Next Article in Journal
Optimizing a Solution Heat Treatment by Increasing the Cooling Rate of Directional Solidification for Ni-Based Superalloys
Next Article in Special Issue
Optimisation of a Multi-Functional Piezoelectric Component for a Climbing Robot
Previous Article in Journal
Experimental Investigation of Engine Valve Train Friction Considering Effects of Operating Conditions and WPC Surface Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stress–Charge Nonlinear Physical Description and Tensor Symmetries for Piezoelectric Materials

by
A. F. Jaramillo-Alvarado
1,*,
A. Torres Jacome
1,
P. Rosales-Quintero
1,
H. Vazquez-Leal
2,*,
G. Diaz-Arango
3,
J. Huerta-Chua
3 and
J. Martínez-Castillo
4
1
Electronics Department, Instituto Nacional de Astrofísica, Óptica y Electrónica (INAOE), Luis Enrique Erro # 1, Tonantzintla, Puebla 72840, Mexico
2
Electronic Instrumentation Faculty, Universidad Veracruzana, Cto. Gonzalo Aguirre Beltran S/N, Xalapa, Veracruz 91000, Mexico
3
Tecnologico Nacional de Mexico, Instituto Tecnologico Superior de Poza Rica, Luis Donaldo Colosio Murrieta S/N, Arroyo del Maiz, Poza Rica, Veracruz 93230, Mexico
4
Research Center in Micro and Nanotechnology, Universidad Veracruzana, Boca del Río, Veracruz 94294, Mexico
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(9), 3432; https://doi.org/10.3390/ma16093432
Submission received: 9 October 2022 / Revised: 29 October 2022 / Accepted: 31 October 2022 / Published: 28 April 2023
(This article belongs to the Special Issue Piezoelectric Materials and Piezoelectric Robots)

Abstract

:
Nonlinear piezoelectric materials are raised as a great replacement for devices that require low power consumption, high sensitivity, and accurate transduction, fitting with the demanding requirements of new technologies such as the Fifth-Generation of telecommunications (5G), the Internet of Things (IoT), and modern radio frequency (RF) applications. In this work, the state equations that correctly predict the nonlinear piezoelectric phenomena observed experimentally are presented. Furthermore, we developed a fast methodology to implement the state equations in the main FEM simulation software, allowing an easy design and characterization of this type of device, as the symmetry structures for high-order tensors are shown and explained. The operation regime of each high-order tensor is discussed and connected with the main nonlinear phenomena reported in the literature. Finally, to demonstrate our theoretical deductions, we used the experimental measurements, which presented the nonlinear effects, which were reproduced through simulations, obtaining maximum percent errors for the effective elasticity constants, relative effective permittivity, and resonance frequencies of 0.79 % , 2.9 % , and 0.3 % , respectively, giving a proof of the potential of the nonlinear state equations presented for the unifying of all nonlinear phenomena observed in the piezoelectric devices.

1. Introduction

Piezoelectric materials have been used in several application fields because their performance and set of physical properties meet the requirements in a wide scope of applications. Since the discovery of the piezoelectric effect by the Curie brothers in the 1880s, these types of materials were mainly used in transduction applications, until the 1970s, when their implementation in radio frequency (RF) applications was developed [1], and currently, the semiconductor manufacturing process allows their use in applications where the transduction between mechanical and electric fields is mandatory at the micro-scale.
Amorphous piezoelectric materials are used in applications where miniaturization is not required, and for this reason, currently, crystalline piezoelectric materials dominate the market and industry, mainly with microelectromechanical system (MEMS) devices due to the reproducibility of their physical and system properties. Consequently, the research on these materials is focused on crystalline composites that have high chemical resistance, high breakdown voltages, and high rigidity for RF applications.
Since its discovery, several fabrication techniques have been developed to obtain piezoelectric materials, where the chemical-based techniques have been of interest due to the requirements of thin-film technologies [2]. Deposition techniques such as metal–oxide chemical vapor deposition (MOCVD) [3] and chemical solution deposition (CSD) [4] are current research topics. Furthermore, there are CMOS-compatible deposition techniques, since these processes have low fabrication temperatures, such as sputtering-based techniques, which can obtain high levels of crystallinity [5,6], being an ideal fabrication process to apply the nonlinear phenomena of piezoelectric materials in a new scope of applications [7,8,9,10].
Currently, the main applications of piezoelectric materials are embedded in the MEMS scope, because they use the accurate transduction capability to implement them in several types of applications such as micro and nano-resonators [11,12], energy harvesters [13], accelerometers [14], wearable devices [15], micro- and nano-actuators [16], and sensors for gasses [17] and electrostatic charge [18]. In general, the applications cited share demanding requirements such as low power consumption, high sensitivity, accurate transduction, great chemical resistance, and good enough electrical and mechanical properties, where all of these conditions are met by piezoelectric materials. The modeling of the mentioned devices using the linear description of traditional state equations [19] gives acceptable errors by its predictions; nevertheless, under relatively high electric fields (> 10 6 V/m) and deformation, the physical behavior of materials is not predicted correctly [7,20], and the need for a complete first-principles physical description of the nonlinear phenomena for piezoelectric materials emerges as a mandatory tool for new designs in demanding applications of the industry, such as the Fifth-Generation of telecommunications (5G) and the Internet of Things (IoT).
There are applications that use the nonlinear properties with the same targets as the linear applications exposed above such as actuators [21,22], energy harvesters [23], sensors [24], memories [25], and tunable devices [7,26,27]. In all of these works, the physical and electrical behavior of the system is explained through mathematical models [10,28,29,30,31] or first-principles deductions (a specific thermodynamic formulation) [32,33,34,35], where the models are only valid for a specific geometry disposition or layer stack, while the physical formulations are general, but very difficult to solve analytically. The case of the hysteresis nonlinear effect is a special topic since its behavior has remnant fields after time; its formulation in deformation–charge form and micro-mechanical modeling was exposed in [36,37] respectively. In the models cited, the core concept used is the algebraic or complex expansion of the material parameters, resulting in adjustments of the macroscopic magnitudes of the physicalsystem, e.g., resonance frequency, effective material constants, lumped elements of equivalent circuits, and quality factor, among others. All of these results produce an imbalance of the state equations, being the core problem of models for nonlinear piezoelectric applications since the introduction of adjustment parameters in the material constants reproduces the macroscopic behavior of the phenomena; nevertheless, the physical behavior of the effect is not described by the equations. In contrast, the first-principles formulations are based on the balance of the microscopic states of the physicalsystem, resulting in the prediction of the physical behavior of the macroscopic states, being a complete physical description of the nonlinear effects where the state equations remain balanced; consequently, the solver’s calculus is more difficult and time consuming. Due to this, to include this type of device within the integrated circuit (IC) industry, mathematical tools are needed that allow fast design and manufacturing processes, such as the finite-element-method (FEM)-based design accompanied by compatibility with the main IC fabrication processes, such as CMOS, PD-SOI, FD-SOI, and FinFET. In summary, a thermodynamic formulation with easy implementation in leading FEM simulation software (e.g., COMSOL Multiphysics and COVENTOR) is mandatory for the inclusion of nonlinear piezoelectric devices within the semiconductor industry, and that set of demanding characteristics for the mathematical and physics tools are contained in the formulation presented in this work.
To simulate the nonlinearities and physical behavior of the piezoelectric materials, it is necessary to know the nonlinear state equations with an easy methodology to include them in the FEM solver’ calculus; consequently, the symmetry structure of high-order tensors must be given as well. Despite this, the methodologies found in the literature to implement nonlinear behaviors in leading FEM simulation software are complicated to carry out, and at the same time, the symmetry structures cannot be found (only some components for a few types of materials [32]). For these reasons, the nonlinear applications reviewed cannot be explained by a unified set of equations with known tensor structures, making the industrial adoption of these types of applications more difficult despite their advantages.
Taking into account the above discussion, in this work, we present a complete physical description of the nonlinear behavior of piezoelectric materials, obtained through the deduction from first-principles of the nonlinear state Equations (until third-order phenomena), the transformation laws required, and the symmetry structures of the tensors, for each of the thirty-two point groups of symmetry (all types of crystalline materials). Furthermore, a methodology with an easy way to implement the state equations and high-order tensors components in the main FEM simulation software is presented, allowing designing and manufacturing devices that can be used in the 5G, IoT, and RF application scopes. Finally, this work gives the MEMS scientific community all the mathematical and physics tools needed to research new types of applications and optimizations for nonlinear piezoelectric devices.

