2.1. Piezoelectric Softening
The idea of evaluating the piezoelectric properties from unpoled samples is based on the fact that the elastic constants undergo a softening of piezoelectric origin in the ferroelectric state, such that the compliance tensor
can be written as [
11]
where
is the compliance in the paraelectric (PE) state, and
and
are the piezoelectric and reciprocal dielectric susceptibility tensors. From this expression, and with the usual approximation for ferroelectrics
, it appears that the compliance of the FE state is softer than that of the PE state of the quantity
and this suggests the possibility of evaluating the piezoelectric constants from the piezoelectric softening. When measuring an elastic modulus of a ceramic, one probes a polycrystalline average of the tensors, and we may write for simplicity
, which makes even more clear that
also for unpoled ceramics, where
is null but
is not. The reason is that each domain undergoes its softening according to the local direction of the spontaneous polarization and applied stress. Since the elastic constants are centrosymmetric tensors, domains with opposite directions of the polarization undergo the same change in the compliance; therefore, the corresponding changes in polarization under the probing stress cancel out but the strains do not.
The above expressions are equivalent to the well known relationships between the compliances at constant field,
, and constant dielectric displacement,
[
3,
12]:
where
k is the electromechanical coupling factor and the permittivity is at constant stress
. These can be rewritten as
which coincides with Equation (
2). In fact, in a purely elastic measurement there is no external field
E,
, and the compliance of the PE phase is
, namely the compliance measured keeping
constant and, being also
0, this is equivalent to keeping
constant.
We describe now a simplified demonstration [
13] of Equation (
2), without considering the tensorial nature of the formulas, in order to provide a clear physical picture of the origin of the softening in the FE state. The compliance, as measured in a purely elastic experiment with no applied electric field, is defined as
where
and
are strain and stress,
is the compliance at constant polarization
P, coinciding with the total compliance in the PE phase with
, and the second term takes into account the fact that, in the FE state, the probing stress modulates the polarization, which in turn modulates the strain. This already makes clear that the additional softening in the FE state is due to a combination of direct and converse piezoelectric effects. The equilibrium spontaneous polarization
under stress can be found by exploiting the fact that it minimizes the free energy
. Actually, under the conditions of applied stress, the elastic Gibbs energy
[
14] is minimized, since its differential is
with
rather than
as independent variable, and
is found from
The description of the FE state is contained in the free energy
, which generally is expanded in a series of even powers of
P, according to the Landau theory of phase transitions. For deducing Equation (
2) we do not need to specify the form of
, but only that of the coupling
between
P and
. The simplest form of such a coupling term, in a material with centrosymmetric (non piezoelectric) PE phase, is [
15]
to be added to Equation (
6), where
Q is the electrostrictive coupling constant, and the electrostrictive strain is, according to Equation (
7),
. A piezoelectric coupling
cannot be introduced, because
G must describe both the FE and PE states, and we are dealing with materials with a, generally cubic, PE state invariant under inversion, and a term
changes sign under inversion. The piezoelectric strain arises in the FE state with spontaneous polarization
, as shown in
Figure 1. In fact, the electrostrictive strain,
, is quadratic in
P in the PE state around
, but is almost linear with slope
in the FE state with spontaneous polarization
and spontaneous strain
. This is at the basis of the converse piezoelectric effect,
d, whose coefficient is
, because
, the dielectric susceptibility.
The direct piezoelectric effect is, by definition,
, and it is easy to verify that it is the same coefficient as for the converse effect. In fact,
and its stress derivative can be found from the condition of minimum elastic Gibbs energy, Equations (
6) and (
8)
which can be derivated with respect to
yielding
because
F is a function of
P only. In addition,
and the last term can be neglected because the probing stress
is small, so that we find
with the same coefficient as the converse effect.
We can now substitute our results for the FE state,
and
into Equation (
5):
Finally,
so that in the FE state one generally has
and Equations (
2) and (
10) coincide.
This result is general, since no hypothesis has been done on the form of , but it does not tell us what kind of temperature dependence we expect from and , and hence d and ; yet, it is useful to anticipate this temperature dependence, since the elastic measurement must be extended above and below , in order to recognize and subtract the non-ferroelectric contribution .
