversibly via hopping between different minima of the energy landscape since this is a basic assumption in the derivation of the Rayleigh law. Thus, the center of mass of the domain walls is not captured in a potential valley as in the pinned regime and we conclude that the Rayleigh-like frequency range corresponds to a regime with coexisting center of mass motion and relaxational motion of internal modes. In analogy to Ref. 15, we will denote this regime as the stochastic regime of complex relative permittivity.

In an impedance spectroscopy experiment, the complex impedance Z(f)=Z(f)iZ(f) is measured as a function of frequency f, typically in the range of f=10106 Hz. The complex impedance is related to the complex relative permittivity by Ref. 25,

(8)

with the geometrical capacitance C0. For a parallel-plate capacitor, C0=ε0Ad with the vacuum permittivity ε0, electrode area A, and plate distance d. Separating real and imaginary parts in Eq. (8) yields

(9)

By inserting the logarithmic frequency dependence Eq. (4) into Eq. (9), we obtain the equivalent-circuit element ZDW, which models the impedance of a ferroelectric due to pinning of domain walls in a potential valley [cf. Eq. (3)] of a random energy landscape,

(10)

with εr,0, εr,0 and Δεr as free parameters. Consequently, ZDW will be called the “domain wall pinning element” in the following.

Impedance spectroscopy allows one to extract the free parameters εr,0, εr,0, and Δεr by fitting the measured impedance over the full frequency range to the impedance of an appropriate equivalent-circuit model of the capacitor stack, including the ferroelectric. Thus, the different contributions (from electrodes, interfaces, and the ferroelectric layer) to the measured impedance of the capacitor stack are deconvoluted by equivalent-circuit fitting with the domain wall pinning element. The deconvolution of impedance data by equivalent-circuit fitting with the domain wall pinning element will be demonstrated below for a bulk PZT reference sample before we apply this to a capacitor stack which includes an epitaxial BCZT film.

Measurements at different excitation field amplitudes E0 then allow one to determine the field dependence of the dispersion parameters εr,0(E0), εr,0(E0), and Δεr(E0). Subsequently, the field- and frequency-dependent dielectric response εr(E0,f) of the ferroelectric is obtained by inserting the field-dependent dispersion parameters into Eq. (4). For comparison, the conventional approach for the analysis of complex relative permittivity measurements is described in  Appendix A.

Note that for systems in which the assumption of a broad distribution of relaxation times and/or Θ1 is not fulfilled, similar equivalent-circuit elements can be obtained by inserting the corresponding frequency dispersion into Eq. (9).

A ceramic BCZT target was used to grow 200-nm-thick epitaxial single crystalline films on (001)-oriented Nb-doped (0.5 wt%) SrTiO3 (Nb:STO) single crystal substrates (SurfaceNet GmbH, Rheine, Germany) by means of pulsed laser deposition (PLD). We used a 248 nm KrF excimer laser with an energy of 110 mJ and 5 Hz repetition rate. For the deposition on Nb:STO substrates, the temperature was fixed at 700°C with an oxygen partial pressure pO2=0.1mbar. After deposition, the films were heated up to 800°C and kept there for 15 min before cooling down to 700°C. Subsequently, pO2 was increased to 5 mbar and the films were cooled down to room temperature with a rate of 10 K/min.

The microstructure of the films was investigated by x-ray diffraction (XRD) to obtain the crystallographic orientation and to check for possible secondary phases. One sample was further analyzed by scanning transmission electron microscopy (STEM) to investigate possible crystal defects in BCZT, which may act as domain wall pinning-centers.

For electrical measurements, 3 mm diameter disk-shaped Au electrodes (100-nm-thick) were deposited on top of the BCZT films by electron beam evaporation through a shadow mask. Impedance spectra of the capacitor stack from f=10 Hz to 1 MHz were collected at different excitation field amplitudes E0 using a Solartron 1260 impedance analyzer together with a probe station. For each measurement, E0 was successively increased from 2.5 to 25 kV/cm (applied root mean square voltage VRMS=0.050.5 V) with a step-size of 1.25 kV/cm, which results in the measurement of the impedance under 19 different E0. Furthermore, impedance spectroscopy measurements were carried out on a 1.6 mm-thick disk-shaped commercial soft PZT ceramic with an electrode diameter of 8.06 mm (with a different impedance analyzer as utilized for the BCZT films).

In this section, we present XRD data and results from electrical measurements from one representative epitaxial BCZT device. The STEM results were obtained from a different epitaxial BCZT device which was fabricated with the same process parameters (see Sec. III).

Figure 1 shows the XRD θ–2θ diffraction pattern for a BCZT thin film deposited on a (001)-oriented Nb:SrTiO3 (Nb:STO) single crystal. No traces of secondary phases are observable within the scanning range. However, we cannot rule out traces of secondary phases at higher angles of 2θ between 60° and 100°. The structural phase diagram of BCZT shows four possible crystallographic phases, i.e., cubic, tetragonal, orthorhombic, and rhombohedral.26 Due to the small variations in the lattice parameters of the possible phases, a rigorous specification of the crystallographic phase of the films is not accessible from the XRD data. Thus, we denote the film structure as pseudo-cubic, which represents the four possible phases. For simplicity, we use cubic metrics for the Miller indices and crystal directions. The film grown on a Nb:STO substrate is epitaxial and the full width half maximum (FWHM) of the rocking curve of the (002)-BCZT peak (the inset in Fig. 1) has a value of ω=0.08°, indicating high crystalline quality of the film. The results are in good agreement with BCZT thin films grown by PLD on undoped STO27 and SrRuO3 coated STO.28 

FIG. 1.