2. Stress–Charge Nonlinear Formulation

A suitable thermodynamic representation for including the nonlinear effect within FEM simulators is the stress–charge formulation due to the characteristics of direct solvers, since the physical behavior of the electrical permittivity and elasticity constants are well-known parameters of crystalline materials; in the literature can be also found references to perform the energy and dissipation calculus [38].
The following deductions are focused on crystalline materials. The theoretical development starts from first-principles using the Voigt form for mechanical tensors, the Einstein sum convention, and the recommended notation for point groups of symmetry by the International Union of Crystallography (IUCr) [39]. The entropy and the temperature contributions were neglected due to the solid phase of materials, the low power dissipation (around 10 mW/mm 2 ), and the nonlinear perturbative operation regime of the devices. In the next sections, we discuss the experimental limits that govern the theoretical development presented.
From the eight possible formulations [19], we used the thermodynamic potential of the electric Gibbs function [40], the total differential of which is defined for the piezoelectric effect as
d G 2 = D k d E k + T λ d S λ ,
where D k , E k , T λ , and S λ are the electric displacement vector, electric field, stress field, and deformation field, respectively. Therefore, considering the properties of the total differential of a multivariable function, the total differentials for dependent variables are
d T λ = C λ μ d S μ e k λ d E k , d D i = ϵ i j d E j + e i μ T d S μ ,
where e k λ T are the piezoelectric coefficients, C λ μ the elastic constants, and ϵ i j the electrical permittivity.
To deduce the nonlinear formulation, we expanded the tensor coefficients in Equation (2) through a Taylor series centered at zero and took into account that the dependent variables are a function of S λ and E k , then the elastic constants are
C λ μ = T λ S μ = T λ S μ 0 + 2 T λ S μ S ν 0 S ν + 2 T λ S μ E k 0 E k ,
where the high-order derivatives were measured at constant deformation and the electric field equals zero. Through an analogous procedure, we can obtain all coefficient tensors of Equation (2) as a function of S λ and E k .
Since Equation (1) is a total differential, we have
G 2 S λ = T λ , and G 2 E k = D k ,
then, considering that G 2 is a physical magnitude, it is continuous, has an exact differential, and has derivatives up to third-order, and knowing the mixed derivatives equivalence, we obtain
3 G 2 E k S λ S μ = 3 G 2 S λ E k S μ = 3 G 2 S λ S μ E k
Applying Equations (4) and (5), we can define g λ μ k as
2 T λ E k S μ = 2 T λ S μ E k = 2 D k S λ S μ = g λ μ k
Considering the other tensors’ coefficients in Equation (2) and applying the same procedure for Equations (3) to (6), we define the remaining high-order tensors as
2 T λ E j E k = 2 D k S λ E j = 2 D k E j S λ = q k j λ ,
and
2 T λ S μ S ν = t λ μ ν , 2 D i E j E k = r i j k .
Subsequently, by replacing Equations (3), (6), (7), and (8) in (2), we obtain
d T λ = C λ μ d S μ e k λ d E k + g λ μ k d S μ E k + q λ j k E j d E k + t λ μ ν S ν d S μ , d D i = ϵ i j d E j + e μ i T d S μ q i j μ d E j S μ g λ μ i S λ d S μ + r i j k E k d E j
After integrating Equation (9), we finally obtain the nonlinear state equations for the piezoelectric effect considering effects up to third-order
T λ = C λ μ S μ e k λ E k + t λ μ ν 2 S μ S ν + g λ μ k S μ E k + q j k λ 2 E j E k , D i = ϵ i j E j + e i μ T S μ + r i j k 2 E j E k q i j λ E j S λ g λ ν i 2 S λ S ν ,
having
q i j λ 2 E j E k = n = 1 3 q n n λ 2 E n E n
In Equation (10), given the equivalence between the Voigt and traditional mechanical notation, an algebraic factor is not needed; this means
t λ μ ν t i j k l m n , g λ μ m g i j k l m , q j k λ q j k l m λ , μ , ν [ 1 , 6 ]
Equation (10) describes how the exchange of the coupling fields’ magnitudes is performed through the direct and converse piezoelectric effect, while the nonlinear contributions generated by the relatively high electric and deformation fields were considered. These conditions subject the material to mechanical and electrical stress, producing variations in all material parameters, as will be shown in the nonlinear effects section.
In this context, t λ μ ν is the contribution to the stress field due to strong deformations taking importance in the plastic operation regime. Furthermore, t λ μ ν relates the orthogonal deformations S μ and S ν that produce the change of the stress field with respect to the linear approximation. Analogously, r i j k is a correction term for the electric displacement vector as function of very high electric fields E j and E k , so this tensor governs the dielectric polarization when | E i | is around 10 9 V/m. g λ μ k is responsible for the elasto-electric effect (in the literature, also known as nonlinear electrostriction and the electro-elastic effect), since its contribution to the stress field g λ μ k S μ E k provokes an augmentation of the effective elasticity constants, producing a stiffening of the material. In the same way, q i j λ E j S λ contributes to the electric permittivity due to the strains S λ , and it is responsible for the change in the effective permittivity of a material subject to relatively high electric fields. Finally, The last quadratic terms of Equation (10) are a contribution to the stress and electric displacement field, modifying the value of the coupling piezoelectric coefficients e k λ and e i μ T , respectively.
With this approach, the state equations presented remain balanced, while considering the nonlinear effects, and therefore, the physical behavior of the microscopic and macroscopic states of the physics system are predicted correctly.

3. Transformation Laws

To obtain the symmetry structure of any tensor, we need to know the transformation laws and the symmetry generators a i j of each crystal type (point group of symmetry). Taking into account the recommended notation for crystal classes and point groups by IUCr [39], a i j belongs to special orthonormal group S O ( 3 ) , since it represents a generic 3D rotation. Furthermore, the transformation laws for the high-order tensors must meet the constraints of the positive energy and generate stable states for the system (e.g., the vanishing of the total torque about the origin), and their symmetry structure must only depend on the point group of the material. Then, to deduce the transformation laws for each high-order tensor, we start with the example of the calculus of the transformation law for the electrical permittivity of any material. The transformations laws for the electric displacement vector and electric field are
D i = a i j D j , E i = a i j E j ,
where the superscript means a transformed magnitude. Now, knowing the law for the polarization of a material:
D i = ϵ i j E j ,
the target is to obtain an equivalent equation in terms of transformed magnitudes, so using Equation (13), we obtain
D i = ϵ k j a i k ( a l j ) 1 E l ,
where the transformation law for electrical permittivity is deduced from the symmetry condition, which means that, after transformation, the tensor form (structure) remains invariant:
ϵ i l = ϵ k j a i k ( a l j ) 1
For the q i j λ high-order tensor, we need the transformation law for the deformation field:
S μ = N μ ν S ν ,
where N μ ν is a function of the symmetry generator a i j [38]. Then, using
D k = q k j λ E j S λ ,
we obtain
D i = a i k q k j λ ( a l j ) 1 ( N μ λ ) 1 E j S λ ,
where the transformation law obtained for the q i j λ tensor is
q i l μ = q k j λ a i k ( a l j ) 1 ( N μ λ ) 1
Through an analogous deduction, the transformation laws for the nonlinear tensors in the Equation (10) can be obtained, and they are shown below:
t λ μ ν = t α β γ M λ α N β μ 1 N γ ν 1 , g λ μ k = g β ν m M λ β N ν μ 1 a m k 1 , q j k λ = q l m β M λ β a l j 1 a m k 1 , r i j k = r l m n a i l a m j 1 a n k 1
At this point, the reader can notice that there are two ways to obtain the transformation laws, one per state equation in (10). Both ways have equivalent results knowing the properties of a symmetry generator a i j (belongs to the S O ( 3 ) group) and the M and N matrices [38]:
( a i j ) 1 = ( a i j ) T = a j i , a n d ( N 1 ) i j = ( M T ) i j