2.3. Anharmonic Stiffening and Electrostrictive Coefficients
The background compliance
is not constant, but undergoes anharmonic stiffening on cooling. The result is an almost linear decrease of
, at least down to 100 K, in the absence of other structural transitions. For example, the Young’s modulus
of polycrystalline SrTiO
increases of 2.9% per 100 K, while that of BaTiO
of ≳
per 100 K [
10]. This effect can be estimated by extending the elastic measurement well above
, in order to extrapolate
from a temperature region unaffected by the precursor FE fluctuations, where only the linear anharmonic stiffening is observed. In BaTiO
, the precursor fluctuations persist even 200 K above
. Similarly, in systems with relaxor-like characteristics,
should be extrapolated from above the Burns temperature. Nevertheless, the anharmonic stiffening is a small effect, less important than other sources of uncertainty.
Also the electrostrictive coefficients
Q are expected to increase slightly and linearly with temperature, since they also can be deduced by simple anharmonic models of inter-ionic potential [
16]. Indeed, they are found to be little [
17,
18] or mildly [
19] dependent on temperature, and composition in solid solutions, [
18,
20,
21] so that, for simple FE transitions as described by Equation (
11), one expects that the steplike softening below
remains nearly constant in a wide temperature range. Actually, one is not interested in the dependence of
Q on
T, because it contributes to that of
much less than
and
, and the use of Equation (
2) does not require the separate knowledge of
and
Q.
2.4. Fluctuations and Thermoelastic Effect
Fluctuations are source of uncertainty, since they are not included in the treatment for deducing
, and, though they affect also the piezoelectric and dielectric constants, there is no guarantee that
s,
and
d are equally affected, and therefore that Equation (
2) is valid also in the presence of fluctuations. The effect of fluctuations is schematically shown in
Figure 2c, together with that of thermoelastic relaxation [
22]
where
is the coefficient of thermal expansion and
the specific heat. The thermoelastic effect occurs only in measurements where the sample vibration induces a inhomogeneous hydrostatic component of strain, as is the case of flexure. Since
is generally very peaked around
, also
is peaked there. In BaTiO
, it can be estimated that
only very close to
and otherwise negligible. Even though a sizeable decay of
below
may be an intrinsic effect from a
more complicated than Equation (
11) (see
Section 2.6), the contribution of fluctuations is difficult or impossible to quantify, and it is better to not consider the temperature region close to
.
2.5. Flexoelectric and Surface Effects
Other mechanisms that may induce polarization in the PE phase are the flexoelectric and surface effects [
23], important in thin films, but generally neglected in ceramics. These effects are difficult to measure quantitatively, with experimental values of the flexoelectric coefficients sometimes exceeding the theoretical ones by two orders of magnitude [
23,
24]. Recently, it has been proposed that also the surface of ceramics may be polarized in the PE state, with
pointing outside the surface, so that the sample is overall unpoled, but under bending or inhomogeneous stress it exhibits enhanced effective flexoelectric coefficents [
24]. The polarization would be due to flexoelectric effect from the inhomogeneous stress of the grains near the surface, which are less constrained perpendicularly to the surface down to several micrometers.
These effects, together with fluctuations, result in the formation of spontaneous polarization above
with null space average, so that they cannot be revealed with macroscopic measurements of the piezoelectric effect, but they contribute to the elastic and dielectric softenings. The elastic softening stands out of the weak linear anharmonic background (
Figure 2b and Figure 4b), while the dielectric softening is certainly more difficult to separate from the Curie-Weiss rise of the permittivity toward
. It cannot be excluded that Equation (
2) partly holds also above
, but the piezoelectric coefficient
d in the PE state would be defined locally, with little practical value.
2.6. Additional Terms in the Expansion of and Multiple Ferroelectric Transitions
In general, a faithful representation of the FE transition may require additional terms in the power expansion of
, with respect to Equation (
11). For example, a first order transition requires that
and at least a positive term
. A FE transition followed by other transitions where the spontaneous polarization
changes direction, as in BaTiO
, requires many more terms, fully expanded in terms of the components of the vector
in order to introduce the anisotropy. The explicit expressions for
in such cases may become very cumbersome and, in the few cases where they have been obtained, they have been analyzed numerically for selected sets of the parameters [
20,
21,
25].