XRD θ–2θ scan of the BCZT thin film on Nb:STO. The inset shows the rocking curve of the (002) BCZT peak.

FIG. 1.

XRD θ–2θ scan of the BCZT thin film on Nb:STO. The inset shows the rocking curve of the (002) BCZT peak.

Close modal

The epitaxial growth of BCZT on Nb:STO was also confirmed by STEM. Moreover, our STEM analysis revealed the occurrence of edge-type misfit dislocations at the epitaxial Nb:STO/BCZT interface, characterized by the Burgers vector b=a[100]. Here, a denotes the lattice constant along the [100] direction within the BCZT unit cell (Fig. 2). These misfit dislocations act as possible domain wall pinning centers, which was already observed in similar epitaxial PZT heterostructures.29 

FIG. 2.

Cross-sectional STEM image of the Nb:STO/BCZT interface viewed along the [010] zone axis. The enlarged section shows an edge-type misfit dislocation with Burgers vector b=a[100].

FIG. 2.

Cross-sectional STEM image of the Nb:STO/BCZT interface viewed along the [010] zone axis. The enlarged section shows an edge-type misfit dislocation with Burgers vector b=a[100].

Close modal

We first utilized impedance data, collected on a soft bulk PZT ceramic reference sample, to validate the domain wall pinning element ZDW [cf. Eq. (10)]. The PZT reference sample is well-known to exhibit the characteristic logarithmic frequency-dispersion due to domain-wall pinning.21Figure 3 shows the measured impedance response of the PZT reference sample in a Bode plot.30 First, the full equivalent-circuit depicted in the inset of Fig. 3 was used to fit the data.31 Here, the series resistance R0 corresponds to the resistance of the contacts and electrodes. The R1C1-element represents the interfaces to the electrodes. The resistance R2 and ZDW model the ferroelectric PZT layer. R2 is attributed to losses due to possible leakage currents, whereas the losses due to domain wall motion are contained in ZDW. Additionally, the dispersion parameters are contained in the domain wall pinning element. The element L0 models the inductance of the cables used.30 The fitting procedure yields values for R1 and C1 in the range of 1011Ω and 1015 F, respectively, and the other fit parameters remain unaffected by setting R1 and C1 to infinity. This corresponds to a vanishing interface contribution to the impedance response, as expected for a mm-thick bulk sample.10 Additionally, no change in the fit parameters could be observed for setting L00. Fitting with the remaining equivalent-circuit of the PZT reference sample yields the following values for the fit parameters: R0=77Ω, R2=18GΩ, εr,0=1400, εr,0=24.5, and Δεr=15.9. Setting further R00 and R2, only changes εr,0 slightly (25.1 instead of 24.5). Thus, the full equivalent-circuit depicted in the inset of Fig. 3 can be reduced to ZDW. The obtained fit with ZDW [cf. Eq. (10)] is depicted as solid lines in Fig. 3. The fit is in good agreement with the measured impedance. However, one might argue that it is not surprising that our model works with bulk PZT, as one can expect that in this kind of sample, the domain wall motion dominates over other extrinsic mechanisms. Hence, to further verify the correct extraction of material properties by domain wall pinning element modeling, we numerically added the impedance of circuit elements R0=1000Ω, R1=5000Ω, C1=100 pF, and L0=1 mH to the measured impedance of the PZT ceramic as depicted in the inset of Fig. 4(a). The added circuit elements represent extrinsic contributions to the impedance response, which may arise from contacts and interfaces.10,11,13 The added circuit elements result in a radically different behavior of the impedance at frequencies >1 kHz, as compared to the impedance measured on the PZT ceramic reference sample [cf. Fig. 4(a)]. Thus, converting the impedance data into complex permittivity data using the standard procedure, Eq. (8) also results in a radically different behavior of the permittivity as can be seen in Fig. 4(b). Here, the real part of the permittivity of the PZT sample with added elements even becomes negative at high frequencies, which is by no means a reflection of the true permittivity of the PZT reference sample. However, the extracted permittivity from the equivalent circuit fit to experimental data plus added elements is in good agreement with the measured permittivity of the PZT ceramic [cf. Fig. 4(b)], and the obtained fit parameters εr,0=1400, εr,0=24.5, and Δεr=15.9 are almost identical to the parameters obtained by fitting the impedance of the reference PZT sample with ZDW in Fig. 3. Moreover, the fitting procedure yields fit parameters of R0=1066Ω, R1=5040Ω, C1=103 pF, and L0=1 mH which are in very good agreement with the numerically added values. Altogether, the data presented in Fig. 4 demonstrate the correct deconvolution of different contributions to the measured impedance response by domain-wall pinning element modeling.

FIG. 3.

Logarithmic Bode plot of impedance modulus |Z| and phase angle obtained by impedance spectroscopy (symbols) on the soft bulk PZT reference sample. The equivalent circuit in the inset was used to fit the data and can be reduced to a fit with the domain wall pinning element ZDW (solid lines). The obtained fit parameters are εr,0=1400, εr,0=25.1, and Δεr=15.9.

FIG. 3.

Logarithmic Bode plot of impedance modulus |Z| and phase angle obtained by impedance spectroscopy (symbols) on the soft bulk PZT reference sample. The equivalent circuit in the inset was used to fit the data and can be reduced to a fit with the domain wall pinning element ZDW (solid lines). The obtained fit parameters are εr,0=1400, εr,0=25.1, and Δεr=15.9.

Close modal
FIG. 4.