4. Symmetry Structure for High-Order Tensors

The structure of the tensors can be calculated from Equation (21), the generator symmetry a i j for the specific crystal type, and a last mathematical constraint
2 D k S λ S μ = g λ μ k = g μ λ k , 2 T λ E k E j = q k j λ = q j k λ , 2 D i E j E k = r i j k = r i k j , 2 T λ S μ S ν = t λ μ ν = t λ ν μ ,
based on the mixed derivatives theorem. Then, selecting a transformation law from Equation (21) for the desired high-order tensor structure, a specific point group of symmetry (e.g., 6 mm), and applying Equation (23) in the transformation law selected, we obtain an undetermined algebraic linear system, which, after being solved, we obtain the structure of the tensor in terms of a few unique components, which represents the contribution of the specific tensor to the nonlinear behavior of the piezoelectric material. This procedure to obtain the symmetry structure of the high-order tensors was tested through obtaining the symmetry structure of known tensors for the thirty-two point groups; specifically, the elasticity constants, electrical permittivity, piezoelectric coupling coefficients, and r i j k tensor were reproduced; the last one is the only high-order tensor, whose complete symmetry structure has been published [41].
Table 1 presents a first approximation of the high-order tensors for two common piezoelectrics, P Z T 5 H and aluminum nitride ( A l N ), which belong to the 4 mm and 6 mm point groups, respectively. These results were obtained after reviewing the literature and noticing that, when an excitation signal provokes the appearance of the nonlinear effects [7,10,33], we suppose a variation around 2 % for dependent variables with respect to the linear approximation.
Table 2 presents the symbols and particular numeration for the thirty-two point groups of symmetry; this numeration is used in the tables where the tensor structures are shown, and all high-order tensor components are introduced only by subscripts. The symmetry structures of tensors q j k λ and g λ ν i are shown in Table 3 and Table 4, respectively; the first column contains the component of the high-order tensor and the following columns its corresponding value for a specific point group. The symmetry structure of q j k λ depends only on the Laue symmetry group. All types of crystalline materials have q j k λ and t λ μ ν different from zero in at least one component, and the g λ ν i and r i j k tensors are null if the material does not exhibit linear piezoelectric behavior (this means they are centrosymmetric crystals), with the only exception of point group 432, where g λ ν i is not zero and r i j k remains null. The symmetry structure of t λ μ ν for some point groups is shown in Appendix A in Table A1; the point groups not included are H I and R I I ; they need a separate complete analysis, and due to this, they are postponed for a future work.

5. Nonlinear Effects of Piezoelectric Materials

The nonlinear phenomena of the piezoelectric effect take importance when the material is subject to relatively high electric fields and strong deformations, and its consequences can be classified into two categories. First is the change of the mechanical and electrical properties such as the change of electrical permittivity, elasticity constants, and piezoelectric coupling coefficients. Second is the behavior variation of the physicalsystem response due to the modified material parameters, in particular the arising of the hysteresis behavior, changes in the electromechanical coupling factor, a shift of the resonance frequency, and the modification of the capacitance of the devices.

5.1. Variation of Mechanical and Electrical Properties

The change of the electrical permittivity in a piezoelectric material is produced by strong deformations or high temperatures [42] and can be induced by exciting the material with a relatively high electric field, the converse piezoelectric effect producing the deformations needed. This physical behavior can be observed from the nonlinear state equations, since the tensor q j k λ in the D i equation modifies the total polarization, and this can be integrated into a unique term with the electrical permittivity as follows:
ϵ i j e f f = ϵ i j q i j λ S λ ,
where ϵ i j e f f is the effective electrical permittivity.
The change of the elasticity constants is due to exposing the material to relatively high electric fields, which provokes a change in the interatomic electronic forces due to deformations, consequently causing a variation of the stiffness of the material. Furthermore, this phenomenon is included in the state equations through the modification of the total stress induced by the contribution of the deformations and can be formulated as effective elasticity constants as
C λ μ e f f = C λ μ + g λ μ k E k
Finally, due to the power balance of the nonlinear state equations, the variation of the piezoelectric coefficients is a consequence of the imbalance produced by the two last phenomena discussed, where the variation in the transduced power produced by the first nonlinear effect is compensated by the second, then the effective piezoelectric coupling coefficients are
e j λ e f f = e j λ q k j λ 2 E k , e i μ T e f f = e i μ T g λ μ i 2 S λ ,
where e j λ e f f and e i μ T e f f must be used in the T λ and D i state equations, respectively.