This is indeed the case of many FE perovskites of practical interest, for example with a morphotropic phase boundary between rhombohedraland tetragonal phases, as in PZT and PZT-based solid solutions, or with a sequence of FE transitions, as in BaTiO
and BaTiO
-based solid solutions. The additional FE instabilities enhance the various susceptibilities, and this is exploited for improving the piezoelectric properties [
26,
27]. It is also possible to show in a simplified manner, that the compliance, and hence also the piezoelectric coefficients, is peaked at transitions where the spontaneous polarization changes direction, due to almost linear coupling between the transverse component of
and a shear strain [
27,
28]. Therefore, until the additional FE transitions are all describable by a same
and the fluctuations and thermoelastic effects are not important,
will present additional peaks and anomalies, but Equation (
2) should remain valid.
2.7. Additional Structural Transitions
Equation (
2) is no more valid if other types of distortion modes are concomitantly present, in addition to the ferroelectric mode. The typical case in perovskites is an octahedral tilt mode, as in Na
Bi
TiO
(NBT) or PZT with less than 50% Ti. The tilt mode mode is described by an additional order parameter
, for example the angle of rotation of the octahedra. Then, the free energy
F contains additional terms with even powers of
, possible FE-tilt coupling terms containing both
P and
, and the coupling between
and stress
is exactly of the form (
8) of the electrostrictive coupling,
[
29,
30]. As a consequence, if the coupling between FE and tilt modes is weak, with two transitions well separated in temperature, the compliance will undergo two well separated softenings at the two transitions, and the previous treatment is still valid at the FE transition. It is impossible to distinguish which is the FE transition from the elastic measurement alone, but it is very easy from the dielectric susceptibility. In fact, the latter presents a huge peak of Curie-Weiss type at the FE transition, but only a small step at the tilt transition, if this occurs within the ferroelectric phase. This is due to the coupling between polarization and tilting, which, for symmetry reasons, to the lowest order is biquadratic [
30],
, and simply renormalises the
term in the free energy (
11) as
and the susceptibility in the FE phase, Equation (
13), from
to
. Below the tilt transition transition temperature
,
passes from 0 to
producing a step in
which is positive for cooperative (
) and negative for competitive tilt-polarization coupling, but very small compared to the FE peak.
As an example,
Figure 3 shows the compliance and dielectric susceptibility of PbZr
Ti
O
[
31], where the softening at
is smaller but of the same order of magnitude as that at
, but the anomalies in the dielectric susceptibility (notice the logarithmic scale) leave no doubt on the different nature of the two transitions: polar ferroelectric at
and nonpolar antiferrodistortive at
. In this case the softening below
should be purely piezoelectric down to
.
In NBT and NBT-rich perovskites, instead, the tilt and polar modes are strongly coupled and act together at the structural transitions [
32], so that there is no way of extracting the piezoelectric contribution from the softening [
33].
2.8. Depolarization Field and Influence of the Measurement Frequency
Up to now we neglected the presence of the depolarization fields
created by the polarization charges at the domain walls and grain boundaries. Let us first consider a uniformly polarized (monodomain) sample, where
. The application of an external probing stress
changes the polarization through the piezoelectric effect of
, which in turn changes the depolarization field by
; finally, through the converse piezoelectric effect, an additional strain is generated:
. This corresponds to a stiffening of the compliance
, known as piezoelectric stiffening [
34], and exactly cancelling the piezoelectric softening,
, in the case of uniform polarization [
11].
On the other hand, if the sample is unpoled and the vibration stress is uniform over a length scale including several domains,
, because
. This condition holds for forced or resonant vibrations of macroscopic samples, with sizes of several millimeters and more, but for ultrasound experiments, the wavelength must be sufficiently long to include several domains, otherwise, partial piezoelectric stiffening may occur, which reduces
. An extreme case are the Brillouin scattering experiments, where the probe acoustic waves at ∼50 GHz have
m. In this case, the wavelength probes regions within single domains with uniform
, and
is considerably or totally cancelled by
. A comparison between the compliance of PbZr
Ti
O
measured on a resonating bar at kHz and with Brillouin scattering [
11], shows that in the latter experiment only a small peak at
, due to fluctuations, is present, but no piezoelectric softening; instead, the tilt transition at
produces a step in both measurements, because it does not involve changes in polarization and hence no piezoelectric softening and stiffening.
In addition to the averaging over several domains in the unpoled state, a further reduction of the depolarization field may be effected by the free charges from ionized defects, which neutralize the polarization charges at the DW.