(a) Logarithmic Bode plot of impedance modulus |Z| and phase angle measured on the PZT ceramic sample (blue symbols) and impedance data obtained by numerically adding circuit elements R0=1000Ω, R1=5000Ω, C1=100 pF, and L0=1 mH to the measured impedance of the PZT ceramic (black symbols). The equivalent circuit in the inset was used to fit the data (solid lines). (b) Comparison of the complex relative permittivity εr(f) obtained by the conventional approach (symbols) and extracted by equivalent-circuit fitting (solid lines).

FIG. 4.

(a) Logarithmic Bode plot of impedance modulus |Z| and phase angle measured on the PZT ceramic sample (blue symbols) and impedance data obtained by numerically adding circuit elements R0=1000Ω, R1=5000Ω, C1=100 pF, and L0=1 mH to the measured impedance of the PZT ceramic (black symbols). The equivalent circuit in the inset was used to fit the data (solid lines). (b) Comparison of the complex relative permittivity εr(f) obtained by the conventional approach (symbols) and extracted by equivalent-circuit fitting (solid lines).

Close modal

Collected impedance spectroscopy data (for a selection of three different E0) on the epitaxial BCZT thin-film device are shown in Fig. 5(a). A validation procedure to rule out artifacts in the measured impedance spectra is given in  Appendix B. The equivalent circuit depicted in Fig. 5(b) [i.e., the same circuit as shown in the inset of Fig. 3(a)] was used to fit the collected impedance spectra. Results of the fits can be seen in the Bode plot shown in Fig. 5(a) (for three values of E0). In all cases, the fits are in good agreement with the measured data, indicating that the utilized equivalent circuit is a suitable model for the thin film capacitors. It should be noted that the full equivalent-circuit [cf. Fig. 5(b)] is necessary to fit the impedance data of the BCZT device in contrast to the PZT reference sample [cf. Fig. 3], which is due to the fundamental differences in the impedance response of thin-film and bulk samples.10 Fit parameters obtained from equivalent-circuit fits at three different excitation field amplitudes are summarized in Table I. The standard errors are calculated from maximum likelihood estimation.32 Note that the equivalent-circuit fitting procedure could also be utilized to fit impedance data measured with a fixed (small) excitation field amplitude E0 superimposed by a dc-bias voltage VDC. This measurement mode can be utilized, e.g., to analyze the properties of possible Schottky barriers at the ferroelectric–metal interfaces, by extracting the interfacial capacitance C1 from the equivalent-circuit fits at different dc-bias voltages. Here, the interfacial capacitance due to Schottky barrier formation is expected to be proportional to (VDC+Vbi)1/2, with the apparent built-in potential Vbi.33 However, impedance spectroscopy and equivalent-circuit fitting does not reveal the electronic transport mechanism (e.g., thermionic injection, space-charge limited conduction) across metal–ferroelectric interfaces. Thus, a comprehensive analysis of transport properties across metal–ferroelectric interfaces also requires current–voltage (I–V) measurements.

FIG. 5.

(a) Logarithmic Bode plot of impedance modulus |Z| and phase angle obtained by impedance spectroscopy (symbols) for a selection of different E0 on the epitaxial BCZT thin film capacitor. The equivalent circuit in (b) was used to fit the data (solid lines).

FIG. 5.

(a) Logarithmic Bode plot of impedance modulus |Z| and phase angle obtained by impedance spectroscopy (symbols) for a selection of different E0 on the epitaxial BCZT thin film capacitor. The equivalent circuit in (b) was used to fit the data (solid lines).

Close modal
TABLE I.

Fit parameters for the epitaxial BCZT capacitor stack at three different excitation field amplitudes E0. The estimated standard errors for R2 are below 1 kΩ.

E0 (kV/cm)R0 (Ω)R1 (Ω)C1 (nF)R2 (MΩ)ɛr,0Δɛrɛr,0(μH)
2.5 18 ± 1 2889 ± 4 150.5 ± 0.3 29 457.9 ± 0.2 7.27 ± 0.04 10.65 ± 0.03 4.8 ± 0.8 
12.5 16 ± 1 2880 ± 3 151.8 ± 0.2 21 493.7 ± 0.1 8.41 ± 0.03 13.23 ± 0.02 5.3 ± 0.5 
22.5 14 ± 3 2838 ± 7 155.1 ± 0.6 13 526.4 ± 0.4 11.41 ± 0.09 18.23 ± 0.07 5.3 ± 1.4 
E0 (kV/cm)R0 (Ω)R1 (Ω)C1 (nF)R2 (MΩ)ɛr,0Δɛrɛr,0(μH)
2.5 18 ± 1 2889 ± 4 150.5 ± 0.3 29 457.9 ± 0.2 7.27 ± 0.04 10.65 ± 0.03 4.8 ± 0.8 
12.5 16 ± 1 2880 ± 3 151.8 ± 0.2 21 493.7 ± 0.1 8.41 ± 0.03 13.23 ± 0.02 5.3 ± 0.5 
22.5 14 ± 3 2838 ± 7 155.1 ± 0.6 13 526.4 ± 0.4 11.41 ± 0.09 18.23 ± 0.07 5.3 ± 1.4 

Subsequently, the complex relative permittivity εr(f) of the BCZT film (i.e., without contributions from the electrodes and ferroelectric/electrode interfaces) was calculated from the obtained fit parameters according to Eq. (4) for each field amplitude E0. Due to domain wall pinning element modeling, the extracted frequency-dependent permittivity of the BCZT layer is in agreement with the theory of interface pinning in random systems.14 This is different to previous studies on the dielectric dispersion in ferroelectrics, where purely phenomenological CPEs were utilized to extract permittivity data.11,12,34,35 The results for the extracted BCZT permittivity for a selection of three different excitation field amplitudes are depicted in Fig. 6. We clearly observe [cf. Fig. 6 and Table I] that εr(f) and the parameters εr,0, εr,0, and Δεr depend on the field amplitude E0, which indicate that dielectric nonlinearity and frequency dispersion are coupled in the BCZT films.