5.2. Change Response of the PhysicalSystem

The literature shows how some piezoelectric devices that are subject to relatively high electric fields have a shift of their resonance frequency; this is produced by the variation of the electrical permittivity and the elasticity constants phenomenon explained before [20,43,44,45]. Generally, the resonance frequency of a piezoelectric resonator depends on its geometric length and the specific material, often calculated as
f r = 1 2 λ d C λ μ D e f f ρ ,
where λ d is the wavelength of the device, ρ the density of the material, and
C λ μ D e f f = C λ μ e f f + ( e k λ ) 2 ϵ i j ,
s the effective elasticity constant (in some cases, it can be called the effective Young’s modulus), requiring all subscripts to match the main oscillation mode of the studied device. If we analyze the relative change of the resonance frequency (Equation (27)), we can obtain
d f r f r = d C λ μ D e f f 2 C λ μ D e f f d ρ 2 ρ d λ d λ d
From Equation (29), it can be noticed how the shift of the resonance frequency is a consequence of the changes of the effective elasticity constants, density, and wavelength of the device, where the last two terms are well known, so they can be neglected [31], because the piezoelectric materials are non-centrosymmetric crystals and the transverse/longitudinal dilatation does not provoke the measured order of magnitude for nonlinear effects.
The variation in the capacitance of the devices is explained through the change in the electrical permittivity phenomenon. Normally, the value of the capacitance of devices that have a dielectric as a piezoelectric material is
C e f f = ϵ i i e f f A t ,
where C e f f is the effective capacitance of the device, A is the electrodes’ contact area, t is the thickness of the piezoelectric, and ϵ i i e f f is the effective electrical permittivity over the i axis. Performing an analogous relative variation analysis, then
d C e f f C e f f = d ϵ i i e f f ϵ i i e f f + d A A d t t ,
where the last two terms can be neglected, inclusive of the nonlinear effects regime. This is due to the absolute displacement of particles because the order is 0.1 Å (theoretical prediction) for the piezoelectric materials under these conditions; hence, d A and d t are not comparable with d ϵ i i e f f , since its variation is of the order of thousandths [20].
The calculus for the electromechanical coupling factor k e f f 2 is defined for resonant applications of piezoelectric materials and depends on the oscillation mode of the device, material properties, and specific device geometry. k e f f 2 is a measure of the exchange of power transduced between the mechanical and electrical fields, and for the most common devices, it has an expression of the form [38,46]
k e f f 2 = ( e x 5 e f f ) 2 C 44 e f f ϵ x x e f f ,
for a device with Z-shear oscillation mode and a wave in the X-propagation axis. In Equation (32), the most significant variation, following the discussion above, comes from the effective elasticity constants [31]; hence, when a shift of the resonance frequency occurs, the electromechanical coupling factor increases its value, while the effective elasticity constants decrease. Therefore, for applications where power transduction is the main goal (e.g., energy harvester, microphones, etc.), it can be deduced using Equations (28) to (32) that a negative external electric field increases the performance of the device [13].
The appearance of the hysteresis behavior in the piezoelectric materials as a soft ferroelectric effect is a well-known phenomenon; this is produced by two main causes, the alignment of the dipoles in the unit cells of the material with respect to an external electric field and the change of the domain walls [47,48,49,50]. The change in the domain walls produces a spontaneous strain, inducing additional stress and polarization, and the alignment of the unit cells corresponds to the spontaneous polarization field having a contribution to the strain field. The correspondence between a cyclic electric field and the response of the polarization field and deformation field results in a hysteresis loop and butterfly loop, respectively [51]. The thermodynamic formulation presented only considers the spontaneous strain produced by high electric fields induced due to the inverse piezoelectric effect (last term of the T λ state equations), but it is only one of the theoretical treatments needed for a complete description of the hysteresis behavior.
In the context of all the experimental evidence exposed and discussed, the several nonlinear effects in piezoelectric materials take importance in different regimes. We describe the limit of the formulation presented as a function of the importance of the high-order tensors for their respective regime of operation, where the nonlinear electric contributions take precedence over the mechanical ones [52,53]. Taking as independent physical magnitudes the electrical and deformation field, if the material is subject to excitements of an order of magnitude under 10 6 V/m and 10 6 , respectively, the linear formulation would be enough. From there, the g λ ν i and q j k λ tensors must be taken into account, where the r i j k domain makes electric contributions with electric fields above 10 9 V/m, and t λ μ ν is only required starting from the plastic regime. Finally, the hysteresis behavior appears as a soft ferroelectric effect for some specific piezoelectric crystals with excitements of the order above 10 6 V/m and 10 4 for the electric and deformation fields, respectively. It is necessary to bear in mind that, currently it is not clear what the starting point for the hysteresis behavior for any piezoelectric material is, since this effect belongs to the point group of the material or is induced by very high electric or deformation fields. The last discussion only applies to crystalline piezoelectric materials that are subject to nonlinear perturbative excitements.

6. Experimental Validation: Simulation

To validate the theoretical development performed in this work, we chose a reference that showed the nonlinear behavior of the piezoelectric devices under a relatively high electric field, since this is the simplest method to induce the nonlinear phenomena. The reference to reproduce is [28], where a solidly mounted resonator (SMR) was fabricated and characterized using A l N as a piezoelectric material; the fabrication details can be found in the reference. Measurements were performed with an Advantest R3767 S-Parameter analyzer, with the DC offset generated by a Keithley K327 and connected through a bias-T, and finally, the data acquisition was performed with the Picoprobe ECP18 GS-200 PP. Therefore, to implement the nonlinear state equations deduced, we can start neglecting the contributions of the t λ μ ν and r i j k tensors, since the operation regime and nonlinear behavior of the device are dominated by the linear description and the tensors g λ ν i and q j k λ [52,53].
To include the nonlinear state equations within the simulations in an easy way, we chose the following formulation:
T λ = C λ μ e f f S μ e k λ e f f E k , D i = ϵ i j e f f E j + e i μ T e f f S μ ,
where we used Equations (25) and (26), since this implementation included the power balance between the physics magnitudes of interest ( T λ and D i ) within the FEM simulator. Generally, the FEM simulators allow us to set the electrical relative permittivity as an input parameter, so we rewrite ϵ i j e f f as
ϵ i j e f f = ϵ 0 ϵ i j r e f f = ϵ 0 ϵ i j r q i j λ r S λ ,
where ϵ 0 is the vacuum electrical permittivity, ϵ i j r is the relative permittivity in the linear regime, ϵ i j r e f f is the effective relative permittivity, and the last term is defined as
q i j λ r = q i j λ ϵ 0
The SMR devices have the main oscillation mode, which confines the mechanical waves within the device; based on this, the algebraic tensor development of Equation (33) results in the only components of g λ ν i and q j k λ that must be taken into account to be g 333 , q 331 r , and q 333 r . The symmetry structure taken from Table 3 and Table 4 was the 6 mm one, since the piezoelectric material was A l N . The values obtained for the high-order tensors from the simulations were
g 333 = 80 N / V m , a n d q 331 r = q 333 r = 120
The SMR device was powered by an S-Parameter analyzer with a DC bias added with a bias-T through the signal probe of the RF microprobes. To reproduce the experimental setup, the simulated device was connected to an RF source with a DC voltage overlap, to calculate the whole interest frequency spectrum as a function of the DC bias. Figure 1a shows a transversal cut of the device simulated. In Figure 1b, the impedance of the device simulated for different DC biases is presented, and there, we can observe how the frequency response depends on the external DC electric field (EDEF), since it augments the stiffness of the material when positive voltages are applied, increasing the elasticity constants’ values, and consequently, the resonance frequency increases as well; this behavior’s prediction was performed by Equation (27). The effective elasticity constant C 33 D e f f obtained from the measurements and simulations is shown in Figure 2a, where the maximum percent error obtained was 0.79 % . The stiffening of the material was proportional to the EDEF due to the negative sign of g 333 ; consequently, the resonance frequency had the same dependency. This can be observed in Figure 2b, where the resonance frequencies measured and simulated, for several values of the EDEF, are presented; there, the maximum percent error obtained was 0.3 % . The behavior obtained from the measurements and simulations for the relative effective permittivity is exposed in Figure 3a, where the linear inverse dependence between the EDEF and the permittivity can be observed, as predicted by Equation (24); the maximum percent error obtained was 2.9 % . As can be expected, the slopes in Figure 2a and Figure 3a correspond to the values of g 333 and q 331 , having the correct prediction for the trend behavior observed experimentally. Finally, the behavior of the electromechanical coupling factor is shown in Figure 3b, where the predictions of Equations (28) to (32) are corroborated, since the maximum value for k e f f 2 was obtained under negative voltages for the EDEF; this behavior was not reported by the experimental reference, but it was obtained from the simulations. In Table 5 is shown the average and maximum percent errors obtained from the simulations with respect to the measurements; there, the maximum percent error for the effective elasticity constants, effective relative permittivity, and resonance frequencies were 0.79 % , 2.9 % , and 0.3 % respectively. These errors were caused by the difference between the physical material parameters and those used in the simulations; furthermore, the inaccuracy in the extremes of the values of the EDEF was due to the divergence problems that are present in the direct solver of the FEM software; this can be observed mainly in Figure 1b and Figure 3a. Nevertheless, in the scope of the simulations performed, the maximum percent error obtained for any material parameter or physicalsystem parameter was 1.1 % ; this shows the accuracy of the state equations presented to predict the main nonlinear phenomena of piezoelectric materials through a unified set of state equations, which can be included in FEM simulators easily.