2.9. Extrinsic Domain Wall Contributions
All the above formulas take into account the intrinsic piezoelectric effect, without movement of the domain walls (DW); yet, the DW motion may considerably enhance the magnitude of the measured piezoelectric coefficients [
12,
35], especially for high stress levels and low frequencies. The relationship (
1) between dielectric, elastic and piezoelectric coefficients does not necessarily hold also for effective coefficients, that include the effect of the DW motion, even if this is within the limit of linear response. For simplicity, let us drop the tensor notation and consider appropriate components, e.g., the longitudinal components for an extensional or flexural vibration of a bar. Then Equation (
1) can be written as
where the superscript in
is dropped and it is put in evidence that only the part
of
arising from the FE transition, Equation (
13), must be considered (in practice for most ferroelectrics it is
). A similar relationship holds for the contributions from defects with elastic and dielectric dipoles, causing dielectric, anelastic and piezoelectric relaxations of amplitudes
,
and
[
27,
36,
37]:
However, even assuming that the linear contribution from DW motion satisfies this relation, the total effective coefficients, , and , do not necessarily obey . It can be shown that this is true only if , but there is no apparent reason for such a relationship to be true in general.
Therefore, Equation (
1) strictly holds only for the intrinsic piezoelectric coefficients, without contributions from DW motion, and it seems safe to limit the analysis of the piezoelectric activity in terms of elastic softening only to measurements with low stress amplitude and linear response, where the DW motion is reversible and minimal. At large stresses, where the elastic response is nonlinear, one probes the irreversible motion of DW, with pinning and unpinning processes, corresponding to the regions of the hysteresis loops with steep slopes. In these regions, the piezoelectric coefficients are anyway ill-defined.
2.10. Polycrystalline Average of Unpoled Ceramic: A Well Defined State
The elastic moduli of ceramic samples are angular averages of the single crystal material constants, and therefore Equation (
2) must be averaged over all domain orientations. The great advantage of using unpoled ceramics is that the orientations are really random, unless a particular texture exists, and therefore the orientation averages are simple. Instead, the orientational averages for poled samples are more complicated (see e.g., Ref. [
38]) and provide theoretical bounds to the maximum polarization, but the actual degree of poling of a sample depends on several factors: (i) the degree of non
DW motion and switching, which generally is well below 100%, due to internal mechanical constraints; (ii) DW pinning; (iii) non-uniform internal fields, due to porosity, microstructure and defects; (iv) insufficient poling field, due to instrumental limitations, and the geometry and resistance to fracture of the sample. As a consequence, it is possible that the degree of poling is considerably below the theoretical maximum, but it is difficult or impossible to establish how much.
It can be concluded that the unpoled state of an untextured ceramic is much better defined than its poled state, and therefore better suited to study the intrinsic behaviour of the material. The orientational averages of
in terms of the single crystal material constants for randomly oriented domains are easy to calculate, and the results for tetragonal, orthorhombic and rhombohedral symmetries are [
10]:
where the Voigt index notation is adopted.
Unfortunately, the single crystal materials constants are generally unknown for most ferroelectrics, and, in order to use the above formulas for quantitative comparisons, one must express the single crystal
and
coefficients in terms of the effective
and
measured on ceramics. This is done, for example, in Ref. [
38], and we report below the result for tetragonal symmetry, assuming full poling and adopting the original notation
instead of
2.11. Experimental Verification and Porosity
Up to now there is only one experimental verification of the validity of Equation (
2), based on ceramic BaTiO
[
10]. The reason it is difficult to carry out a quantitative verification is connected with the difficulty of finding a material with reliably known piezoelectric and dielectric tensors. In addition, the technique available in our laboratory requires samples shaped as bars longer than 2 cm (but DMA, RUS and ultrasonic experiments can be performed on much smaller samples). The ceramic samples must also possess the same physical properties as the crystals, apart from the unavoidable porosity, and this excludes the usual piezoelectric materials, whose properties are heavily dependent on composition, additives and preparation protocols. The material of choice is therefore pure BaTiO
, though it has the complication of strong FE fluctuations extending 200 K above
, and two additional FE transitions below
.
