Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

How space-charge behaviour at grain boundaries in electroceramic oxides is modified by two restricted equilibria

A. L. Usler *, F. Ketter and R. A. De Souza
Institute of Physical Chemistry, RWTH Aachen University, 52056 Aachen, Germany. E-mail: usler@pc.rwth-aachen.de; desouza@pc.rwth-aachen.de

Received 1st December 2023 , Accepted 8th February 2024

First published on 16th February 2024


Abstract

Determining the space-charge potential at grain boundaries in oxides by various experimental methods bears the promise of providing a comprehensive, quantitative description of interfacial defect chemistry. In this study, we draw attention to the problem of unifying data measured in different temperature ranges. We focus on unifying data from elevated-temperature electrical methods, such as impedance spectroscopy and current–voltage measurements, with data from room-temperature imaging techniques, such as Scanning Probe Microscopy (SPM), Transmission Electron Microscopy (TEM), and Atom Probe Tomography (APT). By means of continuum simulations, we calculate the space-charge potential Φ0 at grain boundaries in the model electroceramic oxide acceptor-doped SrTiO3, taking into account, first, a restricted equilibrium that leads to frozen-in acceptor-dopant profiles, and subsequently, a restricted equilibrium that leads to frozen-in bulk oxygen-vacancy concentrations. Our results indicate non-trivial differences between experimental values of Φ0 obtained from electrical and from imaging methods, differences that arise from the different measurement temperatures and that are aggravated by the restricted equilibria. We also show that grain-boundary widths determined from elemental acceptor-cation profiles will not, on principle, agree with the electrical width extracted from impedance spectroscopy data.


1 Introduction

Grain boundaries are regions of perturbed lattice structure. Although the volume fraction of such regions may be very small, their presence in electroceramic oxides can influence, if not govern, the materials' overall behaviour. In polycrystalline samples of the model electroceramic oxide1 SrTiO3, for example, grain boundaries change the material's plasticity,2,3 alter its superconductive properties,4,5 give rise to varistor behaviour,6–8 facilitate resistive switching,9 and strongly increase its electrical resistance.7,8,10–16 The latter effect in particular is characteristic of grain boundaries across a variety of ionic and mixed ionic–electronic conducting electroceramics, such as systems based on CeO2,17–20 ZrO2,19,21–23 LaGaO3,24,25 BaZrO3,26–30 and BaCeO3.31,32

The high grain-boundary resistance in acceptor-doped SrTiO3 is commonly explained by the presence of attendant space-charge layers.7,11,12,33–35 Specifically, the grain boundaries are assumed to exhibit an excess positive charge that is compensated by adjacent negative space-charge layers in which the mobile positive charge carriers (oxygen vacancies and electron holes) are depleted. Such models readily explain why the grain-boundary effect persists in the absence of impurity phases.7,11,15,36–38 In addition, they predict that oxygen diffusion is hindered at grain boundaries;39,40 they explain why, at grain boundaries, oxide-ion currents are hindered more strongly than electronic currents;41,42 they predict the conductivity changes upon reduction to the nanoscale;38,43–46 and they can account for the varistor behaviour of SrTiO3 ceramics.7,8,47

The strength of a space-charge effect is typically quantified by the space-charge potential Φ0, that is, the difference in the electric potential between the grain boundary and the bulk. Electrical measurements yield Φ0 most easily from the ratio of the grain-boundary and bulk time constants14 obtained by means of impedance spectroscopy. It has also been proposed that Φ0 may be calculated from the nonlinearity of the current–voltage curve.48–50 A second characteristic of space-charge layers is their extent, [small script l]gb. In particular, the electrical grain-boundary width can be extracted from the ratio of grain-boundary and bulk capacitances obtained by impedance spectroscopy.

Imaging techniques can not only provide estimates of Φ0 and [small script l]gb, but also, in comparison with the electrical methods, direct access to the spatial profiles of the electric potential and of elemental concentrations. Scanning probe microscopy (SPM),51,52 for example, can be used to probe the electric potential profile across a grain boundary. Atom probe tomography (APT) measurements can provide an atomic-scale imaging of a grain boundary's chemical composition,53–55 from which the space-charge density can be obtained (and thus the space-charge potential). In transmission electron microscopy (TEM), various techniques are available for characterising grain-boundary space-charge layers: electric potential profiles may be calculated from the transmitted electrons' phase shift (electron holography);56–59 local electric fields can be probed by monitoring the electron beam's deflection in differential phase-contrast (DPC) imaging;60–62 and dopant concentration profiles across grain boundaries can be obtained, e.g., by electron energy loss spectroscopy (EELS). Such results will generally be obtained, however, for one specific grain boundary and may, therefore, not be representative of all grain boundaries in a ceramic sample.36,63 Nevertheless, the results from SPM,64 DPC,62 and EELS65–67 measurements are qualitatively consistent with a positive electric potential barrier at grain boundaries in acceptor-doped SrTiO3 bicrystals.

Various sources of errors in the extraction of either Φ0 or [small script l]gb have been discussed in the literature, both for electrical68–73 and for imaging67,74–77 techniques. But regardless of these technical shortcomings, a fundamental issue sets the results apart: the different experimental techniques operate in different temperature ranges. Electrical measurements are typically conducted at elevated temperatures (>500 K), while imaging techniques are typically applied at room temperature or below. While different temperature intervals are not problematic per se, they do constitute a problem when they are confined to different defect-kinetics regimes. Simple extrapolation of data from one temperature interval to another then becomes non-trivial. Two aspects are particularly important in that regard: first, the acceptor cations' migration is frozen-in at some temperature (between 1000 K and 1300 K in SrTiO3, depending on the cooling rate), such that the accumulation profile is frozen-in (see Fig. 1a); second, the oxygen surface-exchange reaction becomes kinetically hindered78 at some temperature (between 550 K and 750 K in SrTiO3, again depending on the cooling rate), such that the oxygen-vacancy content of a sample is frozen-in (see Fig. 1b). NB: the oxygen-vacancy mobility may still be appreciable down to much lower temperatures;78 in fact, the room-temperature conductivity in acceptor-doped SrTiO3 is predominantly ionic.79–81 When space-charge properties are probed in different temperature ranges, the outcomes must, therefore, be expected to differ, and the same is true for samples that have been subjected to different heat treatments. A sample that is cooled down slowly will generally reach equilibrium (chemical or electrochemical) down to lower temperatures than a sample that is quenched rapidly. The first of these two restricted equilibria, frozen-in acceptor profiles in space-charge zones, has been recognised for diverse electroceramic oxides.19,33,35,72,73,82–85 The second, negligible oxygen exchange, has been considered quantitatively in bulk defect-chemical models of diverse electroceramic oxides;27,78,86–94 it has not been included, however, in models of interfacial defect chemistry.


image file: d3cp05870k-f1.tif
Fig. 1 Illustrative considerations of two restricted equilibria: (a) we take the average cation residence time as a proxy for the time for the acceptor-dopant profile in a space-charge region to come to equilibrium. We calculate, therefore, the jump frequency at a field strength, and from this, the residence time. The jump frequency for image file: d3cp05870k-t1.tif was calculated from diffusion data,95 taking into account both the enhanced image file: d3cp05870k-t2.tif concentration in the space-charge layer (given Φ0 = 0.5 V) and the acceleration of transport by the space-charge layer's inherent electric field (assuming a field strength of |E| = 1 MV cm−1).96,97 (b) We estimate the chemical equilibration time for a sample of 1 mm thickness on the basis of surface-exchange coefficients kδ from ref. 40.

In this study, we examine the effects of the two above-mentioned, consecutive restricted equilibria on grain-boundary space-charge layers in acceptor-doped SrTiO3 by means of continuum simulations. First, we discuss in a general way how space-charge properties evolve with decreasing temperature. By inspecting the dependence of Φ0 on the temperature, we then demonstrate the differences that can arise when Φ0 values gathered at different temperatures are compared, and we assess the influence of different thermal histories on Φ0. Finally, we estimate different measures of a grain-boundary width (chemical and electrical) from our simulations and discuss their differences.