7. Conclusions

In this work, we presented the nonlinear state equations for piezoelectric materials obtained from first-principles, conserving the power balance exchange between the dependent physical magnitudes T λ and D i and having a unified set of equations that predicts the behavior of the nonlinear phenomena. Furthermore, we showed how we obtained the transformation laws and the symmetry structures for the r i j k , g λ μ k , and q i j λ tensors, while the calculation procedure was demonstrated with known tensor structures ( C λ μ , e λ k , and ϵ i j ). The physical connection and explanation for the nonlinear phenomena experimentally observed in the piezoelectric material were exposed, remarking on the excitement conditions that made each phenomenon appear, where, under an external DC electric field less of than 10 9 V / m , the nonlinear phenomena were dominated by the change in the relative effective permittivity and effective elasticity constants through the g λ μ k and q i j λ tensors. The elastoelectric effect does not appear in non-piezoelectric materials ( g λ μ k is null), but the electrostrictive effect and nonlinear piezoelectric behavior remained within the material since q i j λ was not zero, except for point group 432, where q i j λ = 0 and g λ μ k was not null. A fast methodology for the implementation of the nonlinear state equations in the main FEM simulation software was exposed and demonstrated; this was carried out through the reproduction of an experimental reference, where the main nonlinearities of the piezoelectric effect were measured. The maximum percent errors obtained from the simulations were 0.79 % , 2.9 % , and 0.3 % for the effective elasticity constants, relative effective permittivity, and resonance frequencies. This proved the effectiveness of the nonlinear stress–charge formulation presented, taking into account that the symmetry structure of each high-order tensor was shown (Table 3, Table 4 and Table A1). The design and simulation in the leading FEM simulators of nonlinear piezoelectric devices with a complete physical description are now possible.

Author Contributions

Conceptualization, A.T.J., P.R.-Q. and J.M.-C.; methodology, A.T.J. and P.R.-Q.; review and editing, H.V.-L., J.H.-C. and J.M.-C.; software, G.D.-A.; investigation, formal analysis, data curation and validation, A.F.J.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article.

Acknowledgments

A.F. Jaramillo Alvarado thanks Instituto Nacional de Astrofísica, Óptica y Electrónica (INAOE) for its support through the scholarship Beca Colaboración.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The symmetry structure of t λ μ ν as function of the Laue group is shown in Table A1, where the expressions used are exposed in Table A2. The components of the tensor are listed as three numbers in a row, which correspond to their subscripts. The Laue groups H I and R I I need a separate complete analysis; therefore, their structures are delayed for a future analysis.
Table A1. Symmetry structure of sixth-ranked tensor t λ μ ν for the nonlinear behavior of the stress field due to the strong deformation field. The expressions used are shown in Table A2.
Table A1. Symmetry structure of sixth-ranked tensor t λ μ ν for the nonlinear behavior of the stress field due to the strong deformation field. The expressions used are shown in Table A2.
Laue Group
Comp.NMOTIITIRIHIICICII
111111111111111111111111111111
121121121121121121121121121121
131131131131131131131131121131
1411410000141000
1511511510000000
161161001610016100
221221221221221221221221221221
231231231231231231231231231231
2412410000241000
2512512510000000
261261002610026100
331331331331331331331331221331
3413410000341000
3513513510000000
361361003610036100
441441441441441441441441441441
451451004510045100
4614614610000000
551551551551551551551551551551
5615610000561000
661661661661661661661661551661
112112112112221221AA221331
122122122122121121BB121131
132132132132231231231231231231
1421420000C000
1521521520000000
16216200−26100−16100
222222222222111111DD111111
232232232232131131131131121121
2422420000E000
2522522520000000
26226200−16100-26100
332332332332331331331331221221
3423420000−341000
3523523520000000
36236200−36100−36100
442442442442551551551551551661
45245200−45100−45100
4624624620000000
552552552552441441441441441441
5625620000F000
662662662662661661GG551551
113113113113113113113113221221
123123123123123123123123231231
133133133133133133133133121121
1431430000143000
1531531530000000
1631630016300000
223223223223113113113113221331
233233233233133133133133121131
2432430000−143000
2532532530000000
26326300−16300000
333333333333333333333333111111
34334300000000
3533533530000000
36336300000000
443443443443443443443443551551
45345300000000
4634634630000000
553553553553443443443443551661
5635630000143000
663663663663663663HH441441
1141140000114000
1241240000124000
1341340000134000
144144144144144144144144144144
1541540000015400
1641641640000000
2242240000I000
2342340000−134000
244244244244244244244244244244
254254002540025400
2642642640000000
33433400000000
344344344344344344344344244344
354354003540035400
3643643640000000
4444440000444000
4544544540000000
4644640046400U00
5545540000−444000
564564564564564564JJ564564
6646640000124000
1151151150000000
1251251250000000
1351351350000000
14514500−25400−25400
155155155155244244244244244344
1651650000K000
2252252250000000
2352352350000000
24524500000−15400
255255255255144144144144144144
2652650000L000
3353353350000000
34534500−35400−35400
355355355355344344344344244244
3653650000134000
4454454450000000
4554550000−444000
465465465465564564JJ564564
5555555550000000
56556500−46400−U00
6656656650000000
1161160011600Y00
12612600000X00
1361360013600−36100
1461461460000000
1561560000−E/2 + 241/2000
166166166166166166D/2−111/
4−221/4
D/2−111/
4−221/4
244244
22622600−11600−3Y−4 × 26100
23623600−1360036100
2462462460000000
2562560000−C/2 + 141/2000
266266266266166166N/2N/2244344
33633600000000
3463463460000000
3563560000341000
366366366366366366OO144144
446446004460045100
456456456456456456PP564564
4664660000M/2000
55655600−44600−45100
5665665660000000
66666600000X00
Table A2. Abbreviated expressions used in Table A1 to show the t λ μ ν components.
Table A2. Abbreviated expressions used in Table A1 to show the t λ μ ν components.
ExpressionEquivalenceExpressionEquivalenceExpressionEquivalence
A121/2−111/4 + 3/4 × 221 + 661J(244−144)/2R−151/2−251/2−461
B111/4 + 121/2 + 221/4−661K(114 + 3 × 124)/2S251/2−151/2
C561−241/2−141/2L(114−124)/2T141/2−241/2
D3/4 × 111 + 121/2−221/4 + 661M241/2−141/2 + 561U(154−254)/2
E−141/2−241/2−561N(3 × 111)/4−121/2−221/4−661V−164/2−(3 × 264)/2
F(141−241)/2O(131−231)/2W264/2−164/2
G(111−2 × 121 + 221)/4P(551−441)/2X(261−161)/2
H(113−123)/2Q461−251/2−151/2Y−(161 + 3 × 261)/2
I−(114 + 2 × 124)