The comparison between
, Equation (
18), deduced from the Young’s modulus of ceramic bars of BaTiO
measured at kHz and the literature data for
and
is described and discussed in detail in Ref. [
10], and here only a brief account is given. A major problem in comparing materials constants measured on ceramics with single crystal data is porosity, whose effect can be estimated only with great uncertainty. In
Figure 4a the compliances of three different samples from different laboratories (sample 1 measured in Ref. [
39] and samples 2 and 3 in Ref. [
10]) differ of factors up to 4, even though the porosities from the Archimedes method differ at most of a factor of 3 and the dependence of
s on porosity is definitely nonlinear. In fact, the shape of the pores has a major influence and is difficult to assess. Nonetheless, by rescaling in
Figure 4b the softer
curves of the more porous samples on the stiffer curve of sample 1 well above
, one finds a fair agreement between all the curves also below
. This encourages one to use the data of porous ceramics even to evaluate the material intrinsic piezoelectric properties, based on the effect of porosity in the PE phase. Notice that, if one wants to deduce piezoelectric coefficients from the elastic softening and permittivity measured on porous ceramics, also the permittivity must be renormalised for taking into account the porosity, but the effects of porosity on the elastic and dielectric responses may be different, so that there is no guarantee that the same rescaling factor applies to
s and
. For reducing the uncertainty one should reduce the porosity.
The
has been extrapolated from >800 K, since already below that temperature its linear dependence is obscured by the rising precursor softening from the FE fluctuations and polarization induced by flexoelectric and surface effects. After subtraction of
one gets
(
Figure 4c), which can be compared with the averaged expression, Equation (
18), where the single crystal
and
are inserted. Even for BaTiO
, there are no accepted values for these parameters, and the two curves are calculated using values measured on a set of single crystals [
40], and values obtained from a Landau free energy, that reproduce many experimental data on BaTiO
[
41].
The agreement between the measured and calculated
is good only in the plateau of
below
and well above the transition temperature
between the FE tetragonal (T) and FE orthorhombic (O) phases. In this temperature region the fluctuations of the polarization should be minimal, and indeed both the calculated curves cross the experimental ones there. Notice that longitudinal fluctuations of the polarization are present near and above
, but transverse fluctuations are expected above
, where they are associated with the change of direction of the spontaneous polarization (
Figure 4d).
As discussed in
Section 2.6,
should include also the peaked softenings at
and
, and it is not clear if the fluctuations together with surface and flexoelectric effects may explain the difference between the measured and calculated curves, away from the plateau of
. It is true that also the calculated curves do not agree with each other, but there seems to be a systematic difference between the temperature dependencies of the measured and calculated piezoelectric softenings. The motion of domain walls may also be responsible for such deviations, but the analysis of the
and elastic energy loss curves measured at different frequencies [
42] suggests that the contribution of DW relaxation is negligible, at least in curves 2 and 3.
It can be concluded that, at least in a limited temperature region below where complications from fluctuations and multiple FE transitions are excluded, there is quantitative agreement between the piezoelectric softening measured on unpoled ceramic BaTiO and that calculated from the known piezoelectric and dielectric tensors.
In ferroelectrics with a single or at least more isolated FE transition, the situation should be more clear. For example, PbTiO
doped with Ca has only one transitions from C-PE to T-FE, and no tendency to octahedral tilting. Indeed, preliminary measurements (
Figure 5) of the Young’s modulus at kHz show a situation very close to the simple case in
Figure 2c. Further piezoelectric and dielectric measurements are in progress [
43] on samples from the same material, in order to make a quantitative comparison between the softening of
Figure 5 and Equations (
18) with (21)–(25).
2.12. Usefulness of the Elastic Assessment of the Potential Piezoelectric Properties of Unpoled Samples
The use of Equation (
2), and hence Equations (
18)–(
20), to extract the intrinsic piezoelectric coefficients
seems particularly promising in combination with Resonant Ultrasound Spectroscopy (RUS) experiments on unpoled ceramics, from which the full tensors
in the FE state and
in the PE state can be extracted. This type of experiments has been done for extracting the elastic tensor
of unpoled PZT-4 in the FE state [
44], and by piezoelectrically exciting a poled PZT-8 ceramic sample [
9], so obtaining the full piezoelectric and elastic tensors in the FE state. The latter method is certainly a considerable improvement with respect to the traditional methods requiring sets of different samples, but still the piezoelectric coefficients depend on the degree of poling. Measuring unpoled samples over the full FE and PE temperature range would avoid also this source of uncertainty. The analysis of the full resonance spectrum of a poled ceramic disc [
8] is similar to the piezoelectrically excited RUS [
9], though with a larger sample.
At present, few laboratories measure the full elastic tensor with the RUS technique, and the more common acoustic measurements adopt forced (Dynamic Mechanical Analyzer = DMA, torsional pendulum) or resonant vibrations of bars, or ultrasonic propagation in pellets. While the ultrasonic experiments allow longitudinal and transverse waves to be excited, the other methods provide only one type of modulus, Young’s or torsional. In these cases it is impossible to extract the three or more components of , but the piezoelectric softening can still provide useful information. This is particularly true when studying new materials, or the effect of new types of doping or preparation protocols. In these cases, it is possible that varying some composition or preparation parameter, renders the poling of the sample less effective or impossible, for example due to increased conductivity or coercive field. A comparison of the piezoelectric softenings of the series of unpoled samples would provide a reliable indication of the potential piezoelectric response, that each sample would have if it would be fully poled, without the need for actually poling it, and therefore before optimizing composition and process in order to make poling easier.
A possible course of action in similar instances is illustrated in
Figure 6.
Figure 6a shows three hypothetical compliance curves obtained by varying a material parameter, for example doping. Increasing doping lowers
and changes the amplitude of
, but the main differences between the magnitude of the compliances are due to different porosities. Therefore, in
Figure 6b the curves are normalized in order to overlap in the PE phase, where it is assumed that doping has little effect on the elastic properties. This is done by choosing a temperature
well above
, where
increases linearly, due to anharmonic effects, as explained in
Section 2.3. Then, each curve is normalized by
(
Figure 6b). In this manner, the comparison between different
is not or little affected by porosity. The validity of this procedure is confirmed by tests on BaTiO
(
Figure 4a,b). The anharmonic stiffening of
is generally linear with very good approximation down to 100 K, so that it is easy to extrapolate it in FE phase and obtain
. According to Equation (
2) or (
16), a measure of the piezoelectric response is provided by
It is adimensional, because one uses the normalized , in order to eliminate the influence of different porosities, but one may as well normalize the curves to that with lower , presumably from the denser sample: if is of the th curve, than one extracts its from and calculates with the correct dimensions. The point is that, in this context, the absolute value of is of little interest, since it does not correspond to any or , but to a combination of them; what is important is the evolution of , measured in the same manner for the whole series of samples, with changing doping or the process parameters.
Also the dielectric permittivity is influenced by porosity, but in opposite manner with respect to the compliance, being reduced. If
were reduced by the same factor as
s, the effect of porosity would be automatically cancelled in
, without any correction. Unfortunately, this is not necessarily true (see e.g., Equation (
18) in [
45]), so that a quantitative analysis would require the use of Equation (
26), where also
is rescaled in order to cancel the effect of porosity. This is not as easy as for the compliance, since the FE instability causes a Curie-Weiss peak in
, rather than a step below
, whose contribution extends to the PE phase. Therefore, rather than simply renormalising
with its value at some
, one should extract the high frequency limits
and use
in Equation (
26). For qualitative purposes, it seems appropriate to use
or simply
d, and use the variation from sample to sample in
as an indication of the uncertainty introduced by changes in porosity.
Figure 6c shows the permittivity curves
, necessary to obtain
. Only the part below
is useful, and, in case of high conductivity, it is convenient to use high enough a frequency, e.g., 1 MHz, to minimize the lossy conductivity contribution. The final result of this example in
Figure 6d shows that doping decreases
, but the piezoelectric response of the material is even increased, thanks to the increase of
with lower
. This type of information is obtained without poling the sample, and calculating an effective
according to Equation (
26). The conclusion is valid even if doping introduces free carriers or strong DW pinning, which make actual poling impossible. Yet, one knows that the particular type of doping improves the intrinsic piezoelectric properties, and further material engineering can be attempted, in order to make poling easier.
Figure 7 presents another hypothetical instance, where doping transforms the FE transition into charge/orbital order, similarly to the manganite or nickelate perovskites. Below the onset of these types of transitions, a stiffening rather than softening is observed [
46,
47], so that it is easy to see by simple inspection that the FE transition is suppressed by doping.