2 Modelling

The simulations were implemented in Python. Equations for the bulk defect chemistry were solved with standard routines from the SciPy package.98 The space-charge layers were simulated by means of the Finite Element Method (FEM), with the FEniCS package.99–102 Data analysis was carried out with the NumPy,103 SciPy,98 and Matplotlib104 packages.

2.1 Bulk defect chemistry

The bulk defect chemistry of acceptor-doped SrTiO3 is (at moderate pO2 and not too high T) determined by oxygen vacancies image file: d3cp05870k-t3.tif, electrons and electron holes (e′ and h˙), and acceptor cations, both ionised and neutral (image file: d3cp05870k-t4.tif and image file: d3cp05870k-t5.tif).78,79,105,106 The bulk concentrations (cb) of these defects are subject to the constraints of electroneutrality,
 
image file: d3cp05870k-t6.tif(1)
and of mass conservation of the acceptor dopant,
 
image file: d3cp05870k-t7.tif(2)
where cb(AccTi) denotes the bulk doping level. The relationships between the defect concentrations are described by three defect equilibria
 
image file: d3cp05870k-t8.tif(3)
 
nil ⇌ e′ + h˙(4)
 
image file: d3cp05870k-t9.tif(5)
The equilibrium bulk defect concentrations are obtained by solving the corresponding mass-action laws:
 
image file: d3cp05870k-t10.tif(6)
 
cb(e′)cb(h˙) = Keh(7)
 
image file: d3cp05870k-t11.tif(8)
with the equilibrium constants being described as temperature-dependent functions:107,108
 
image file: d3cp05870k-t12.tif(9)
 
image file: d3cp05870k-t13.tif(10)
 
image file: d3cp05870k-t14.tif(11)
These values are characteristic for SrTiO3: a reduction enthalpy between 5.2 eV and 6.1 eV,105,108–111 an effective bandgap of 3.4 eV (which includes the temperature dependence of the effective densities of states),108,112 and charge-transition levels of ∼1 eV for acceptor cations (e.g., Fe, Al) in titanate perovskites.34,78,79,107,110,113–115

Below the critical temperature Tsurfcrit, the oxygen surface-exchange reaction is assumed to be (abruptly) kinetically hindered, such that the oxygen-exchange equilibrium, eqn (3), can no longer be attained. The restricted chemical equilibrium below this temperature is described by solving together eqn (1), (2), (7) and (8), while keeping the oxygen-vacancy concentration fixed at its equilibrium value from T = Tsurfcrit.

2.2 Space-charge layers

Space-charge layer formation in acceptor-doped SrTiO3 is assumed35,116,117 to be due to a difference in the oxygen vacancies' standard chemical potential between the grain-boundary core and the bulk phase (Δμv < 0), an assumption supported by the results of atomistic simulations.118–120 This driving energy leads to image file: d3cp05870k-t15.tif segregating to the grain-boundary core, and thus, to the buildup of an excess positive core charge. Defect–defect interactions are neglected in our model, and hence, Δμv is assumed to be independent of the defect concentrations within the core. Non-zero driving energies for other point defects are possible (within certain bounds35), but for simplicity, they are set to zero. This means that changes in the concentrations of other mobile defects can only occur in response to the electrical field that develops through the re-distribution of the oxygen vacancies. The electrostatics of the system are described by the Poisson equation:
 
image file: d3cp05870k-t16.tif(12)
The relative dielectric permittivity of SrTiO3, εr, is a function of temperature, conforming to Curie–Weiss behaviour.121–126 Specifically, we describe εr(T) with121
 
image file: d3cp05870k-t17.tif(13)

We treat the ionic defects as structural elements (e.g., image file: d3cp05870k-t18.tif and image file: d3cp05870k-t19.tif), rather than as building units127 (e.g., image file: d3cp05870k-t20.tif), both in the bulk and in the grain boundary.35 Consequently, their concentrations must be coupled explicitly with a set of constraints. The ionic defects are subject to site-exclusion constraints that describe the impossibility of a lattice site's double occupation:

 
image file: d3cp05870k-t21.tif(14)
 
image file: d3cp05870k-t22.tif(15)
where γO and γTi are the site densities of the oxygen and titanium sublattices. These densities are calculated from the lattice parameter a:
 
image file: d3cp05870k-t23.tif(16)
with a being obtained as a function of temperature T according to97
 
a/Å = 3.90 + 6.64 × 10−5 K−1·T.(17)

We assume that the grains are large enough for the bulk to be unaffected by concentration changes (i.e., acceptor-dopant accumulation) in the space-charge layers. All bulk concentrations are, therefore, approximated as constant.

Electrons, holes, and oxygen vacancies attain electrochemical equilibrium within the solid under all considered conditions. The acceptor cations are assumed to attain electrochemical equilibrium above Tcatcrit (this is also called the Gouy–Chapman case128,129); for T < Tcatcrit, the redistribution of acceptor cations is modelled as being frozen-in. Possible defect–defect interactions are ignored, so that the defect concentrations in the space-charge layers can be expressed in standard Maxwell–Boltzmann forms:

 
c(h˙) = cb(h˙)·eαϕ(18)
 
c(e′) = cb(e′)·eαϕ(19)
 
image file: d3cp05870k-t24.tif(20)
 
image file: d3cp05870k-t25.tif(21)
with α := e/(kBT). By combining eqn (14) and (20), one obtains
 
image file: d3cp05870k-t26.tif(22)
The appropriate terms for the concentrations of image file: d3cp05870k-t27.tif and image file: d3cp05870k-t28.tif depend on the considered space-charge scenario. For the Gouy–Chapman (GC) case, eqn (15) and (21) must be combined with the additional condition
 
image file: d3cp05870k-t29.tif(23)
yielding the expressions
 
image file: d3cp05870k-t30.tif(24)
 
image file: d3cp05870k-t31.tif(25)
In the restricted-equilibrium (RE) case, the overall concentration profile of acceptor cations, image file: d3cp05870k-t32.tif is fixed, keeping the equilibrium profile from T = Tcatcrit:
 
cRE(AccTi) = ccrit(AccTi).(26)
By combining eqn (26) with eqn (21), one obtains
 
image file: d3cp05870k-t33.tif(27)
Note that eqn (27) may also be applied to the Mott–Schottky130–132 case, by setting ccrit(AccTi) = cb(AccTi).

The grain-boundary core is modelled as an infinitely thin charged plane. The areal charge density in the core, Qc, is taken into account by a Neumann boundary condition:

 
εrε0(∂+ϕ(xgb) − ∂ϕ(xgb)) = Qc,(28)
where ∂+ and ∂ denote the right- and left-hand side derivatives with respect to x, and xgb is the location of the grain boundary. Qc is calculated from formal point-defect concentrations in the core, cc.
 
image file: d3cp05870k-t34.tif(29)
Previous studies have indicated that the structurally distorted region around the grain-boundary plane is several lattice constants wide,11,133,134 and only some of the sites in it are energetically more favourable than the bulk sites.118–120 In our model, we attribute a formal width, wc, to the grain-boundary core, and a formal O-site density, γcO < γO, such that the areal core-site density is lower than it would be in a bulk slab of the same width.35 For convenience, we assume that the number of sites in the Ti sublattice as well as the number of electronic interface states is lowered by the same factor, γcO/γO. Since, for all defect species except image file: d3cp05870k-t35.tif, the segregation energy Δμ is assumed to be zero-valued, the site fractions of these defects are continuous across the grain boundary. That is, we can set
 
image file: d3cp05870k-t36.tif(30)
while the concentration of image file: d3cp05870k-t37.tif in the grain-boundary core is calculated by solving
 
image file: d3cp05870k-t38.tif(31)
subject to a site-exclusion constraint,
 
image file: d3cp05870k-t39.tif(32)
which yields
 
image file: d3cp05870k-t40.tif(33)
As stated before, this equation assumes implicitly that Δμv varies neither with the number of vacancies present in the core nor with changes in core composition through dopant enrichment. Taking such effects into account comprehensively requires the inclusion, for example, of activity coefficients for defect–defect interactions in the treatment, which then in turn requires numerical values for the activity coefficients to be specified (e.g. as a function of defect concentrations and temperature). Since such values are not available, the assumption of a constant Δμv is made out of necessity. And from the point of view of experiment, the assumption is reasonable, since experimental data can be described with a single Δμv value.35,72 In this sense, Δμv is an effective value in a dilute-solution treatment that to some degree reflects defect–defect interactions.

The model parameters for the simulation of the space-charge layers are summarised in Table 1.

Table 1 Model parameters for the simulation of the space-charge layers
Parameter Value Comment
[small script l]cell/μm 20 Simulation cell size
x gb/μm 0 Grain-boundary location
c a b/cm−3 1.6 × 1019 Bulk doping level
pO2/bar 0.2 O2 partial pressure
γ O [Eqn (16)] Bulk O-site density
γ Ti [Eqn (16)] Bulk Ti-site density
a [Eqn (17)] Lattice constant
w c/nm 2 Formal core width
γ cO/γO 0.05 Core O-site density
Δμv/eV −1.5 image file: d3cp05870k-t41.tif segregation energy
T catcrit/K Varied Negligible cation mobility
T surfcrit/K Varied Negligible oxygen exchange


3 Results

We simulated grain-boundary space-charge layers in SrTiO3 over a temperature range that encompasses both critical temperatures, that is, Tcatcrit, the critical temperature above which the acceptor cations achieve electrochemical equilibrium; and Tsurfcrit, the critical temperature above which the chemical equilibrium with the gas phase, eqn (3), is achieved. This leads naturally to the subdivision of the considered temperature range into three regimes: at high temperatures (T > Tcatcrit), the acceptor-cation profiles reach electrochemical equilibrium, and the oxygen sublattice is in chemical equilibrium with oxygen in the gas phase (full equilibrium, FE). At intermediate temperatures (Tsurfcrit < T < Tcatcrit), the cation distributions are frozen-in, but chemical equilibrium for oxygen is still attained (one restricted equilibrium, 1RE). At low temperatures (T < Tsurfcrit), oxygen exchange is also kinetically hindered, and the bulk oxygen-vacancy concentration is frozen-in (two restricted equilibria, 2RE).

Exemplary solutions of the Poisson equation [eqn (12)] for the relevant boundary conditions are shown in Fig. 2. The temperatures (T/K = {1400, 800, 300}) were chosen so as to represent the three different regimes (FE, 1RE, 2RE). Note that the absolute values of Φ0 depend on the choice of the grain-boundary parameters Δμv and wcγcO. The aim of our quantitative analysis is not a precise prediction of these values, but the identification of trends in the Φ0(T) behaviour. Visual inspection of the top row indicates that the space-charge potentials at the three different temperatures are not equal, and neither are the space-charge thicknesses: Φ0 increases with decreasing temperature (0.35 V at 1400 K; 0.47 V at 800 K; 0.63 V at 300 K), and [small script l]gb increases (≈20 nm at 1400 K; ≈35 nm at 800 K; ≈80 nm at 300 K). These results clearly demonstrate that, if Φ0 and [small script l]gb were obtained experimentally for the three different regimes (probably with different methods), and if all possible measurement errors were eliminated, the values obtained for Φ0 and [small script l]gb would differ considerably. We show later that, even though the trend is clear in this case, i.e., Φ0 and [small script l]gb increasing with decreasing T, data extrapolation from one regime to another is non-trivial.


image file: d3cp05870k-f2.tif
Fig. 2 Simulated grain-boundary space-charge layers in acceptor-doped SrTiO3. Profiles of the electric potential, ϕ, and of the point-defect concentrations, cdef, are depicted for three temperatures that represent the three relevant space-charge regimes: full equilibrium (FE) in the range Tcatcrit < T; one restricted-equilibrium (1RE) in the range Tsurfcrit < T < Tcatcrit; and two restricted equilibria (2RE) in the range T < Tsurfcrit, with Tcatcrit = 1000 K and Tsurfcrit = 650 K. Space-charge layer formation is assumed to originate in image file: d3cp05870k-t42.tif segregation to the grain-boundary core, driven by a difference Δμv = −1.5 eV in the image file: d3cp05870k-t43.tif standard chemical potential between grain boundary and bulk.

The second row of Fig. 2 shows the behaviour of the ionic defects within the space-charge layers. One perceives that the behaviour of the acceptor cations changes once the acceptor distribution is frozen-in at Tcatcrit. While the neutral acceptors, Acc×Ti, may be neglected in the FE case and exhibit an essentially uniform concentration profile, they are depleted in the RE regimes (weakly at T = 800 K and extremely strongly at T = 300 K in Fig. 2), even though they are effectively neutral. This depletion occurs because of the shift of the ionisation equilibrium, eqn (5). The concentration profile of image file: d3cp05870k-t44.tif, in turn, develops a plateau region in the outer parts of the space-charge layer, which is clearly visible at 300 K.

The concentration profiles of the electronic defects h˙ and e′ are shown in the bottom row of Fig. 2. These profiles indicate that the bulk concentrations of electronic defects fall drastically with decreasing temperature: between 1400 K and 300 K, cb(h˙) decreases by 15 orders of magnitude, and cb(e′) decreases by 30 orders of magnitude. With cb(h˙) = 1 × 103 cm−3 and cb(e′) = 5 × 10−16 cm−3, the material may be considered free of electronic defects at 300 K. Already at 800 K, however, the electronic defects do not contribute substantially to the charge density in the space-charge layer; compared to image file: d3cp05870k-t45.tif, c(h˙) is lower by more than a factor 400, and c(e′) by more than 11 orders of magnitude, throughout the space-charge layer.

3.1 The dependence of Φ0 on temperature and on thermal history

In order to demonstrate the problem of unifying data between the different regimes, we show in Fig. 3 the variation in Φ0 across a temperature range that includes both Tcatcrit and Tsurfcrit. Different thermal histories of a sample were modelled by varying both Tcatcrit and Tsurfcrit. To illustrate the influence of the freezing-in processes, we also calculated the temperature-dependent behaviour without the two processes being frozen-in (as indicated by the dashed and dotted lines, respectively). In Fig. 3, one sees, starting from high temperatures in the FE regime, that Φ0 increases weakly with decreasing temperature. If the sample now experiences the freezing-in of the acceptor-cation profiles at Tcatcrit/K = 1100, 1000 or 900, Φ0 increases sharply. There is a clear difference between the actual behaviour with the frozen-in AccTi profile (1RE) and the behaviour that would be obtained if the FE regime could be extended down to lower temperatures. In Fig. 3, one also sees, upon further cooling down of the sample through Tsurfcrit/K = 750, 650 or 550, another sharp increase in Φ0. And similarly, there is a difference between the actual behaviour with frozen-in oxygen-vacancy concentration (2RE) and the behaviour that would result if the sample could still equilibrate with the gas phase (for more detailed explanations, see ESI, Fig. S1). The abrupt changes in the Φ0(T) curve's behaviour, both at Tcatcrit and Tsurfcrit, reflect the simplified model of the processes being abruptly frozen-in. In reality, the transitions between the different regimes must be expected to take place more smoothly, possibly with hysteresis (arising from whether the sample is heated or cooled).
image file: d3cp05870k-f3.tif
Fig. 3 Simulated space-charge potentials, Φ0, as a function of temperature, given different thermal histories, which are represented by a variation of Tcatcrit and Tsurfcrit, with Tcatcrit/K = {900, 1000, 1100} and Tsurfcrit/K = {550, 650, 750}.

It is evident in Fig. 3 that values of Φ0 gathered in different temperature ranges (FE, 1RE, 2RE) will generally differ from one another (see also ESI, Fig. S2). In the case of Tcatcrit = 1100 K and Tsurfcrit = 550 K, for instance, electrical measurements carried out in the range of, say, 600 < T/K < 800 would probe values of Φ0 between 0.48 V and 0.51 V, whereas room-temperature imaging techniques would probe a value of 0.62 V. This may seem a small difference, but constructing a detailed, quantitative model requires a more precise set of data for Φ0(T) than Φ0 is (0.5 to 0.6) V. The most problematic aspect about the differences, however, is not their magnitude—it is the non-trivial behavior as a function of temperature. If only data for the 1RE regime for 600 < T/K < 800 were available, it would appear that Φ0 approached a plateau, and one would be inclined to take this plateau value, of ≈0.51 V, for all temperatures down to room temperature (Φ0 does not vary linearly with temperature, so that a linear extrapolation would be inappropriate). This value is, however, substantially lower than the actual (2RE) value of 0.62 V. This difference may also seem small, but macroscopic grain-boundary parameters, such as the grain-boundary resistance, depend exponentially on αΦ0, so the changing behaviour of Φ0(T) at a critical temperature implies a non-trivial temperature dependence also of these observables. In fact, if the grain-boundary resistance at a low temperature (below Tsurfcrit) should, for instance, be predicted by extrapolating a set of electrical resistances obtained at temperatures >Tsurfcrit, the extrapolated values would deviate strongly from the true ones. This illustrates how the non-trivial Φ0(T) behaviour can cause severe errors even when a value of Φ0 is not extracted at all from the experimental data. The different temperature ranges, and in particular, the kinetic hindrance of cation migration and oxygen surface exchange must be taken into account in order to achieve a consistent description of grain-boundary properties.

It has been pointed out before that results obtained for a single grain boundary will not generally be representative of a sample's grain-boundary population.36,37,63 Based on the considerations above, we extend this statement: such results will be quantitatively reproduceable, for the given grain-boundary type, only if the measurements are interpreted in conjunction with the sample's thermal history. Two microstructurally identical samples would exhibit, at the same temperature, different grain-boundary properties if they had previously been subjected to different thermal histories. If, for instance, one sample was cooled down slowly after sintering and the other was quenched quickly, the cation profiles at the grain boundaries within these two samples will correspond to Tcatcrit values well below and close to the sintering temperature, respectively. The variation of thermal histories is not merely a matter of different sample preparation routes, however. Even if, within a single ceramic sample, all grain boundaries showed identical behaviour in the FE regime (and this is unlikely for SrTiO336,37), the grain boundaries would show different behaviour in the 1RE regime on account of having different Tcatcrit, arising from the sample's outer regions cooling more rapidly than its centre.

The considerations above illustrate that choosing appropriate values of Tcatcrit and Tsurfcrit is not a simple task, and as a result, the uncertainty in Tcatcrit and Tsurfcrit will generally propagate to an uncertainty in Φ0, as indicated, e.g., by the variation in Φ0 at 300 K in Fig. 3. Compared to the deviations that would result from disregarding the freezing-in of the processes, however, this variation is relatively small: specifically, for the different (Tcatcrit, Tsurfcrit) combinations, differences in Φ0 of about ±2.5% are obtained, whereas Φ0 is predicted to be ∼20% lower in the scope of the 1RE regime, and ∼25% lower in the scope of the FE regime. That is, although only rough estimates of these critical temperatures will typically be available, the accuracy of the space-charge model can still be improved considerably by taking the two restricted equilibria into account.

3.2 Comparing different measures of the space-charge layer width

Like Φ0, the space-charge layer width [small script l]gb is also temperature-dependent (see Fig. 2), so values obtained in different temperature ranges are not directly comparable. In the case of [small script l]gb, however, there is an additional problem: it is not a uniquely defined quantity.135 Rather, there are various measures of a grain-boundary's thickness. Let us first set aside the grain boundary's structural width, which represents the extent of the region of disrupted lattice periodicity.136,137 It is typically obtained from electron micrographs, and values are of the order of a few lattice constants.11,133,134 In a space-charge picture, it may be identified with the width of the grain-boundary core. In terms of space-charge widths, there are three possibilities. The first two, the electrical width138,139 and the diffusional width,135,140 correspond to the regions over which transport properties are altered by the space-charge layers. Specifically, the electrical width probes the extent of the charge-carrier hindrance across a grain boundary and is typically obtained from capacitance data (Cgb/Cb) of an impedance spectroscopy measurement; in contrast, the diffusional width probes the extent of enhanced transport along a grain boundary, and its determination is not trivial.135 The third possibility, the chemical width,136–138 is itself not a uniquely defined property. Instead, a chemical width can be attributed, in principle, to every chemical species in the system, by taking the width of the grain-boundary feature in the respective concentration profile. It can be determined, for example, from the dopant concentration profiles measured by EELS or APT. Reported chemical widths are typically of the order of several nanometers.65–67,141 Electrical widths, on the other hand, show a strong variation; in acceptor-doped SrTiO3, thicknesses have been reported in the range of tens to hundreds of nanometers.7,8,14,33

To study the differences between chemical and electrical widths in more detail, we extracted both quantities from a set of simulated space-charge configurations at T/K = {500, 700, 1000}. A chemical width of ∼20 nm was estimated by visual inspection of the AccTi concentration profile (see Fig. 4a). Although the concentration profiles of image file: d3cp05870k-t46.tif and Acc×Ti depend on the temperature, this value remains unchanged, since c(AccTi) is frozen-in below Tcatcrit = 1000 K. The electrical grain-boundary width is typically obtained from Cgb/Cb. In a previous study,73 we found that, for a dilute solution of charge carriers, the electrical width can be estimated from the inflection points of the local conductivity profile, σloc(x) (see Fig. 4b). This was calculated as image file: d3cp05870k-t47.tif, with the electrical mobilities (uv, uh, ue) taken from ref. 79. The electrical width obtained in this way increases from 11 nm at T = 1000 K to 27 nm at T = 700 K, and further to 46 nm upon cooling to T = 500 K. Consistent with the above-mentioned differences between experimentally measured chemical and electrical widths, we find that the chemical width underestimates the extent of the space-charge layer for T < Tcatcrit, most severely (by a factor >2) at the lowest temperature. Note that the results presented here presuppose that electrochemical equilibrium of the acceptor-cation profile is actually established at Tcatcrit. In reality, this may not always be the case, especially if an accelerated sintering method (such as spark plasma sintering143,144 or flash sintering145) is applied,84,146,147 and it is reasonable to assume that the discrepancies between chemical and electrical width can be even larger in such cases.


image file: d3cp05870k-f4.tif
Fig. 4 Illustration of the discrepancies between different measures of the space-charge width for exemplary restricted-equilibrium space-charge layers at T/K = {500, 700, 1000}, with Tcatcrit = 1000 K and Tsurfcrit = 650 K: (a) a chemical width is estimated from the concentration profiles of AccTi, image file: d3cp05870k-t48.tif profiles are shown for comparison; (b) the electrical width is estimated from local-conductivity profiles, σloc, calculated from the concentration profiles as image file: d3cp05870k-t49.tif.

The substantial difference between the chemical and electrical widths is indicative of a fundamental disparity in the information that they convey. Typically, both quantities will be obtained at a temperature well below Tcatcrit, and hence, the AccTi profile will then be frozen-in. As a consequence, the chemical width of the AccTi profile reflects the state of the space-charge layer at Tcatcrit, rather than its state at the measurement temperature T.§ The electrical width, in turn, probes primarily the width of the image file: d3cp05870k-t50.tif depletion layer, which is assumed reasonably to reach electrochemical equilibrium.