References

  1. Berlincourt, D. Piezoelectric Crystals and Ceramics. In Ultrasonic Transducer Materials; Mattiat, O.E., Ed.; Springer: Boston, MA, USA, 1971; pp. 63–124. [Google Scholar] [CrossRef]
  2. Shilpa, G.D.; Sreelakshmi, K.; Ananthaprasad, M.G. PZT thin film deposition techniques, properties and its application in ultrasonic MEMS sensors: A review. IOP Conf. Ser. Mater. Sci. Eng. 2016, 149, 012190. [Google Scholar] [CrossRef]
  3. Micard, Q.; Condorelli, G.G.; Malandrino, G. Piezoelectric BiFeO3 Thin Films: Optimization of MOCVD Process on Si. Nanomaterials 2020, 10, 630. [Google Scholar] [CrossRef] [PubMed]
  4. Godard, N.; Grysan, P.; Defay, E.; Glinšek, S. Growth of 100-oriented lead zirconate titanate thin films mediated by a safe solvent. J. Mater. Chem. C 2021, 9, 281–287. [Google Scholar] [CrossRef]
  5. Wang, Y.; Cheng, H.; Yan, J.; Chen, N.; Yan, P.; Yang, F.; Ouyang, J. Large piezoelectricity on Si from highly (001)-oriented PZT thick films via a CMOS-compatible sputtering/RTP process. Materialia 2019, 5, 100228. [Google Scholar] [CrossRef]
  6. Liu, W.; Xu, W.; Wang, W.; He, L.; Zhou, J.; Radhakrishnan, K.; Yu, H.; Ren, J. RF reactive sputtering AlN thin film at room temperature for CMOS-compatible MEMS application. In Proceedings of the 2017 Joint IEEE International Symposium on the Applications of Ferroelectric (ISAF)/International Workshop on Acoustic Transduction Materials and Devices (IWATMD)/Piezoresponse Force Microscopy (PFM), Atlanta, GA, USA, 7–11 May 2017; pp. 52–55. [Google Scholar] [CrossRef]
  7. Chen, C.; Shang, Z.; Gong, J.; Zhang, F.; Zhou, H.; Tang, B.; Xu, Y.; Zhang, C.; Yang, Y.; Mu, X. Electric Field Stiffening Effect in c-Oriented Aluminum Nitride Piezoelectric Thin Films. ACS Appl. Mater. Interfaces 2018, 10, 1819–1827. [Google Scholar] [CrossRef]
  8. Wang, W.; Weinstein, D. Acoustic Bragg reflectors for Q-enhancement of unreleased MEMS resonators. In Proceedings of the 2011 Joint Conference of the IEEE International Frequency Control and the European Frequency and Time Forum (FCS) Proceedings, San Francisco, CA, USA, 2–5 May 2011; pp. 1–6. [Google Scholar] [CrossRef]
  9. Saddik, G.N.; Boesch, D.S.; Stemmer, S.; York, R.A. Strontium titanate DC electric field switchable and tunable bulk acoustic wave solidly mounted resonator. In Proceedings of the 2008 IEEE MTT-S International Microwave Symposium Digest, Atlanta, GA, USA, 15–20 June 2008; pp. 1263–1266. [Google Scholar] [CrossRef]
  10. Chen, Q.; Zhang, T.; Wang, Q.M. Frequency-temperature compensation of piezoelectric resonators by electric DC bias field. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2005, 52, 1627–1631. [Google Scholar] [CrossRef]
  11. Piazza, G.; Stephanou, P.J.; Pisano, A.P. Piezoelectric Aluminum Nitride Vibrating Contour-Mode MEMS Resonators. J. Microelectromechanical Syst. 2006, 15, 1406–1418. [Google Scholar] [CrossRef]
  12. Lin, B.T.; Lee, W.H.; Shieh, J.; Chen, M.J. Ferroelectric AlN ultrathin films prepared by atomic layer epitaxy. In Proceedings of the Behavior and Mechanics of Multifunctional Materials XIII, Denver, CO, USA, 3–7 March 2019; Naguib, H.E., Ed.; International Society for Optics and Photonics, SPIE: Denver, Colorado, 2019; Volume 10968, pp. 287–293. [Google Scholar] [CrossRef]
  13. Liu, H.; Zhong, J.; Lee, C.; Lee, S.W.; Lin, L. A comprehensive review on piezoelectric energy harvesting technology: Materials, mechanisms, and applications. Appl. Phys. Rev. 2018, 5, 041306. [Google Scholar] [CrossRef]
  14. Wang, L.P.; Wolf, R.; Wang, Y.; Deng, K.; Zou, L.; Davis, R.; Trolier-McKinstry, S. Design, fabrication, and measurement of high-sensitivity piezoelectric microelectromechanical systems accelerometers. J. Microelectromechanical Syst. 2003, 12, 433–439. [Google Scholar] [CrossRef]
  15. Wu, Y.; Ma, Y.; Zheng, H.; Ramakrishna, S. Piezoelectric materials for flexible and wearable electronics: A review. Mater. Des. 2021, 211, 110164. [Google Scholar] [CrossRef]
  16. Liu, J.; Liu, Y.; Zhao, L.; Xu, D.; Chen, W.; Deng, J. Design and Experiments of a Single-Foot Linear Piezoelectric Actuator Operated in a Stepping Mode. IEEE Trans. Ind. Electron. 2018, 65, 8063–8071. [Google Scholar] [CrossRef]
  17. Wen, W.; Shitang, H.; Shunzhou, L.; Minghua, L.; Yong, P. Enhanced sensitivity of SAW gas sensor coated molecularly imprinted polymer incorporating high frequency stability oscillator. Sens. Actuators B Chem. 2007, 125, 422–427. [Google Scholar] [CrossRef]
  18. Zhu, Y.; Lee, J.E.Y.; Seshia, A.A. A Resonant Micromachined Electrostatic Charge Sensor. IEEE Sens. J. 2008, 8, 1499–1505. [Google Scholar] [CrossRef]
  19. Tichy, J.; Erhart, J.; Kittinger, E.; Prívratskȥ, J. Fundamentals of Piezoelectric Sensorics; Springer: Berlin/Heidelberg, Germany, 2010; pp. 55–68. [Google Scholar]
  20. Ben Hassine, N.; Mercier, D.; Renaux, P.; Chappaz, C.; Basrour, S.; Defay, E. Linear variation of aluminum nitride capacitance versus voltage induced by a piezoelectric-electrostrictive coupling. J. Appl. Phys. 2008, 104, 034110. [Google Scholar] [CrossRef]
  21. Hall, D.A. Review Nonlinearity in piezoelectric ceramics. J. Mater. Sci. 2001, 36, 4575–4601. [Google Scholar] [CrossRef]
  22. Asanuma, H.; Komatsuzaki, T. Nonlinear piezoelectricity and damping in partially-covered piezoelectric cantilever with self-sensing synchronized switch damping on inductor circuit. Mech. Syst. Signal Process. 2020, 144, 106867. [Google Scholar] [CrossRef]
  23. Stanton, S.C.; McGehee, C.C.; Mann, B.P. Nonlinear dynamics for broadband energy harvesting: Investigation of a bistable piezoelectric inertial generator. Phys. D Nonlinear Phenom. 2010, 239, 640–653. [Google Scholar] [CrossRef]
  24. Jamshidi, R.; Jafari, A.A. Nonlinear vibration of conical shell with a piezoelectric sensor patch and a piezoelectric actuator patch. J. Vib. Control. 2022, 28, 1502–1519. [Google Scholar] [CrossRef]
  25. Xu, L.; Mu, X.; Chen, X.G.; Zhang, H.Y.; Xiong, R.G. Organic Enantiomeric Ferroelectrics with High Piezoelectric Performance: Imidazolium l- and d-Camphorsulfonate. Chem. Mater. 2021, 33, 5769–5779. [Google Scholar] [CrossRef]
  26. Tatarenko, A.S.; Murthy, D.V.B.; Srinivasan, G. Hexagonal ferrite-piezoelectric composites for dual magnetic and electric field tunable 8–25 GHz microstripline resonators and phase shifters. Microw. Opt. Technol. Lett. 2012, 54, 1215–1218. [Google Scholar] [CrossRef]
  27. Serrano, D.E.; Tabrizian, R.; Ayazi, F. Electrostatically tunable piezoelectric-on- silicon micromechanical resonator for real-time clock. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2012, 59, 358–365. [Google Scholar] [CrossRef] [PubMed]
  28. van Hemert, T.; Reimann, K.; Hueting, R.J.E. Extraction of second order piezoelectric parameters in bulk acoustic wave resonators. Appl. Phys. Lett. 2012, 100, 232901. [Google Scholar] [CrossRef]
  29. Wang, Q.M.; Zhang, Q.; Xu, B.; Liu, R.; Cross, L.E. Nonlinear piezoelectric behavior of ceramic bending mode actuators under strong electric fields. J. Appl. Phys. 1999, 86, 3352–3360. [Google Scholar] [CrossRef]