FIG. 6.

Complex relative permittivity εr(f) of the BCZT film for a selection of three different excitation field amplitudes E0.

FIG. 6.

Complex relative permittivity εr(f) of the BCZT film for a selection of three different excitation field amplitudes E0.

Close modal

From the equivalent-circuit fits, the field dependence of the dispersion parameters εr,0(E0), εr,0(E0), and Δεr(E0) can be extracted and the results are depicted in Fig. 7.

FIG. 7.

Field dependence of the dispersion parameters εr,0(E0), εr,0(E0), and Δεr(E0) of the BCZT film. The solid lines are linear fits in the region between the two vertical dashed lines, which mark the threshold field ET and the partial switching field EPS.

FIG. 7.

Field dependence of the dispersion parameters εr,0(E0), εr,0(E0), and Δεr(E0) of the BCZT film. The solid lines are linear fits in the region between the two vertical dashed lines, which mark the threshold field ET and the partial switching field EPS.

Close modal

We observe that all dispersion parameters exhibit a linear dependence on the field amplitude E0, i.e., in the guise of the following linear equations:

(11)

To specify the coefficients in Eq. (11), we insert Eq. (11) into Eq. (4), which yields

(12)

Comparing the real part εr(E0,f) of Eq. (12) with Eqs. (5) and (6) gives rise to the identifications a0εr,in,0, b0Δεr,in, aα0, and bΔα. It should be noted that similar identifications were previously demonstrated to be valid for a PZT-based ferroelectric ceramic.21 However, the imaginary part of the material response was not considered in Ref. 21. Here, comparison of the frequency-independent imaginary part in Eq. (12), with the imaginary part of the Rayleigh law Eq. (1), yields the additional identifications a0εr,in and aα.

Altogether, the field dependence of the dispersion parameters Eq. (11) can then be written in the following form:

(13)

where the coefficient Δεr,in is the zero-field contribution to the logarithmic dispersion strength and the field-dependent contribution is given by ΔαE0.

Subsequently, linear fits according to Eq. (13) were carried out in the range 6.25kV/cmE016.25kV/cm as indicated by the vertical dashed lines in Fig. 7. The starting point for the linear fits at E0=6.25kV/cmET was determined by the criterion that all subsequent data points are required to have a higher value compared to their left neighbor. The end point for the linear fits was then chosen by the longest fitting range for which every linear fit has a correlation factor R20.980, which is fulfilled for E0=16.25kV/cmEPS. At higher fields, all three curves in Fig. 7 show an upward curvature which we attribute to the partial switching of domains.20 For the linear fits, the estimated standard errors from the previous equivalent-circuit fitting procedure were taken into account, which is also indicated by the error bars within the fitting range [cf. Fig. 7]. The results of the fits according to Eq. (13) are summarized in Table II.

TABLE II.

Extracted parameters from the linear fits according to Eq. (13). The estimated uncertainties from previous fitting steps were taken into account.

ɛr,in,0α0′ (cm/kV)ɛr,inα″ (cm/kV)Δɛr,inΔα′ (cm/kV)
463.9 ± 0.1 2.41 ± 0.01 9.31 ± 0.02 0.318 ± 0.002 6.09 ± 0.03 0.187 ± 0.003 
ɛr,in,0α0′ (cm/kV)ɛr,inα″ (cm/kV)Δɛr,inΔα′ (cm/kV)
463.9 ± 0.1 2.41 ± 0.01 9.31 ± 0.02 0.318 ± 0.002 6.09 ± 0.03 0.187 ± 0.003 

From the extracted fit parameters, the functional form of the frequency-dependent reversible and irreversible Rayleigh parameter of the BCZT thin film can be immediately obtained according to Eq. (5) and the results are depicted in Fig. 8. Here, the vertical dotted line atfR1 kHz indicates the Rayleigh frequency for the BCZT thin film, which we calculated according to Eq. (7). As discussed in Sec. II, fR is part of the Rayleigh-like frequency range, where the center of mass of the domain walls can move irreversibly via hopping between potential minima corresponding to the stochastic regime.

FIG. 8.

Logarithmic frequency dispersion of reversible εr,in(f) and irreversible α(f) Rayleigh parameters of the BCZT film. The vertical dotted line at 1 kHz marks the Rayleigh frequency fR. The irreversible Rayleigh parameter changes to negative values at f063 kHz, which is indicated by the vertical dashed line.

FIG. 8.

Logarithmic frequency dispersion of reversible εr,in(f) and irreversible α(f) Rayleigh parameters of the BCZT film. The vertical dotted line at 1 kHz marks the Rayleigh frequency fR. The irreversible Rayleigh parameter changes to negative values at f063 kHz, which is indicated by the vertical dashed line.

Close modal

The dashed line at f063 kHz marks the region where the irreversible Rayleigh parameter α is zero and changes sign to negative values. This might indicate a transition from the stochastic regime to a pinned regime, in which the center of mass motion of the domain walls cannot follow the fast driving field, and, thus, the domain walls are captured in a potential valley.15 Similar experimental indications for the freezing of the irreversible center of mass motion of the domain walls at higher frequencies were reported previously.24 

By inserting the functional form of the field dependent frequency dispersion parameters [cf. Fig. 7 and Table II] into Eq. (4), we finally obtain the functional form of the field and frequency dependent dielectric response in the epitaxial BCZT thin film, which is given by

(14)

Note that εr,in in Eq. (14) is not zero, which could be an indication for an internal bias or stress field.10 

The coupling between field dependence and frequency dispersion in the BCZT thin film arises explicitly in Eq. (14) as the product of E0 and the f-dependent term. Thus, this gives rise to the identification of the coefficient Δα as the coupling strength between dielectric nonlinearity and frequency dispersion, which is defined as the change of logarithmic frequency dispersion strength with field amplitude—or equivalent—as the change of the irreversible Rayleigh parameter with the logarithm of frequency, i.e.,

(15)

which has a value of 0.187 cm/kV in the epitaxial BCZT thin film. In fact, the negative values of the irreversible Rayleigh parameter are a direct consequence of the coupling between frequency dispersion and dielectric nonlinearity in the BCZT thin film.

Altogether, our analysis yields the schematic diagram depicted in Fig. 9 for the different domain wall motion regimes in the epitaxial BCZT thin film. Note that the diagram constructed in Fig. 9 is not a generic diagram for any epitaxial BCZT film. In fact, it is a schematic diagram of different domain-wall motion regimes, which can now be identified within the framework of the extended Rayleigh analysis.

FIG. 9.

Schematic diagram of the different domain wall motion regimes of the epitaxial BCZT thin film. The threshold field ET=6.25kV/cm and the partial switching field EPS=16.25kV/cm are obtained from the field dependence of the dispersion parameters [cf. Fig. 7] and f063 kHz is determined from the frequency dispersion of the irreversible Rayleigh parameter [cf. Fig. 8].

FIG. 9.

Schematic diagram of the different domain wall motion regimes of the epitaxial BCZT thin film. The threshold field ET=6.25kV/cm and the partial switching field EPS=16.25kV/cm are obtained from the field dependence of the dispersion parameters [cf. Fig. 7] and f063 kHz is determined from the frequency dispersion of the irreversible Rayleigh parameter [cf. Fig. 8].

Close modal

At low field amplitudes 0<E0<ET, the center of mass of the domain walls cannot overcome the potential well Eq. (3) of the random energy landscape; however, domain wall segments can jump between metastable states with close energies, which corresponds to the non-coupled pinning regime. This observation has similarities to what has been found theoretically for disordered ferromagnets.36 By increasing the field amplitude above the threshold field ET, the center of mass of the domain walls can additionally jump between different potential minima, resulting in a coexistence of irreversible center of mass motion and relaxational motion of internal modes in the stochastic regime. Note that this coexistence is not simply additive due to the coupling of dielectric nonlinearity and frequency dispersion. As a consequence of coupling, the irreversible Rayleigh parameter becomes zero at f063 kHz and changes to negative values in the pinned coupling regime. Note that in this coupled-pinning regime, domain wall segments can still jump between metastable states, which is indicated by the logarithmic frequency dispersion observed in this regime. The domain wall dynamics in the coupled-pinning regime is not fully understood and remains as a task of a unified theory of dielectric nonlinearity and frequency dispersion, which is not yet available.

In conclusion, this work demonstrates several key aspects. We introduce the domain wall pinning element ZDW based on the theory of interface pining in random systems, which includes the characteristic logarithmic frequency dispersion due to domain wall pinning in a random energy landscape and hence models the impedance response of a ferroelectric below the coercive field. The domain wall pinning element can be utilized to replace purely phenomenological CPEs in modeling the impedance response of ferroelectrics and extracting material properties. Moreover, we demonstrate the practical application of this new element to extract the field- and frequency-dependent dielectric response in an epitaxial lead-free ferroelectric BCZT thin film embedded in a capacitor stack and we reveal its superiority over the conventional approach. In addition, we perform an extended Rayleigh analysis which results in the quantification of the coupling strength and the explicit functional form εr(E0,f) of the coupled dielectric response in the BCZT film. Finally, we present a schematic diagram of the different domain wall motion regimes in the BCZT film and discuss the corresponding domain wall dynamics.

The present work is intended to serve as a guideline for future work on ferroelectric materials, which includes to reveal the effect of temperature and phase transitions,37 external bias fields,4 dopants,23 ion-bombardment,38,39 and substrate clamping24 on domain wall dynamics and related material properties. In terms of ferroelectric devices, our approach may help to characterize and optimize the response of ferroelectric memristors, which is also of practical importance for future interdisciplinary applications of ferroelectrics in bioelectronics, such as ferroelectric microelectrodes40 in hybrid neurocomputational systems.

We acknowledge Dragan Damjanovic for providing data on a ceramic reference sample and gratefully acknowledge fruitful discussions with J. Driscoll (Cambridge), C. Warres (NMI), C. Hofer, and J. C. Meyer and technical support by M. Turad and R. Löffler (LISA+). This work was partially funded by the Bundesministerium für Bildung und Forschung (BMBF) under Grant No. 13GW0123E and by the Europäische Fonds für regionale Entwicklung (EFRE) under Grant No. 712303.

The authors have no conflicts to disclose.

Maximilian T. Becker: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Writing – original draft (lead); Writing – review and editing (equal). Claus J. Burkhardt: Funding acquisition (lead); Project administration (lead); Resources (equal). Reinhold Kleiner: Formal analysis (supporting); Writing – review and editing (supporting); Resources (equal). Dieter Koelle: Formal analysis (supporting); Writing – original draft (supporting); Writing – review and editing (equal); Resources (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

For the conventional approach, the measured complex impedance Z=ZiZ of the capacitor stacks is converted into complex permittivity data of the capacitor stacks using the formalism described in Ref. 25, resulting in

(A1)

The resulting εr,stack(f) plots are shown in Fig. 10 for a selection of three different E0. For comparison, the extracted complex permittivity εr(f) of the BCZT layer from the equivalent-circuit fits [cf. Fig. 6 and Eq. (4)] is depicted. The capacitor stack exhibits a Debye-like dielectric relaxation at high frequencies (105 Hz) due to resistive losses in the electrodes.13 Moreover, the dielectric spectrum of the epitaxial thin film capacitor indicates an additional Debye-like relaxation process at lower frequencies (100Hz). We attribute this second relaxation process to an extrinsic electrode/film interface effect, which was previously reported to occur in other ferroelectric thin films grown on Nb:STO substrates.12 This is also consistent with the theoretical discussion of potential well Eq. (3) within a random energy landscape, which leads to a distribution of relaxation times in the BCZT layer and hence the BCZT layer is not expected to exhibit a Debye-like dielectric relaxation which corresponds to a single relaxation time and hence one dominant relaxation frequency fr. The relaxation frequency can be calculated from the obtained interface fit parameters R1 and C1 according to the relation fr=1/2πR1C1. The calculated relaxation frequencies are in agreement with the peaks in the imaginary part of permittivity of the capacitor stack [cf. Fig. 10] at frequencies (100 Hz), which indicates the correct deconvolution of film and interface contributions by our equivalent-circuit fits. Moreover, for BCZT thin films grown on platinized Si substrates, the second Debye-like dielectric relaxation was absent.41 Furthermore, we note that the slightly negative values of εr,stack(f) at frequencies around 1 MHz are due to the inductance of the cables used.30 

FIG. 10.

Measured permittivities εr,stack(f) (full symbols) and εr,stack(f) (open symbols) of the BCZT capacitor stack for a selection of different excitation field amplitudes E0. For clarity, the symbols are connected by thin lines which do not represent fits to the data. For comparison, the extracted permittivities of the BCZT layer from the equivalent-circuit fits [cf. Fig. 6 and Eq. (4)] are also shown (solid and dashed thick lines).

FIG. 10.

Measured permittivities εr,stack(f) (full symbols) and εr,stack(f) (open symbols) of the BCZT capacitor stack for a selection of different excitation field amplitudes E0. For clarity, the symbols are connected by thin lines which do not represent fits to the data. For comparison, the extracted permittivities of the BCZT layer from the equivalent-circuit fits [cf. Fig. 6 and Eq. (4)] are also shown (solid and dashed thick lines).

Close modal

It is clear that the extraction of frequency dispersion of permittivity from [cf. Eq. (A1) and symbols in Fig. 10] does not reflect the true dielectric dispersion of the BCZT thin film since there is no deconvolution of electronically distinct components forming the capacitor heterostructure. The same is true for the extraction of Rayleigh parameters, which has been recently demonstrated.10 

Impedance spectra might be afflicted with artifacts. To rule out artifacts introduced by experimental setup conditions in the measured impedance spectra, we use the two-pole Hilbert transform (Z-HIT) described in Ref. 42. This approach is beneficial compared to methods based on Kramers–Kronig (KK) relations43,44 since the KK relations are strictly defined within the frequency range between zero and infinity and hence, the extrapolation of the measured impedance data is a unavoidable task which may lead to erroneous results.42 The Z-HIT avoids the necessity of an extrapolation of the frequency range and allows one to reconstruct the modulus of the impedance |Z| vs f from the measured values of the less artifact-prone10,42 phase angle. The results of the Z-HIT for three different excitation field amplitudes E0 are depicted in Fig. 11. In all cases, the reconstructed curves match fairly well with the measured data, which indicates that the measured impedance spectra are not afflicted with artifacts and the investigated system is stable, causal, and (sufficiently) linear10 and exhibits finite values over the entire frequency range.

FIG. 11.

Comparison of measured (symbols) and Z-HIT-reconstructed (black solid lines) impedance modulus |Z| of the BCZT capacitor stack vs frequency f for (a) E0=2.5 kV/cm, (b) E0=12.5 kV/cm, and (c) E0=22.5 kV/cm.

FIG. 11.

Comparison of measured (symbols) and Z-HIT-reconstructed (black solid lines) impedance modulus |Z| of the BCZT capacitor stack vs frequency f for (a) E0=2.5 kV/cm, (b) E0=12.5 kV/cm, and (c) E0=22.5 kV/cm.

Close modal
1.
D.
Damjanovic
, “
Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin films and ceramics
,”
Rep. Progr. Phys.
61
,
1267
(
1998
).
2.
D.
Damjanovic
,
S. S. N.
Bharadwaja
, and
N.
Setter
, “
Toward a unified description of nonlinearity and frequency dispersion of piezoelectric and dielectric responses in Pb(Zr, Ti)O3
,”
Mater. Sci. Eng. B
120
,
170
(
2005
).
3.
N.
Bassiri-Gharb
and
S.
Trolier-McKinstry
, “
Dielectric nonlinearity of Pb(Yb1/2Nb1/2)O3PbTiO3 thin films with (100) and (111) crystallographic orientation
,”
J. Appl. Phys.
97
,
064106
(
2005
).
4.
N.
Bassiri-Gharb
,
I.
Fujii
,
E.
Hong
,
S.
Trolier-McKinstry
,
D. V.
Taylor
, and
D.
Damjanovic
, “
Domain wall contributions to the properties of piezoelectric thin films
,”
J. Electroceram.
19
,
49
(
2007
).
5.
S.
Trolier-McKinstry
and
P.
Muralt
, “
Thin film piezoelectrics for MEMS
,”
J. Electroceram.
12
,
7
(
2004
).
6.
Y.
Saito
,
H.
Takao
,
T.
Tani
,
T.
Nonoyama
,
K.
Takatori
,
T.
Homma
,
T.
Nagaya
, and
M.
Nakamura
, “
Lead-free piezoceramics
,”
Nature
432
,
84
(
2004
).
7.
W.
Liu
and
X.
Ren
, “
Large piezoelectric effect in Pb-free ceramics
,”
Phys. Rev. Lett.
103
,
257602
(
2009
).
8.
A.
Chanthbouala
,
V.
Garcia
,
R. O.
Cherifi
,
K.
Bouzehouane
,
S.
Fusil
,
X.
Moya
,
S.
Xavier
,
H.
Yamada
,
C.
Deranlot
,
N. D.
Mathur
,
M.
Bibes
,
A.
Barthelemy
, and
J.
Grollier
, “
A ferroelectric memristor
,”
Nat. Mater.
11
,
860
(
2012
).
9.
S.
Boyn
,
J.
Grollier
,
G.
Lecerf
,
B.
Xu
,
N.
Locatelli
,
S.
Fusil
,
S.
Girod
,
C.
Carretero
,
K.
Garcia
,
S.
Xavier
,
J.
Tomas
,
L.
Bellaiche
,
M.
Bibes
,
A.
Barthelemy
,
S.
Saighi
, and
V.
Garcia
, “
Learning through ferroelectric domain dynamics in solid-state synapses
,”
Nat. Commun.
8
,
14736
(
2017
).
10.
T.
Schenk
,
M.
Hoffmann
,
M.
Pesic
,
M. H.
Park
,
C.
Richter
,
U.
Schroeder
, and
T.
Mikolajick
, “
Physical approach to ferroelectric impedance spectroscopy: The Rayleigh element
,”
Phys. Rev. Appl.
10
,
064004
(
2018
).
11.
R.
Schmidt
,
W.
Eerenstein
,
T.
Winiecki
,
F. D.
Morrison
, and
P. A.
Midgley
, “
Impedance spectroscopy of epitaxial multiferroic thin films
,”
Phys. Rev. B
75
,
245111
(
2007
).
12.
R.
Schmidt
,
J.
Ventura
,
E.
Langenberg
,
N. M.
Nemes
,
C.
Munuera
,
M.
Varela
,
M.
Garcia-Hernandez
,
C.
Leon
, and
J.
Santamaria
, “
Magnetoimpedance spectroscopy of epitaxial multiferroic thin films
,”
Phys. Rev. B
86
,
035113
(
2012
).
13.
M.
Tyunina
, “
Size effects and dielectric behaviour in ferroelectric heterostructures
,”
J. Phys.: Condens. Matter
18
,
5725
(
2006
).
14.
T.
Nattermann
,
Y.
Shapir
, and
I.
Vilfan
, “
Interface pinning and dynamics in random systems
,”
Phys. Rev. B
42
,
8577
(
1990
).
15.
A. A.
Fedorenko
,
V.
Mueller
, and
S.
Stepanow
, “
Dielectric response due to stochastic motion of pinned domain walls
,”
Phys. Rev. B
70
,
224104
(
2004
).
16.
H.
Kronmüller
, “
Statistical theory of Rayleigh’s law
,”
Z. Angew. Phys.
30
,
1
(
1970
).
17.
O.
Boser
, “
Statistical theory of hysteresis in ferroelectric materials
,”
J. Appl. Phys.
62
,
1344
(
1987
).
18.
D.
Damjanovic
, “
Stress and frequency dependence of the direct piezoelectric effect in ferroelectric ceramics
,”
J. Appl. Phys.
82
,
1788
(
1997
).
19.
D. V.
Taylor
and
D.
Damjanovic
, “
Evidence of domain wall contribution to the dielectric permittivity in PZT thin films at sub-switching fields
,”
J. Appl. Phys.
82
,
1973
(
1997
).
20.
D. A.
Hall
, “
Rayleigh behaviour and the threshold field in ferroelectric ceramics
,”
Ferroelectrics
223
,
319
(
1999
).
21.
D.
Damjanovic
, “
Logarithmic frequency dependence of the piezoelectric effect due to pinning of ferroelectric-ferroelastic domain walls
,”
Phys. Rev. B
55
,
R649
(
1997
).
22.
V.
Mueller
,
H.
Beige
, and
H. P.
Abicht
, “
Non-debye dielectric dispersion of barium titanate stannate in the relaxor and diffuse phase-transition state
,”
Appl. Phys. Lett.
84
,
1341
(
2004
).
23.
L.
Yang
,
H.
Fang
,
L.
Zheng
,
J.
Du
,
L.
Wang
,
X.
Lu
,
W.
,
R.
Zhang
, and
W.
Cao
, “
Evaluation of reversible and irreversible domain wall motion in relaxor ferroelectrics: Influence of acceptor ions
,”
Appl. Phys. Lett.
114
,
232901
(
2019
).
24.
F.
Griggio
,
S.
Jesse
,
A.
Kumar
,
O.
Ovchinnikov
,
T. N.
Jackson
,
D.
Damjanovic
,
S. V.
Kalinin
, and
S.
Trolier-McKinstry
, “
Substrate clamping effects on irreversible domain wall dynamics in lead zirconate titanate thin films
,”
Phys. Rev. Lett.
108
,
157604
(
2012
).
25.
R.
Gerhardt
, “
Impedance and dielectric spectroscopy revisited: Distinguishing localized relaxation from long-range conductivity
,”
J. Phys. Chem. Solids
55
,
1491
(
1994
).
26.
D. S.
Keeble
,
F.
Benabdallah
,
P. A.
Thomas
,
M.
Maglione
, and
J.
Kreisel
, “
Revised structural phase diagram of (Ba0.7Ca0.3TiO3)–(BaZr0.2Ti0.8O3)
,”
Appl. Phys. Lett.
102
,
092903
(
2013
).
27.
N. D.
Scarisoreanu
,
F.
Craciun
,
A.
Moldovan
,
V.
Ion
,
R.
Birjega
,
C.
Ghica
,
R. F.
Negrea
, and
M.
Dinescu
, “
High permittivity (1x)Ba(Zr0.2Ti0.8)O3x(Ba0.7Ca0.3)TiO3 (x=0.45) epitaxial thin films with nanoscale phase fluctuations
,”
ACS Appl. Mater. Interfaces
7
,
23984
(
2015
).
28.
Q.
Lin
,
D.
Wang
, and
S.
Li
, “
Strong effect of oxygen partial pressure on electrical properties of 0.5Ba(Zr0.2Ti0.8)O30.5(Ba0.7Ca0.3)TiO3 thin films
,”
J. Am. Ceram. Soc.
98
,
2094
(
2015
).
29.
D.
Su
,
Q.
Meng
,
C. A. F.
Vaz
,
M.-G.
Han
,
Y.
Segal
,
F. J.
Walker
,
M.
Sawicki
,
C.
Broadbridge
, and
C. H.
Ahn
, “
Origin of 90° domain wall pinning in Pb(Zr0.2Ti0.8)O3 heteroepitaxial thin films
,”
Appl. Phys. Lett.
99
,
102902
(
2011
).
30.
J. R.
Macdonald
and
E.
Barsoukov
,
Impedance Spectroscopy
, 3rd ed. (
Wiley
,
New York
,
2018
).
31.
A Wolfram Mathematica script to fit impedance data with an equivalent-circuit model containing the domain wall pinning element can be made available to interested readers upon reasonable request addressed to the corresponding author.
32.
P. G.
Hoel
,
Introduction to Mathematical Statistics
, 3rd ed. (
Wiley
,
New York
,
1962
).
33.
L.
Pintilie
and
M.
Alexe
, “
Metal-ferroelectric-metal heterostructures with Schottky contacts. I. Influence of the ferroelectric properties
,”
J. Appl. Phys.
98
,
124103
(
2005
).
34.
F. D.
Morrison
,
D. J.
Jung
, and
J. F.
Scott
, “
Constant-phase-element (CPE) modeling of ferroelectric random-access memory lead zirconate-titanate (PZT) capacitors
,”
J. Appl. Phys.
101
,
094112
(
2007
).
35.
X.
Vendrell
,
J.
Ramírez-González
,
Z.-G.
Ye
, and
A. R.
West
, “
Revealing the role of the constant phase element in relaxor ferroelectrics
,”
Commun. Phys.
5
,
9
(
2022
).
36.
T.
Nattermann
,
V.
Pokrovsky
, and
V. M.
Vinokur
, “
Hysteretic dynamics of domain walls at finite temperatures
,”
Phys. Rev. Lett.
87
,
197005
(
2001
).
37.
A.
Piorra
,
V.
Hrkac
,
N.
Wolff
,
C.
Zamponi
,
V.
Duppel
,
J.
Hadermann
,
L.
Kienle
, and
E.
Quandt
, “
(Ba0.85Ca0.15)(Ti0.9Zr0.1O3) thin films prepared by PLD: Relaxor properties and complex microstrucure
,”
J. Appl. Phys.
125
,
244103
(
2019
).
38.
S.
Saremi
,
R.
Xu
,
F. I.
Allen
,
J.
Maher
,
J. C.
Agar
,
R.
Gao
,
P.
Hosemann
, and
L. W.
Martin
, “
Local control of defects and switching properties in ferroelectric thin films
,”
Phys. Rev. Mater.
2
,
084414
(
2018
).
39.
S.
Saremi
,
J.
Kim
,
A.
Ghosh
,
D.
Meyers
, and
L. W.
Martin
, “
Defect-induced (dis)order in relaxor ferroelectric thin films
,”
Phys. Rev. Lett.
123
,
207602
(
2019
).
40.
M. T.
Becker
, “
Charge injection capacity of ferroelectric microelectrodes for bioelectronic applications
,”
AIP Adv.
11
,
065106
(
2021
).
41.
M.
Becker
,
C. J.
Burkhardt
,
B.
Schröppel
,
R.
Kleiner
, and
D.
Koelle
, “
Rayleigh analysis and dielectric dispersion in polycrystalline 0.5(Ba0.7Ca0.3)TiO30.5Ba(Zr0.2Ti0.8)O3 ferroelectric thin films by domain-wall pinning element modeling
,”
J. Appl. Phys.
128
,
154103
(
2020
).
42.
C. A.
Schiller
,
F.
Richter
,
E.
Gülzow
, and
N.
Wagner
, “
Validation and evaluation of electrochemical impedance spectra of systems with states that change with time
,”
Phys. Chem. Chem. Phys.
3
,
374
(
2001
).
43.
M.
Urquidi-Macdonald
,
S.
Real
, and
D. D.
Macdonald
, “
Applications of Kramers-Kronig transforms in the analysis of electrochemical impedance data—III. Stability and linearity
,”
Electrochim. Acta
35
,
1559
(
1990
).
44.
P.
Agarwal
,
M. E.
Orazem
, and
L. H.
Garcia-Rubio
, “
Application of measurement models to impedance spectroscopy III. Evaluation of consistency with the Kramers-Kronig relations
,”
J. Electrochem. Soc.
142
,
1885
(
1995
).