4 Discussion

When determining and comparing data from different experimental methods, conflicting data may demand that the results be prioritised by their levels of reliability. For the extraction of Φ0 and [small script l]gb, one may be inclined to trust imaging results more than electrical measurements, since they do not require a preconceived space-charge model (e.g., Gouy–Chapman or Mott–Schottky). Since such results are generally obtained for a single grain boundary, however, they will not be representative of the variety of grain-boundary types and their different properties within a ceramic's grain-boundary population,15,63 and the same goes for electrical measurements on a specific bicrystal. Electrical measurements on ceramic samples, in contrast, do probe the grain-boundary population representatively, but at the cost of losing microscopic information:70 the arcs in impedance spectra, for instance, are typically associated with the microscopic electrical properties in the scope of the brick-layer model139 for the microstructure. This simplification has been demonstrated to yield good estimates of microscopic properties, provided that the grain-size distribution is monomodal and not too wide.8,14,68,69,148,149Φ0 is typically extracted from electrical data by closed-form expressions that neglect defect interactions and that require the interfacial defect chemistry to conform to a Mott–Schottky model. In concentrated solutions, such expressions have been shown to be unsuitable.85,150 In dilute solutions, however, Φ0 may be extracted with an error of only a few percent.73 The validity of the employed space-charge model can be put to a test with different consistency checks.71,73 By applying these formulas to electrical data obtained for acceptor-doped SrTiO3, plausible values of Φ0 have consistently been obtained in the literature.12,14,35,37,41

Imaging techniques, in contrast, have not yet succeeded at providing plausible estimates of Φ0 in acceptor-doped SrTiO3. The electrostatic grain-boundary barriers observed by SPM are unreasonably low (several tens of mV) in view of their being up to a few micrometers wide.64 Elemental profiles from EELS have only been reported for the acceptor dopants, and, while the observed accumulation of AccTi at the grain boundaries11,65–67 is consistent with the presence of depletion layers, Φ0 was not obtained from this (and if it was, it would refer to the state of the system at Tcatcrit, rather than at the measurement temperature, as discussed in the previous section). The results of a recent DPC study62 are also consistent with the presence of depletion layers, but Φ0 could not be extracted quantitatively. Lastly, electron holography yields a highly incoherent picture: the reported phase shifts at the grain boundaries are typically negative,67,75,76,151 although positive phase shifts have also been reported.74,152 Not only does this oppose the extensive evidence in favour of a depletion-layer model for acceptor-doped samples, but the calculated space-charge potentials are also unphysically large (∼−4 V) in some cases:67,151 for a semiconductor, such as SrTiO3, with a bandgap of 3 eV, Φ0 is constrained to lie between ∼+3 V and ∼−3 V,35 and these upper and lower bounds can only be reached if the Fermi level in the bulk is at the edge of the valence or conduction band, respectively.

These inaccuracies indicate that the mentioned imaging techniques are prone to disturbing influences. In SPM, the potential is measured at a distance from the surface, and hence, surface roughness and unknown surface charges can make a reconstruction of potential differences within the sample difficult, if not impossible.35,64 With regard to DPC, in turn, it has been argued that quantitative measurements could only be achieved if the TEM specimen was scaled down to nanometer dimensions, at which the space-charge layer properties would be compromised by finite-size effects.62 The severe inconsistencies in electron-holographic measurements point towards a conceptual problem: beside the space-charge potential, also topographic, structural and chemical changes in the grain-boundary region are reflected in the electron phase shift. If these influences cannot be neglected, the phase shift at the grain boundary will not be a measure of Φ0.67,74–77

Given pristine grain boundaries in dilute solid solutions, electrical measurements are currently the most reliable way to assess the space-charge potential, Φ0. In the development and refinement of imaging techniques for the analysis of space-charge properties, the accuracy of a measurement should, therefore, be appraised by comparing it with electrical measurements.

A joining of imaging techniques and electrical methods can also be highly useful in identifying the relationships between grain-boundary structure, composition, and conductivity. It has been demonstrated recently that a thorough treatment of imaging data can help bridge the gap between microscopic observations and macroscopic conductivity.63 Further investigations on the microscopic structure–composition–conductivity relations are a promising route in making ceramic conductivities more predictable. Unless both electrical and imaging measurements are carried out under the same conditions (e.g., by appropriate in situ approaches), such endeavours will generally face the issue of unifying data from different temperature ranges.

A direct comparison of elevated-temperature electrical data with room-temperature data from imaging techniques implicitly presupposes that one specific value can be attributed to Φ0 and/or [small script l]gb for a given material. Neither Φ0 nor [small script l]gb, however, are material-specific properties. They depend, of course, on material-specific parameters, but also on the thermodynamic conditions (T, pO2), on the grain-boundary type,36,37,63 and on the sample's thermal history.

5 Conclusions

We simulated grain-boundary space-charge layers in SrTiO3 at different temperatures, with a focus on the characteristic temperature ranges of electrical methods, on the one hand, and imaging techniques, on the other hand. Various thermal histories were represented by varying the critical temperatures of cation redistribution, Tcatcrit, and of the oxygen surface exchange, Tsurfcrit.

We treated the trends in Φ0(T) as a proxy for grain-boundary properties as a function of temperature. In the case considered, we found that Φ0 generally increased with decreasing temperature, but at the onset of a restricted equilibrium, at Tcatcrit and then at Tsurfcrit, the increase became substantially more pronounced. From our results, we identified some general problems that may arise when data gathered by different experimental approaches are considered together.

1. An extrapolation of grain-boundary properties across the transitions between restricted equilibria (e.g., from electrical measurement conditions to room temperature) will lead to large errors.

2. The critical temperatures can be difficult to estimate. Even when only a very rough estimate can be made, however, taking the frozen processes into account will strongly enhance the accuracy of a space-charge model.

3. When the acceptor-cation redistribution is frozen-in, the chemical width from acceptor-cation profiles will underestimate the space-charge layer's thickness. The electrical grain-boundary width from impedance spectroscopy is a more reliable measure of its full extent.

In summary, our results demonstrate that probing the behaviour of grain-boundary space-charge layers in electroceramics in different temperature ranges will naturally yield values of characteristic parameters that are different and not trivially related. Owing to these non-trivial differences, the grain-boundary properties of an electroceramic oxide as a function of temperature may be understood and predicted only within the scope of a space-charge model that takes into account both restricted equilibria.

Author contributions

Adrian L. Usler: methodology, software, investigation, writing – original draft, visualisation, writing – review & editing. Fabian Ketter: investigation, writing – review & editing, visualisation. Roger A. De Souza: conceptualization, supervision, writing – original draft, writing – review & editing, project administration, funding acquisition.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by RWTH Aachen University and by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) within the Priority Programme (SPP1959) “FieldsMatter” under project DE 2854/9-2 and within the Sonderforschungsbereich (collaborative research centre) SFB 1548, FLAIR (Fermi Level Engineering applied to oxide electroceramics).

Notes and references

  1. R. A. De Souza, J. Fleig, R. Merkle and J. Maier, Int. J. Mater. Res., 2022, 94, 218–225 Search PubMed.
  2. D. Singh, M. Lorenzo-Martin, G. Chen, F. Gutiérrez-Mora and J. L. Routbort, J. Eur. Ceram. Soc., 2007, 27, 3377–3384 CrossRef CAS.
  3. L. Porz, M. Scherer, M. Höfling, A. Nakamura, W. Rheinheimer and J. Rödel, J. Mater. Sci., 2023, 58, 2430–2438 CrossRef CAS.
  4. E. R. Pfeiffer and J. F. Schooley, Phys. Lett. A, 1969, 29, 589–590 CrossRef CAS.
  5. H. Suzuki, H. Kurosawa, J. Chen, Y. Hirotsu, K. Nakajima, T. Yamashita, H. Myoren and Y. Osaka, Supercond. Sci. Technol., 1991, 4, 479 CrossRef CAS.
  6. T. R. N. Kutty and S. Philip, Mater. Sci. Eng., B, 1995, 33, 58–66 CrossRef.
  7. I. Denk, J. Claus and J. Maier, J. Electrochem. Soc., 1997, 144, 3526 CrossRef CAS.
  8. S. Rodewald, J. Fleig and J. Maier, J. Am. Ceram. Soc., 2001, 84, 521–530 CrossRef CAS.
  9. D.-H. Kwon, S. Lee, C. S. Kang, Y. S. Choi, S. J. Kang, H. L. Cho, W. Sohn, J. Jo, S.-Y. Lee, K. H. Oh, T. W. Noh, R. A. De Souza, M. Martin and M. Kim, Adv. Mater., 2019, 31, 1901322 CrossRef PubMed.
  10. J. Maier, G. Schwitzgebel and H.-J. Hagemann, J. Solid State Chem., 1985, 58, 1–13 CrossRef CAS.
  11. F. Noll, W. Münch, I. Denk and J. Maier, Solid State Ion., 1996, 86, 711–717 CrossRef.
  12. M. Vollmann, R. Hagenbeck and R. Waser, J. Am. Ceram. Soc., 1997, 80, 2301–2314 CrossRef CAS.
  13. J. R. Jurado, M. T. Colomer and J. R. Frade, J. Am. Ceram. Soc., 2000, 83, 2715–2720 CrossRef CAS.
  14. J. Fleig, S. Rodewald and J. Maier, J. Appl. Phys., 2000, 87, 2372–2381 CrossRef CAS.
  15. R. A. De Souza, J. Fleig, J. Maier, O. Kienzle, Z. Zhang, W. Sigle and M. Rühle, J. Am. Ceram. Soc., 2003, 86, 922–928 CrossRef CAS.
  16. R. Meyer and R. A. De Souza, Electrochemical Society Proceedings, 2003, p. 220 Search PubMed.
  17. D. W. Wang and A. S. Nowick, J. Solid State Chem., 1980, 35, 325–333 CrossRef CAS.
  18. R. Gerhardt and A. S. Nowick, J. Am. Ceram. Soc., 1986, 69, 641–646 CrossRef CAS.
  19. X. Guo and R. Waser, Prog. Mater. Sci., 2006, 51, 151–210 CrossRef CAS.
  20. H. J. Avila-Paredes and S. Kim, Solid State Ion., 2006, 177, 3075–3080 CrossRef CAS.
  21. J. Bauerle, J. Phys. Chem. Solids, 1969, 30, 2657–2670 CrossRef CAS.
  22. T. Van Dijk and A. J. Burggraaf, Phys. Status Solidi A, 1981, 63, 229–240 CrossRef CAS.
  23. M. J. Verkerk, B. J. Middelhuis and A. J. Burggraaf, Solid State Ion., 1982, 6, 159–170 CrossRef CAS.
  24. K. Huang, R. S. Tichy and J. B. Goodenough, J. Am. Ceram. Soc., 1998, 81, 2576–2580 CrossRef CAS.
  25. H. J. Park and S. Kim, J. Phys. Chem. C, 2007, 111, 14903–14910 CrossRef CAS.
  26. K. D. Kreuer, Solid State Ion., 1999, 125, 285–302 CrossRef CAS.
  27. H. G. Bohn and T. Schober, J. Am. Ceram. Soc., 2000, 83, 768–772 CrossRef CAS.
  28. F. Iguchi, N. Sata, T. Tsurui and H. Yugami, Solid State Ion., 2007, 178, 691–695 CrossRef CAS.
  29. P. Babilo, T. Uda and S. M. Haile, J. Mater. Res., 2007, 22, 1322–1330 CrossRef CAS.
  30. S. Tao and J. T. S. Irvine, J. Solid State Chem., 2007, 180, 3493–3503 CrossRef CAS.
  31. S. M. Haile, D. L. West and J. Campbell, J. Mater. Res., 1998, 13, 1576–1595 CrossRef CAS.
  32. W. G. Coors and D. W. Readey, J. Am. Ceram. Soc., 2002, 85, 2637–2640 CrossRef CAS.
  33. M. Vollmann and R. Waser, J. Am. Ceram. Soc., 1994, 77, 235–243 CrossRef.
  34. R. Hagenbeck and R. Waser, J. Appl. Phys., 1998, 83, 2083–2092 CrossRef CAS.
  35. R. A. De Souza, Phys. Chem. Chem. Phys., 2009, 11, 9939–9969 RSC.
  36. R. A. De Souza, J. Fleig, J. Maier, Z. Zhang, W. Sigle and M. Rühle, J. Appl. Phys., 2005, 97, 053502 CrossRef.
  37. Z. Zhang, W. Sigle, R. A. De Souza, W. Kurtz, J. Maier and M. Rühle, Acta Mater., 2005, 53, 5007–5015 CrossRef CAS.
  38. P. Balaya, M. Ahrens, L. Kienle, J. Maier, B. Rahmati, S. B. Lee, W. Sigle, A. Pashkin, C. Kuntscher and M. Dressel, J. Am. Ceram. Soc., 2006, 89, 2804–2811 CrossRef CAS.
  39. I. Denk, F. Noll and J. Maier, J. Am. Ceram. Soc., 1997, 80, 279–285 CrossRef CAS.
  40. M. Leonhardt, J. Jamnik and J. Maier, Electrochem. Solid-State Lett., 1999, 2, 333 CrossRef CAS.
  41. X. Guo, J. Fleig and J. Maier, J. Electrochem. Soc., 2001, 148, J50 CrossRef CAS.
  42. X. Guo, J. Fleig and J. Maier, Solid State Ion., 2002, 154, 563–569 CrossRef.
  43. P. Balaya, J. Jamnik, J. Fleig and J. Maier, Appl. Phys. Lett., 2006, 88, 062109 CrossRef.
  44. P. Balaya, J. Jamnik, J. Fleig and J. Maier, J. Electrochem. Soc., 2007, 154, P69 CrossRef CAS.
  45. G. Gregori, P. Lupetin and J. Maier, ECS Trans., 2012, 45, 19 CrossRef CAS.
  46. Y. Wu, P. C. Bowes, J. N. Baker and D. L. Irving, J. Appl. Phys., 2020, 128, 014101 CrossRef CAS.
  47. T. Hölbling and R. Waser, J. Appl. Phys., 2002, 91, 3037–3043 CrossRef.
  48. S. K. Kim, S. Khodorov, C.-T. Chen, S. Kim and I. Lubomirsky, Phys. Chem. Chem. Phys., 2013, 15, 8716–8721 RSC.
  49. S. K. Kim, S. Khodorov, I. Lubomirsky and S. Kim, Phys. Chem. Chem. Phys., 2014, 16, 14961–14968 RSC.
  50. S. Kim, S. K. Kim, S. Khodorov, J. Maier and I. Lubomirsky, Phys. Chem. Chem. Phys., 2016, 18, 3023–3031 RSC.
  51. J. M. R. Weaver and H. K. Wickramasinghe, J. Vac. Sci. Technol., B: Nanotechnol. Microelectron.: Mater., Process., Meas., Phenom., 1991, 9, 1562–1565 CAS.
  52. W. Nabhan, B. Equer, A. Broniatowski and G. De Rosny, Rev. Sci. Instrum., 1997, 68, 3108–3111 CrossRef CAS.
  53. D. R. Diercks, J. Tong, H. Zhu, R. Kee, G. Baure, J. C. Nino, R. O'Hayre and B. P. Gorman, J. Mater. Chem. A, 2016, 4, 5167–5175 RSC.
  54. B. Gault, A. Chiaramonti, O. Cojocaru-Mirédin, P. Stender, R. Dubosq, C. Freysoldt, S. K. Makineni, T. Li, M. Moody and J. M. Cairney, Nat. Rev. Methods Primers, 2021, 1, 51 CrossRef CAS PubMed.
  55. D. R. Clark, D. R. Diercks, S. Ricote, T. T. Dearden, N. P. Sullivan, J. W. Medlin, B. P. Gorman and R. P. OHayre, J. Mater. Chem. C, 2023, 11, 5082–5091 RSC.
  56. P. G. Merli, G. F. Missiroli and G. Pozzi, Phys. Status Solidi A, 1975, 30, 699–711 CrossRef.
  57. S. Frabboni, G. Matteucci, G. Pozzi and M. Vanzi, Phys. Rev. Lett., 1985, 55, 2196 CrossRef CAS PubMed.
  58. W.-D. Rau, P. Schwander and A. Ourmazd, Solid State Phenom., 1998, 63, 525–528 Search PubMed.
  59. W.-D. Rau, P. Schwander, F. H. Baumann, W. Höppner and A. Ourmazd, Phys. Rev. Lett., 1999, 82, 2614 CrossRef CAS.
  60. N. Shibata, S. D. Findlay, Y. Kohno, H. Sawada, Y. Kondo and Y. Ikuhara, Nat. Phys., 2012, 8, 611–615 Search PubMed.
  61. K. Müller-Caspary, T. Grieb, J. Müßener, N. Gauquelin, P. Hille, J. Schörmann, J. Verbeeck, S. Van Aert, M. Eickhoff and A. Rosenauer, Phys. Rev. Lett., 2019, 122, 106102 CrossRef PubMed.
  62. C. Yang, Y. Wang, W. Sigle and P. A. van Aken, Nano Lett., 2021, 21, 9138–9145 CrossRef CAS PubMed.
  63. W. J. Bowman, A. Darbal and P. A. Crozier, ACS Appl. Mater. Interfaces, 2019, 12, 507–517 CrossRef PubMed.
  64. B. D. Huey and D. A. Bonnell, Solid State Ion., 2000, 131, 51–60 CrossRef CAS.
  65. Y.-M. Chiang and T. Takagi, J. Am. Ceram. Soc., 1990, 73, 3278–3285 CrossRef CAS.
  66. N. Wilcox, V. Ravikumar, R. Rodrigues, V. Dravid, M. Vollmann, R. Waser, K. Soni and A. Adriaens, Solid State Ion., 1995, 75, 127–136 CrossRef CAS.
  67. S. J. P. Longworth, K. M. Knowles and R. E. Dunin-Borkowski, J. Phys.: Conf. Ser., 2006, 235 CrossRef.
  68. J. Fleig and J. Maier, J. Electrochem. Soc., 1998, 145, 2081 CrossRef CAS.
  69. J. Fleig and J. Maier, J. Eur. Ceram. Soc., 1999, 19, 693–696 CrossRef CAS.
  70. J. Fleig, Solid State Ion., 2000, 131, 117–127 CrossRef CAS.
  71. S. Kim, Phys. Chem. Chem. Phys., 2016, 18, 19787–19791 RSC.
  72. J. P. Parras, C. Cao, Z. Ma, R. Mücke, L. Jin, R. Dunin-Borkowski, O. Guillon and R. A. De Souza, J. Am. Ceram. Soc., 2020, 103, 1755–1764 CrossRef CAS.
  73. A. L. Usler and R. A. De Souza, J. Electrochem. Soc., 2021, 168, 056504 CrossRef.
  74. H. R. Liu, Y. G. Wang, Q. X. Liu and Q. B. Yang, J. Phys. D: Appl. Phys., 2004, 37, 1478 CrossRef CAS.
  75. C. T. Koch, Int. J. Mater. Res., 2010, 101, 43–49 CrossRef CAS.
  76. S. von Alfthan, N. A. Benedek, L. Chen, A. Chua, D. Cockayne, K. J. Dudeck, C. Elsässer, M. W. Finnis, C. T. Koch and B. Rahmati, et al. , Annu. Rev. Mater. Res., 2010, 40, 557–599 CrossRef CAS.
  77. T. Bondevik, H. H. Ness, C. Bazioti, T. Norby, O. M. Løvvik, C. T. Koch and Ø. Prytz, Phys. Chem. Chem. Phys., 2019, 21, 17662–17672 RSC.
  78. R. Waser, J. Am. Ceram. Soc., 1991, 74, 1934–1940 CrossRef CAS.
  79. I. Denk, W. Münch and J. Maier, J. Am. Ceram. Soc., 1995, 78, 3265–3272 CrossRef CAS.
  80. R. A. De Souza and A. H. H. Ramadan, Phys. Chem. Chem. Phys., 2013, 15, 4505–4509 RSC.
  81. R. A. Maier and C. A. Randall, J. Am. Ceram. Soc., 2016, 99, 3350–3359 CrossRef CAS.
  82. J.-S. Lee and D.-Y. Kim, J. Mater. Res., 2001, 16, 2739–2751 CrossRef CAS.
  83. A. Tschöpe, S. Kilassonia and R. Birringer, Solid State Ion., 2004, 173, 57–61 CrossRef.
  84. M. Shirpour, B. Rahmati, W. Sigle, P. A. van Aken, R. Merkle and J. Maier, J. Phys. Chem. C, 2012, 116, 2453–2461 CrossRef CAS.
  85. X. Tong, D. S. Mebane and R. A. De Souza, J. Am. Ceram. Soc., 2020, 103, 5–22 CrossRef CAS.
  86. T. Scherban, W. K. Lee and A. S. Nowick, Solid State Ion., 1988, 28, 585–588 CrossRef.
  87. J. F. Liu and A. S. Nowick, Solid State Ion., 1992, 50, 131–138 CrossRef CAS.
  88. J. Maier, Solid State Ion., 1996, 86, 55–67 CrossRef.
  89. J. Müller, K. D. Kreuer, J. Maier, S. Matsuo and M. Ishigame, Solid State Ion., 1997, 97, 421–427 CrossRef.
  90. H. G. Bohn, T. Schober, T. Mono and W. Schilling, Solid State Ion., 1999, 117, 219–228 CrossRef CAS.
  91. K. Sasaki, J. Claus and J. Maier, Solid State Ionics, 1999, 121, 51–60 CrossRef CAS.
  92. K. Sasaki and J. Maier, J. Eur. Ceram. Soc., 1999, 19, 741–745 CrossRef CAS.
  93. E. Bucher, A. Egger, P. Ried, W. Sitte and P. Holtappels, Solid State Ionics, 2008, 179, 1032–1035 CrossRef CAS.
  94. R. Zohourian, R. Merkle, G. Raimondi and J. Maier, Adv. Funct. Mater., 2018, 28, 1801241 CrossRef.
  95. K. Gömann, G. Borchardt, M. Schulz, A. Gömann, W. Maus-Friedrichs, B. Lesage, O. Kaïtasov, S. Hoffmann-Eifert and T. Schneller, Phys. Chem. Chem. Phys., 2005, 7, 2053–2060 RSC.
  96. A. R. Genreith-Schriever and R. A. De Souza, Phys. Rev. B, 2016, 94, 224304 CrossRef.
  97. D. Kemp and R. A. De Souza, Phys. Rev. Mater., 2021, 5, 105401 CrossRef CAS.
  98. P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wilson, K. J. Millman, N. Mayorov, A. R. J. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, I. Polat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen, E. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, P. van Mulbregt and SciPy 1.0 Contributors, Nat. Methods, 2020, 17, 261–272 CrossRef CAS PubMed.
  99. R. C. Kirby and A. Logg, ACM Trans. Math. Softw, 2006, 32(3), 417–444 CrossRef.
  100. A. Logg and G. N. Wells, ACM Trans. Math. Softw, 2010, 37(2), 20 CrossRef.
  101. A. Logg, K.-A. Mardal and G. Wells, Automated solution of differential equations by the finite element method: The FEniCS book, Springer Science & Business Media, 2012, vol. 84 Search PubMed.
  102. M. S. Alnæs, A. Logg, K. B. Ølgaard, M. E. Rognes and G. N. Wells, ACM Trans. Math. Softw., 2014, 40, 1–37 CrossRef.
  103. C. R. Harris, K. J. Millman, S. J. van der Walt, R. Gommers, P. Virtanen, D. Cournapeau, E. Wieser, J. Taylor, S. Berg, N. J. Smith, R. Kern, M. Picus, S. Hoyer, M. H. van Kerkwijk, M. Brett, A. Haldane, J. Fernández del Río, M. Wiebe, P. Peterson, P. Gérard-Marchant, K. Sheppard, T. Reddy, W. Weckesser, H. Abbasi, C. Gohlke and T. E. Oliphant, Nature, 2020, 585, 357–362 CrossRef CAS PubMed.
  104. J. D. Hunter, Comput. Sci. Eng., 2007, 9, 90–95 Search PubMed.
  105. R. Moos and K. H. Härdtl, J. Am. Ceram. Soc., 1997, 80, 2549–2562 CrossRef CAS.
  106. M. Knight, I. Reimanis, A. Meyer, J.-H. Preusker and W. Rheinheimer, J. Am. Ceram. Soc., 2023, 106, 4740–4751 CrossRef CAS.
  107. C.-J. Shin, H.-I. Yoo and C.-E. Lee, Solid State Ion., 2007, 178, 1081–1087 CrossRef CAS.
  108. R. A. De Souza, F. Gunkel, S. Hoffmann-Eifert and R. Dittmann, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89, 241401 CrossRef.
  109. G. M. Choi and H. L. Tuller, J. Am. Ceram. Soc., 1988, 71, 201–205 CrossRef CAS.
  110. H.-I. Yoo and K. D. Becker, Phys. Chem. Chem. Phys., 2005, 7, 2068–2073 RSC.
  111. F. Horikiri, N. Iizawa, L. Han, K. Sato, K. Yashiro, T. Kawada and J. Mizusaki, Solid State Ion., 2008, 179, 2335–2344 CrossRef CAS.
  112. R. A. De Souza, Adv. Funct. Mater., 2015, 25, 6326–6342 CrossRef CAS.
  113. R. Waser, T. Bieger and J. Maier, Solid State Commun., 1990, 76, 1077–1081 CrossRef CAS.
  114. D. M. Smyth, Curr. Opin. Solid State Mater. Sci., 1996, 1, 692–697 CrossRef CAS.
  115. A. Klein, K. Albe, N. Bein, O. Clemens, K. A. Creutz, P. Erhart, M. Frericks, E. Ghorbani, J. P. Hofmann and B. Huang, et al. , J. Electroceram., 2023, 1–31 CAS.
  116. P. C. McIntyre, J. Am. Ceram. Soc., 2000, 83, 1129–1136 CrossRef CAS.
  117. R. A. De Souza and E. C. Dickey, Philos. Trans. R. Soc. A, 2019, 377, 20180430 CrossRef CAS PubMed.
  118. A. H. H. Ramadan and R. A. De Souza, Acta Mater., 2016, 118, 286–295 CrossRef CAS.
  119. C. Tao, D. Mutter, D. F. Urban and C. Elsässer, Phys. Rev. B, 2021, 104, 054114 CrossRef CAS.
  120. D. Mutter, C. Tao, D. F. Urban and C. Elsässer, Adv. Eng. Mater., 2023, 2201847 CrossRef CAS.
  121. H. E. Weaver, J. Phys. Chem. Solids, 1959, 11, 274–277 CrossRef CAS.
  122. G. Rupprecht and R. O. Bell, Phys. Rev., 1964, 135, A748 CrossRef.
  123. T. Sakudo and H. Unoki, Phys. Rev. Lett., 1971, 26, 851 CrossRef CAS.
  124. R. A. Van der Berg, P. W. M. Blom, J. F. M. Cillessen and R. M. Wolf, Appl. Phys. Lett., 1995, 66, 697–699 CrossRef CAS.
  125. M. Trainer, Eur. J. Phys., 2000, 21, 459 CrossRef CAS.
  126. S. A. Prosandeev, A. E. Maslennikov, W. Kleemann and J. Dec, Ferroelectrics, 2000, 238, 171–178 CrossRef.
  127. J. Maier, Phys. Chem. Chem. Phys., 2009, 11, 3011–3022 RSC.
  128. M. Gouy, J. Phys. Theor. Appl., 1910, 9, 457–468 CrossRef.
  129. D. L. Chapman, Lond. Edinb. Dublin Philos. Mag. J. Sci., 1913, 25, 475–481 CrossRef.
  130. W. Schottky, Z. Phys., 1939, 113, 367–414 CrossRef CAS.
  131. W. Schottky and E. Spenke, Wissenschaftliche Veröffentlichungen aus den Siemens-Werken, 1939, 18, 225–291 Search PubMed.
  132. N. F. Mott, Proc. R. Soc. London, Ser. A, 1939, 171, 27–38 Search PubMed.
  133. O. Kienzle, M. Exner and F. Ernst, Phys. Status Solidi A, 1998, 166, 57–71 CrossRef CAS.
  134. S. Hutt, O. Kienzle, F. Ernst and M. Rühle, Z. Metallkd., 2001, 92, 105–109 CAS.
  135. R. E. Mistler and R. L. Coble, J. Appl. Phys., 1974, 45, 1507–1509 CrossRef CAS.
  136. K. D. Chowdhury, R. Carpenter and W. Braue, Ultramicroscopy, 1992, 40, 229–239 CrossRef.
  137. H. Gu and Y. Shinoda, Interface Sci., 2000, 8, 269–278 CrossRef CAS.
  138. W. G. Morris, J. Vac. Sci. Technol., 1976, 13, 926–931 CrossRef CAS.
  139. T. Van Dijk and A. J. Burggraaf, Phys. Status Solidi A, 1981, 63, 229–240 CrossRef CAS.
  140. J. P. Parras and R. A. De Souza, Acta Mater., 2020, 195, 383–391 CrossRef CAS.
  141. H. Yang, P. G. Kotula, Y. Sato, M. Chi, Y. Ikuhara and N. D. Browning, Mater. Res. Lett., 2014, 2, 16–22 CrossRef CAS.
  142. J. Jamnik and J. Maier, Phys. Chem. Chem. Phys., 2001, 3, 1668–1678 RSC.
  143. Z. A. Munir, U. Anselmi-Tamburini and M. Ohyanagi, J. Mater. Sci., 2006, 41, 763–777 CrossRef CAS.
  144. O. Guillon, J. Gonzalez-Julian, B. Dargatz, T. Kessel, G. Schierning, J. Räthel and M. Herrmann, Adv. Eng. Mater., 2014, 16, 830–849 CrossRef CAS.
  145. M. Cologna, B. Rashkova and R. Raj, J. Am. Ceram. Soc., 2010, 93, 3556–3559 CrossRef CAS.
  146. M. Shirpour, G. Gregori, L. Houben, R. Merkle and J. Maier, Solid State Ion., 2014, 262, 860–864 CrossRef CAS.
  147. J. Bae, Y. Lim, J.-S. Park, D. Lee, S. Hong, J. An and Y.-B. Kim, J. Electrochem. Soc., 2016, 163, F919 CrossRef CAS.
  148. G. Dezanneau, A. Morata, A. Tarancón, M. Salleras, F. Peiró and J. R. Morante, Appl. Phys. Lett., 2006, 88, 141920 CrossRef.
  149. G. Dezanneau, A. Morata, A. Tarancón, F. Peiró and J. R. Morante, Solid State Ion., 2006, 177, 3117–3121 CrossRef CAS.
  150. T. Bondevik, J. M. Polfus and T. Norby, Solid State Ion., 2020, 353, 115369 CrossRef CAS.
  151. V. Ravikumar, R. P. Rodrigues and V. P. Dravid, Phys. Rev. Lett., 1995, 75, 4063 CrossRef CAS PubMed.
  152. Y. G. Wang and V. P. Dravid, Philos. Mag. Lett., 2002, 82, 425–432 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp05870k
On occasion, impedance spectra are calculated directly from cdef(x) or ϕ(x) profiles in the literature, by applying the Maxwell–Wagner approximation. This is not a suitable approximation for space-charge layers, however.142 Impedance spectra can only be obtained from such profiles by means of drift–diffusion simulations; either explicitly time-dependent73 or linearised, in Fourier space.142
§ The equilibrium width of the space-charge layer could be probed, in principle, if valence-state-sensitive measurements of the image file: d3cp05870k-t51.tif and image file: d3cp05870k-t52.tif concentration profiles could be achieved with a very high precision and sensitivity.

This journal is © the Owner Societies 2024
Click here to see how this site uses Cookies. View our privacy policy here.