  30. Wang, Q.M.; Zhang, T.; Chen, Q.; Du, X.H. Effect of DC bias field on the complex materials coefficients of piezoelectric resonators. Sens. Actuators A Phys. 2003, 109, 149–155. [Google Scholar] [CrossRef]
  31. Kusters, J. The Effect of Static Electric Fields on the Elastic Constants of α-Quartz. In Proceedings of the 24th Annual Symposium on Frequency Control, Atlantic City, NJ, USA, 27–29 April 1970; pp. 46–54. [Google Scholar] [CrossRef]
  32. Newnham, R.E. Properties of Materials: Anisotropy, Symmetry, Structure; Oxford University Press: Oxford, UK, 2005; pp. 147–161. [Google Scholar]
  33. Defaÿ, E.; Hassine, N.B.; Emery, P.; Parat, G.; Abergel, J.; Devos, A. Tunability of Alluminum Nitride Acoustic Resonators: A Phenomenological Approach. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2011, 58, 2516–2520. [Google Scholar] [CrossRef]
  34. Hruska, K. Non-linear equations of state of second-order electromechanical effects. Czechoslov. J. Phys. 1964, 14, 309–321. [Google Scholar] [CrossRef]
  35. Chizhikov, S.I.; Sorokin, N.G.; Petrakov, V.S. The elastoelectric effect in the non-centrosymmetric crystals. Ferroelectrics 1982, 41, 9–25. [Google Scholar] [CrossRef]
  36. Landis, C.M. Fully coupled, multi-axial, symmetric constitutive laws for polycrystalline ferroelectric ceramics. J. Mech. Phys. Solids 2002, 50, 127–152. [Google Scholar] [CrossRef]
  37. Delibas, B.; Arockiarajan, A.; Seemann, W. Rate dependent properties of perovskite type tetragonal piezoelectric materials using micromechanical model. Int. J. Solids Struct. 2006, 43, 697–712. [Google Scholar] [CrossRef]
  38. Auld, B.A. Acoustic Fields And Waves In Solids; Wiley-Interscience: Toronto, Canada, 1973; Volume 1, pp. 386–408. [Google Scholar]
  39. Hahn, T.; Klapper, H.; Muller, U.; Aroyo, M.I. Point Groups and Crystal Classes; Internation Union of Crystallography: New York, NY, USA, 2016; Volume A, pp. 720–776. [Google Scholar]
  40. García, L.; Scherer, C. Introduction to Classical Thermodynamics; Trillas: Ciudad de Mexíco, Mexico, 1972; Chapter 9. [Google Scholar]
  41. Clayton, J.D. Nonlinear Mechanics of Crystals; Springer: London, UK, 2011; pp. 543–566. [Google Scholar]
  42. Manjunatha, H.; Naidu, K.C.B.; Kumar, N.S.; Pothu, R.; Boddula, R. Applications of Metal–Organic Frameworks (MOFs) and Their Derivatives in Piezo/Ferroelectrics. In Applications of Metal–Organic Frameworks and Their Derived Materials; John Wiley & Sons, Ltd: Hoboken, NJ, USA, 2020; Chapter 2; pp. 41–51. [Google Scholar]
  43. Kittinger, E.; Tichy, J. Electroelastic effect of crystal rods expressed by fundamental material constants. J. Acoust. Soc. Am. 1987, 83, 647–651. [Google Scholar] [CrossRef]
  44. Hruska, C.K. Determination of the thirdorder piezoelectric constants of quartz using the extentional mode of rods. J. Appl. Phys. 1989, 66, 1071–1074. [Google Scholar] [CrossRef]
  45. Tunable Silicon Bulk Acoustic Resonators with Multi-Face AlN transduction; International Frequency Control and the European Frequency and Time Forum: San Francisco, CA, USA, 2011.
  46. Feldmann, M.; Hénaff, J. Surface Acoustic Waves for Signal Processing; Artech House: Norwood, OH, USA, 1989; Volume 1, pp. 29–41. [Google Scholar]
  47. Bertotti, G.; Mayergoyz, I.D. The Science of Hysteresis; Elsevier: Amsterdam, The Netherlands, 2005; Volume 3, Chapter 4; pp. 337–451. [Google Scholar]
  48. Smith, R.C. A Domain Wall Model for Hysteresis in Piezoelectric Materials. J. Intell. Mater. Syst. Struct. 2000, 11, 62–79. [Google Scholar] [CrossRef]
  49. The Effect Of Static Electric Fields On The Elastic Constants of alpha-Quartz; SPIE Smart Structures + Nondestructive Evaluation: Denver, CO, USA, 2019.
  50. Ang, W.T.; Garmón, F.A.; Khosla, P.K.; Riviere, C.N. Modeling Rate-Dependent Hysteresis in Piezoelectric Actuators. In Proceedings of the 2003 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS 2003) (Cat. No. 03CH37453), Las Vegas, NV, USA, 27–31 October 2003. [Google Scholar]
  51. Abdolali, S.; Jahromi, Z. Nonlinear Constitutive Modeling of Piezoelectric Materials. Master’s Thesis, University of Calagary, Calgary, AB, Canada, 2013. [Google Scholar]
  52. Fulton, C.C.; Gao, H. Electrical Nonlinearity in Fracture of Piezoelectric Ceramics. Appl. Mech. Rev. 1997, 50, S56–S63. [Google Scholar] [CrossRef]
  53. Eitel, R.E.; Shrout, T.R.; Randall, C.A. Nonlinear contributions to the dielectric permittivity and converse piezoelectric coefficient in piezoelectric ceramics. J. Appl. Phys. 2006, 99, 124110. [Google Scholar] [CrossRef]
Figure 1. (a) Transversal cut of the simulated device that was fabricated in [28]; the scales for the axis are different to expose all the layers. (b) The impedance of the devices simulated for an external DC electric field in the range of −2 MV/m to 2 MV/m.
Figure 1. (a) Transversal cut of the simulated device that was fabricated in [28]; the scales for the axis are different to expose all the layers. (b) The impedance of the devices simulated for an external DC electric field in the range of −2 MV/m to 2 MV/m.
Materials 16 03432 g001
Figure 2. (a) Effective elasticity constants C 33 D e f f and (b) resonance frequencies, obtained from the simulations and measurements of the device fabricated in [28], for an external DC electric field in the range of −2 MV/m to 2 MV/m.
Figure 2. (a) Effective elasticity constants C 33 D e f f and (b) resonance frequencies, obtained from the simulations and measurements of the device fabricated in [28], for an external DC electric field in the range of −2 MV/m to 2 MV/m.
Materials 16 03432 g002
Figure 3. (a) Relative effective permittivity ϵ 33 r e f f obtained from the simulations and measurements of the device fabricated in [28] for an external DC electric field in the range of −2 MV/m to 2 MV/m. (b) Electromechanical coupling factor for the device simulated; this parameter was not reported by [28].
Figure 3. (a) Relative effective permittivity ϵ 33 r e f f obtained from the simulations and measurements of the device fabricated in [28] for an external DC electric field in the range of −2 MV/m to 2 MV/m. (b) Electromechanical coupling factor for the device simulated; this parameter was not reported by [28].
Materials 16 03432 g003
Table 1. Estimated order of magnitude of the high-order tensors for the nonlinear effects of A l N and P Z T 5 H (4 mm and 6 mm point groups respectively), using the stress–charge formulation presented.
Table 1. Estimated order of magnitude of the high-order tensors for the nonlinear effects of A l N and P Z T 5 H (4 mm and 6 mm point groups respectively), using the stress–charge formulation presented.
Material ParameterSymbolDefinitionOrderUnits
Elasticity Constant C λ μ T λ S μ 10 11 N/m 2
Relative Electrical Permittivity ϵ i j r D i E j 10 1 1
Electrical correction term of elasticity constants
Strain correction term of piezoelectric coefficient
g λ μ k 2 T λ E k S μ = 2 T λ S μ E k
2 D k S λ S μ
10 1 10 3 C/m 2
Electrical correction term of piezoelectric coefficient
Strain correction term of electrical permittivity
q i j λ 2 D k S λ E j = 2 D k E j S λ
2 T λ E j E k
10 10 10 8 N/Vm
Electrical correction term of permittivity r i j k 2 D i E j E k 10 20 10 22 C/V 2
Strain correction term of elasticity constants t λ μ ν 2 T λ S μ S m u 10 10 10 12 N/m 2
Table 2. Numeration of point groups of crystal systems and Laue groups. The identification number corresponds to that used in Table 3, Table 4 and Table A1 to show the symmetry structure of the high-order tensors for each type of material.
Table 2. Numeration of point groups of crystal systems and Laue groups. The identification number corresponds to that used in Table 3, Table 4 and Table A1 to show the symmetry structure of the high-order tensors for each type of material.
Crystal SystemLaue GroupSymbolId.SymbolId.SymbolId.
TriclinicN11 1 ¯ 2
MonoclinicM23m42/m5
OrthorhombicO2226mm27mmm8
TetragonalTII49 4 ¯ 104/m11
TetragonalTI4 mm13 4 ¯ 2 m144/mmm15
RhombohedralRI32163 m17 3 ¯ m18
RhombohedralRII319 3 ¯ 20422 1 12
HexagonalHII621 6 ¯ 226/m23
HexagonalHI622246 mm25 6 ¯ m26
CubicCII2328m 3 ¯ 296/mmm 2 27
CubicCI432304 3 ¯ m31m 3 ¯ m32
1 Belongs to the TI Laue group. 2 Belongs to the HI Laue group.
Table 3. Symmetry structure of high-order tensor q j k λ for the nonlinear piezoelectric coupling effect.
Table 3. Symmetry structure of high-order tensor q j k λ for the nonlinear piezoelectric coupling effect.
Laue Group
Comp.NMOTIITIRIRIIHIIHICICII
111111111111111111111111111111111111
1211210012100121121000
13113113100001310000
221221221221221221221221221221221221
23123100002312310000
331331331331331331331331331331331331
112112112112221221221221221221331331
12212200−12100−121−121000
1321321320000−1310000
222222222222111111111111111111111111
2322320000−231−2310000
332332332332331331331331331331221221
113113113113113113113113113113221221
1231230000000000
133133133000000000
223223223223113113113113113113331331
2332330000000000
333333333333333333333333333333111111
11411400001141140000
12412412400001240000
1341340013400134134000
2242240000−114−1140000
234234234234234234234234234234234234
3343340000000000
1151151150000−1240000
12512500001141140000
135135135135234234234234234234234234
22522522500001240000
23523500−13400−134−134000
335335335000000000
1161160011600−121−121000
126126126126126126AAAA234234
13613600002312310000
22622600−11600121121000
2362362360000−1310000
3363360000000000
A = (111−221)/2.
Table 4. Symmetry structure of the fifth-ranked tensor g λ μ k for the nonlinear piezoelectric coupling effect. The point groups not shown are zero.
Table 4. Symmetry structure of the fifth-ranked tensor g λ μ k for the nonlinear piezoelectric coupling effect. The point groups not shown are zero.
Point Group of Symmetry
Comp.13467910121314161719212224252628,3130
11111101110000000111011101110011100
12112101210000000121012101210012100
13113101310000000131013101310013100
14114114101410141141141014114101411410141001410
1511510151015115115101510015115115100151000
161161161000000000161161016100000
22122102210000000A0J0A00A00
23123102310000000−1310−1310−13100−13100
2412412410241024124124102412410241241024100241241
2512510251025125125102510025125125100251000
261261261000000000261261026100000
331331033100000000000000000
3413413410341034134134103413410341341034100341−241
3513510351035135135103510034235135100351000
361361361000000000361361036100000
44144104410000000441044104410044100
451451451000000000451451045100000
46146104610461461461046100EEE00E000
55155105510000000−4410−4410−44100−44100
561561561056105615615610561B0BB0B005610
66166106610000000121012101210012100
112112112000000000FF0F00000
122122122000000000GG0G00000
132132132000000000361361036100000
14214201420142251−25102510025125125100251000
15215215201520−241241−2410241−2410−241−2410−24100341−241
16216201620000000C0C0C00C00
222222222000000000HH0H00000
232232232000000000−361−3610−36100000
24224202420242151−15101510015115115100151000
25225225202520−141141−1410141−1410-141−1410−141001410
26226202620000000D0-D0-D00-D00
332332332000000000000000000
34234203420342351−35103510034235135100351000
35235235203520−341341−3410341−3410-341−3410−34100241241
36236203620000000−1310−1310−13100−13100
442442442000000000−451−4510−45100000
45245204520000000441044104410044100
46246246204620−561561−5610561−B0−B−B0−B005610
552552552000000000451451045100000
56256205620562461−461046100EEE00E000
662662662000000000GG0G00000
1131130113011311311301130011311311300113000
12312301230123123001230012312312300123000
1331330133013313313301330013313313300133000
143143143000000000143143014300000
15315301530000000153015301530015300
16316316301630163163163016300000000241241
22322302230223113−11301130011311311300113000
23323302330233133−13301330013313313300133000
243243243000000000−143−1430−14300000
25325302530000000−1530−1530−15300−15300
26326326302630−163163−163016300000000341−241
33333303330333333003330033333333300333000
343343343000000000000000000
353353035300000000000000000
36336336303630036300363000000001410
4434430443044344344304430044344344300443000
45345345304530045300453000000005610
46346304630000000−1530−1530−15300−15300
55355305530553443−44304430044344344300443000
563563563000000000143143014300000
663663066306636630066300III00I000
A = −111−121/2, B = −(141−241)/2, C = −111/2−121(3/2), D = (121−111)/2, E = (151−251)/2, F = 161/2 + 261(3/2), G = (161−261)/2, H = −261/2 161(3/2), I = (113−123)/2, J = −111 + 121/2.
Table 5. Percent errors obtained from the simulations with respect to the measurements for the effective elasticity constant, relative effective permittivity, and resonance frequency.
Table 5. Percent errors obtained from the simulations with respect to the measurements for the effective elasticity constant, relative effective permittivity, and resonance frequency.
Percent Error
QuantitySymbolAverageMaximum
Effective Elasticity Constant C 33 D e f f 0.280.79
Relative Effective Electrical Permittivity ϵ 33 r e f f 0.922.9
Resonance Frequency f r 0.150.3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jaramillo-Alvarado, A.F.; Torres Jacome, A.; Rosales-Quintero, P.; Vazquez-Leal, H.; Diaz-Arango, G.; Huerta-Chua, J.; Martínez-Castillo, J. Stress–Charge Nonlinear Physical Description and Tensor Symmetries for Piezoelectric Materials. Materials 2023, 16, 3432. https://doi.org/10.3390/ma16093432

AMA Style

Jaramillo-Alvarado AF, Torres Jacome A, Rosales-Quintero P, Vazquez-Leal H, Diaz-Arango G, Huerta-Chua J, Martínez-Castillo J. Stress–Charge Nonlinear Physical Description and Tensor Symmetries for Piezoelectric Materials. Materials. 2023; 16(9):3432. https://doi.org/10.3390/ma16093432

Chicago/Turabian Style

Jaramillo-Alvarado, A. F., A. Torres Jacome, P. Rosales-Quintero, H. Vazquez-Leal, G. Diaz-Arango, J. Huerta-Chua, and J. Martínez-Castillo. 2023. "Stress–Charge Nonlinear Physical Description and Tensor Symmetries for Piezoelectric Materials" Materials 16, no. 9: 3432. https://doi.org/10.3390/ma16093432

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop