We first revisit the mathematical modeling of the flexoelectric effect in the context of continuum mechanics at infinitesimal deformations. We establish and clarify the relation between the different formulations, point out theoretical and numerical issues related to the resulting boundary value problems, and present the natural extension to finite deformations. We then present a simple B-spline based computational technique to numerically solve the associated boundary value problems, which can be extended to handle unfitted meshes, hence allowing for arbitrarily-shaped geometries. Several numerical examples illustrate the flexoelectric effect in simple benchmark setups, as well as in new flexoelectric devices and metamaterials engineered for sensing or actuation.
I. INTRODUCTION
Flexoelectricity generally refers to the two-way coupling between strain gradient and electric polarization (direct and inverse effects). Converse flexoelectricity couples polarization gradients with strain. All these effects have been demonstrated experimentally in cantilever thin films or nanobeams and truncated pyramids, under inhomogeneous mechanical fields (direct effect)1,2 and under homogeneous applied electric field (inverse effect),3–6 and in truncated pyramids and in PFM,7 under inhomogeneous applied electric field8 (converse effect). The setups mobilizing flexoelectricity, which necessarily involve non-uniform fields, demand accurate theoretical models for the quantitative evaluation of the effect and, in particular, for the correct interpretation of characterization experiments.9
Phenomenological models for flexoelectricity in crystalline dielectrics were first proposed by Kogan10 after the early studies by Mashkevich and Tolpygo11 and Tolpygo.12 First comprehensive theoretical works by Tagantsev13,14 clarified the distinction between piezoelectricity and flexoelectricity. In the mechanics community, Mindlin15 formalized the converse flexoelectric effect in elastic dielectrics. A complete unified continuum framework, including strain gradient elasticity, both direct and converse flexoelectric couplings, and the polarization inertia effect was proposed later by Sahin and Dost.16 More recently, a simplified framework for isotropic dielectrics was proposed by Maranganti et al.17
Nowadays, various continuum theories of flexoelectricity co-exist in the literature. Some of them are based on extensions to the electromechanical setting of gradient elasticity and other enriched continuum theories.18–24 Other authors consider the couplings with further physics, such as the flexoelectric effect in ferroelectrics,25,26 the coupling with magnetic fields27,28 or photovoltaics,29,30 and the contributions of surface effects.31 General variational principles for flexoelectric materials can be found in Liu,28 Shen and Hu,31 Hu and Shen.32 The reader is referred to Yudin and Tagantsev,33 Nguyen et al.,34 Zubko et al.,35 Krichen and Sharma,36 Wang et al.37 for comprehensive reviews of flexoelectricity in solids. Another focus of recent research is the modeling of flexoelectricity for soft materials (e.g., polymers and elastomers), which requires a finite deformation framework.22,28,38–41
All the aforementioned theories of flexoelectricity can be classified depending on the following considerations:
The choice of state variables describing the flexoelectric effect. For the mechanics, either the displacement gradient or its symmetrized form (i.e., strain) can be used, which give rise to type-I or type-II flexoelectricity, respectively. For the dielectrics, the most natural variable is the electric polarization. However, there exist many theories taking the electric field or the electric displacement instead.
The flexoelectric coupling considered explicitly, either the direct one, the converse, both, or the Lifshitz-invariant form. Section II A further elaborates on this topic.
The dielectric media surrounding the flexoelectric solid being included in the modeling or not. Since Maxwell equations hold in the whole , the electric potential must be solved as well outside the flexoelectric solid, in principle. However, if the dielectric permittivity of the flexoelectric material is large enough as compared to that of the surrounding media, zero-electric displacement around the flexoelectric solid is a reasonable assumption.
We present here an overview of various theoretical continuum formulations for flexoelectricity and the resulting computational models based on the discretization of the corresponding governing equations with smooth approximants. The paper is structured as follows. Section II presents the continuum modeling of type-II flexoelectricity models neglecting the surrounding media. It includes the discussion of flexoelectric couplings at infinitesimal deformations proposed in the literature, establishing the relationships between them, the different variational principles governing the behavior of flexoelectric solids, their associated boundary value problems and corresponding boundary conditions. The extension of the continuum modeling of flexoelectricity to finite deformations is stated and discussed. Section III presents a computational approach based on smooth B-spline basis functions, which can be combined with the immersed boundary method to allow for arbitrarily-shaped geometries. -adaptivity is easily achieved by means of hierarchical refinement. The non-dimensionalization of the resulting equations is briefly commented. In Sec. IV, we use the computational approach discussed in Sec. III to solve the BVPs describing the electromechanical response of a flexoelectric cantilever beam in bending, under both mechanical and electrical input, based on the different formulations presented in Sec. II. A detailed comparison of the electromechanical output is given and several numerical issues are pointed out. In Sec. V, we explore the cantilever beam actuator in the regime of finite deformations, and study the interplay between flexoelectricity and electrostriction for different actuation levels. Then, in Sec. VI, we propose a new flexoelectric device and flexoelectricity-based piezoelectric metamaterials engineered to harness the flexoelectric effect toward sensing/actuation applications. Section VII concludes this paper.
II. CONTINUUM MODELING
In this section, we review the continuum modeling of the flexoelectric effect considering type-II flexoelectricity models at infinitesimal deformations and neglecting the surrounding media. In Sec. II A, we discuss the direct and converse flexoelectric couplings and their relation via null Lagrangians. Section II B presents the variational principles for flexoelectric solids, taking either the electric polarization or the electric potential as primal state variables, and explores their relation via a partial Legendre transform. Section II C formulates the boundary value problems of potential-based variational principles using the direct and the Lifshitz-invariant flexoelectricity models and introduces the weak enforcement of essential boundary conditions via Nitsche’s method. Non-standard boundary conditions, such as sensing electrode conditions, interface conditions and generalized periodicity conditions are briefly discussed as well. Finally, the modeling of flexoelectricity at finite deformations is introduced in Sec. II D.
A. Direct and converse flexoelectricity
is the usual fourth-order elasticity tensor,
is the usual second-order reciprocal dielectric susceptibility tensor,
is the sixth-order strain gradient elasticity tensor, representing the purely non-local elastic effects,
is the fourth-order polarization gradient tensor, representing the purely non-local effects of polarization,
is the direct flexocoupling tensor,
is the polarization gradient-strain coupling tensor introduced by Mindlin in his theory of polarization gradient.15 In a more modern context, it is also known as the converse flexocoupling tensor.
In non-centrosymmetric dielectrics, other relevant physics arise, such as piezoelectricity, the linear two-way coupling between strains and polarization. This coupling can be incorporated to Eq. (1) as an additional contribution to the internal energy density, , with the third-order piezoelectric tensor . The interplay between flexoelectricity and piezoelectricity is very rich and worth studying (see, for instance, Abdollahi and Arias5 and Bhaskar et al.42). Here, we focus on the formulation of flexoelectricity in centrosymmetric dielectrics for the sake of simplicity. More complete phenomenological formulations for flexoelectricity have also been proposed and studied in the literature,43,44 which consider as well the linear couplings between strain and strain gradients, polarization and polarization gradients, and polarization gradients and strain gradients.
The direct and converse flexoelectric effects are present in all the aforementioned formulations, as revealed in the constitutive equations for mechanics and dielectrics in Eq. (8). On the one hand, the direct flexoelectric effect in Eq. (8a) induces an electric field (or polarization) proportional to the strain gradient, whereas the converse effect in Eq. (8b) consists on a contribution to the mechanical stress proportional to the gradient of polarization. Note that both flexoelectric effects are governed by the same flexocoupling tensor .
B. Variational principles
The physics governing the behavior of dielectric materials can be modeled at a continuum level by different variational principles, depending on the state variables that are chosen to describe them. The usual choice for the mechanical state variable is the displacement field , whose symmetric gradient yields the strain field, . However, different options are frequently considered for the electrical state variable: either the electric polarization , the electric displacement , or the electric field . Sections focus on two of them, and , which give rise to two different variational principles.
1. Minimization of the free energy Π[u, P]
2. Optimization of the free enthalpy
3. Equivalence between different variational principles via partial Legendre transform
The variational models based on free energy minimization and the ones based on free enthalpy optimization are related by means of a partial Legendre transform. That is, given one form, the other one is uniquely determined, and it can be directly obtained—under some assumptions that are next discussed—by a partial Legendre transform, thus revealing the relations between the material parameters of each form.
C. Boundary value problems
In this section, we consider the – variational principle on in Sec. II B 2 and derive the corresponding boundary value problems for the Direct and Lifshitz free enthalpy densities in Eqs. (19) and (20).
Let be a physical domain in . The domain boundary, , can be conformed by several smooth portions as [see Figs. 1(a) and 2(a)]. At each point , we define as the outward unit normal vector.
The boundary of the th portion of is a closed curve ( ) or a pair of points ( ) denoted as . The union represents the regions at which is not smooth, and , , represents the part of that is adjacent to and . In case , at each point , we define as the unit co-normal vector pointing outward of , which is orthogonal to the normal vector and to the tangent vector of the curve [see Figs. 1(b) and 1(c)]. In case , is defined on as the outward unit tangent vector to [see Figs. 2(b) and 2(c)].
The jump operator acting on a given quantity is defined on as the sum of that quantity evaluated at both sides of , namely, , where is the value of from [see Fig. 1(a)].
1. Direct flexoelectricity
In practice, in a finite element (FE) context, the functional spaces in Eq. (46) and (48) are approximated by means of a set of linear combinations of basis and test functions. However, and are in general difficult to approximate since they require fulfilling second order Dirichlet conditions for the displacement field (i.e., prescribing its normal derivative). The typical approach to overcome this difficulty is considering , which implies that only second order Neumann boundary conditions are allowed, i.e., .58,64–67 This choice is further justified by the unclear physical interpretation of second order Dirichlet boundary conditions.40
2. Lifshitz-invariant flexoelectricity
3. Weak enforcement of Dirichlet boundary conditions via Nitsche’s method
So far, the functional spaces of the state variables (48) or (61) are chosen such that Dirichlet boundary conditions are automatically fulfilled. In computational flexoelectricity, this implies having an interpolant space on the boundaries where Dirichlet conditions are to be enforced, e.g., by means of body-fitted meshes in a FE approach. However, it is typically difficult to achieve such functional spaces while fulfilling the continuity requirement.59
An alternative to enforce essential boundary conditions without constraining the functional spaces consists on incorporating them into the enthalpy functional, in a way that equilibrium states satisfying the corresponding variational principle necessarily fulfill (in a weak sense) the Dirichlet boundary conditions. We show here how to do so via Nitsche’s method68 due to its simple form and convenient numerical properties (i.e., self-consistency, symmetry, optimal error convergence rates, and preservation of the number of degrees of freedom69) as compared to other alternatives such as the Lagrange multipliers or penalty methods.
Many authors in computational flexoelectricity neglect the edge Dirichlet conditions (43a) on 9,57,58,71 [or analogously (59a) on ]. In the cases where essential boundary conditions are enforced strongly, as in conforming FE or meshless discretizations, this fact has no practical relevance since the strong imposition on the surface ( ) automatically implies the strong imposition on the adjacent edges in ( ) as well. However, it is important to underline that in frameworks where boundary conditions are enforced weakly, dismissing edge conditions is equivalent to considering homogeneous Neumann edge conditions, which is wrong on Dirichlet edges, and can completely spoil the quality of the numerical results.63
4. Beyond standard boundary conditions
So far, we have mentioned that the standard conditions along the boundary of the flexoelectric domain, arising naturally from the derivation of the boundary value problems, are either Dirichlet or Neumann. However, it is worth mentioning other types of boundary conditions that are also frequent in electromechanical boundary value problems:
a. Sensing electrode condition
The electric potential along electrodes placed on the surface of the devices is uniform, since they are made of conducting material. Electrodes in actuation mode correspond to a standard Dirichlet boundary condition for [see Eq. (43c) or (59c)], where the prescribed voltage is constant.
In a discrete boundary value problem with body-fitted discretization, Eq. (72) can be enforced strongly by assigning the same degree of freedom for to all the nodes on the electrode, and Eq. (73) is then automatically fulfilled. Otherwise, they must be enforced weakly, e.g., via Lagrange multipliers72 or a modified version of Nitsche’s method.63
b. Interface condition
Composite materials, i.e., domains composed by multiple materials with different properties, can be used to achieve a specific functionality of an electromechanical device. The paradigmatic example consists on cantilevers bimorphs for piezoelectric transduction,5,73,74 but one can also think of flexoelectricity-based devices with multimaterial stacks46,75,76 or structured materials with geometrically polarized cavities2,77–79 and even conducting/insulating inclusions.80,81
In the context of discrete flexoelectricity boundary value problems, high-order interface conditions must be considered.81 Such conditions correspond to high-order continuity of the state variables and high-order equilibrium, (i) across the interface between two adjacents subdomains, as well as (ii) at the sharp regions of the interfaces, shared by multiple—possibly more than two—subdomains (see Fig. 3).
c. Generalized periodicity condition
Architected materials exhibit an intrinsic microstructure that is replicated periodically in space. In order to numerically compute their overall bulk response, free of finite edge effects, undergoing macroscopic displacement (sensor) or electric field (actuator) conditions, an efficient strategy consists on exploiting the lattice periodicity by studying a single representative volume element (RVE) or unit cell under (high-order) generalized-periodic conditions.82,83 These set of conditions are a generalization of classical (high-order) periodicity conditions to account for non-periodic solution fields with periodic gradients.
Consider an infinite lattice with translational periodicity in . Without loss of generality, the flexoelectricity boundary value problem can be reduced to a single RVE defined as with the proper generalized periodic conditions (i) between a given periodic boundary of the RVE (i.e., , , or ) and its corresponding translational image (i.e., , , or ), as well as (ii) between the intersections of and with the actual boundary of the infinite lattice, denoted as and , respectively (see Fig. 4).
In the context of flexoelectricity, generalized periodicity conditions allow the systematic study and design of flexoelectric metamaterials exhibiting an apparent piezoelectric behavior.83 An elegant way of enforcing high-order generalized-periodic conditions in a discrete boundary value problem consists on simply constructing a high-order generalized-periodic approximation space.82 Alternatively, one can resort to enforcing generalized periodicity conditions (76) weakly.81
D. Flexoelectricity in soft materials: Beyond infinitesimal deformations
The interest on computational flexoelectricity in soft materials has increased in recent years22,38–41,64,84 due to different reasons. On the one hand, a large flexoelectric response is expected, since flexoelectric coefficients of polymers are at least the same order of magnitude as those of hard crystalline materials,85–87 but, in turn, they are much more compliant. On the other hand, electromechanical actuation of polymers by flexoelectricity overcomes the current limitations of traditional actuation (based on the Maxwell stress effect, i.e., an electromechanical coupling inducing a contribution in the stress field quadratic to the electric field): (i) one-way coupling, (ii) dielectric breakdown due to too large electric fields, and (iii) irreversibility of the deformation sign.36,88–90 We present here a brief introduction of computational flexoelectricity in the regime of finite deformations and review its most relevant features.
Consider a deformable dielectric body described by in the reference (or undeformed) configuration and by in the current (or deformed) configuration. The deformation map maps every material point to the spatial point . Relevant deformation measures in this context are the deformation gradient , the Jacobian determinant , and the right Cauchy–Green deformation tensor . In the reference configuration, the strain is represented by the Green–Lagrangian strain tensor , and the electric field in by .
In summary, the regime of finite deformations presents a richer electromechanical behavior as compared to the one in the limit of infinitesimal deformations. On the one hand, one can study the interplay between the electrostrictive and the flexoelectric effects in different scenarios, as illustrated in Sec. V with the actuation of a soft flexoelectric cantilever. On the other hand, the nonlinear definition of the strain field implies geometrical nonlinearity, allowing the occurrence of mechanical instabilities such as buckling, wrinkling or creasing, which can be harnessed toward the design of tunable flexoelectricity-based devices.41
III. NUMERICAL MODELING
The equations of flexoelectricity can only be solved analytically in very simple settings, such as simplified Euler–Bernoulli beam,51,52,91 Timoshenko beam,55 or Cosserat rod41 models. Otherwise, it is necessary to resort to computational flexoelectricity.
The major challenge is to handle the continuity of the state variables required by the fourth-order PDE system. To address this, several numerical alternatives have been proposed, i.e., mesh-free approximations,5,9,58,64,92 isogeometric analysis,39,41,65,84,93 the Argyris triangular element approximation,38 the immersed boundary B-spline method,63 or the C0-interior penalty method,72 among others. Another family of numerical methods are those circumventing the continuity requirement by introducing additional variables, such as mixed formulations,66,67,94 or those based on micromorphic theories of continua.22,40 We refer to Zhuang et al.59 for a comprehensive review on computational approaches to solve flexoelectricity boundary value problems.
In this tutorial, in Sec. III A, we focus first on open B-spline approximations in body-fitted meshes (i.e., a particular case within isogeometric analysis). It is a simple method to successfully resolve the flexoelectricity equations in a smooth approximation space, albeit only rectangular geometries (or cuboidal ones in 3D) are allowed. It is, therefore, a suitable method for the computational modeling of flexoelectricity in beams. In Sec. III B, in order to circumvent this limitation and allow arbitrarily-shaped geometries, we combine this approximation with the immersed boundary method,95,96 involving non-conforming discretizations of the geometry and non-interpolant discretizations of the state variables. Dirichlet boundary conditions must be weakly enforced, and some further numerical treatments in cut cells must be accounted for, such as defining a numerical integration for cut elements and alleviating ill-conditioning produced by degrees of freedom with small support within the domain. Section III C presents the extension to hierarchical B-spline basis, allowing -refinement of the computational mesh, and the adimensionalization of the resulting equations is shown in Sec. III D.
Using these approaches, as compared to the aforementioned ones, yields multiple benefits, i.e., (i) trivial mesh generation, (ii) easy handling of arbitrarily-shaped geometries, (iii) optimal convergence rates for the approximation error, that can be systematically increased using higher-order B-spline bases, (iv) cheap and robust evaluation of basis functions and derivatives, (v) relatively small number of degrees of freedom for a given spatial resolution, and (vi) good scalability and relatively low matrix fill-in due to a compact and structured set of basis functions.
A. Uniform open B-spline basis
B-spline functions97–99 are smooth, positive-valued piece-wise polynomials with compact support. Being the polynomial degree, they are by construction -continuous throughout the domain and form a partition of unity. Since they are polynomials in nature, they are explicitly evaluated and exactly integrated by rich enough numerical quadratures. However, they do not satisfy in general the Kronecker delta property for .
In Eq. (85) and throughout the rest of the paper, we follow the convention that the first basis function corresponds to index and spans from to . Note that other references might consider -based indexing of the basis functions or -based indexing for the knot vector instead.
Different kinds of B-spline bases are obtained depending on the choice of the knot vector. For simplicity, we consider equispaced knots , yielding a uniform B-spline basis where the th B-spline function can be expressed as a translation of the first one as (see Fig. 5).
In order to enforce boundary conditions in a strong way, B-spline bases must satisfy the Kronecker delta property at the boundary, corresponding to the knots and in parametric space. To do so, the basis is modified by knot multiplicity, i.e., , where and are repeated times. In doing so, the continuity of the B-spline basis at these knots is decreased to (i.e., discontinuous), yielding a boundary-interpolant (or open) basis suitable to enforce essential boundary conditions strongly (see Fig. 6).
Using the multivariate B-spline basis functions in Eq. (86) and the geometrical map in Eq. (89), the state variables in the boundary value problem are approximated by means of Eq. (88). Volume, surface, and line integrals are numerically computed by standard Gaussian quadrature rules, and boundary conditions are enforced strongly. In univariate open bases, the DOF corresponding to the first and last basis functions can be directly prescribed with the value of the boundary condition. In multivariate open bases, however, the values of the control variables on the boundary are computed by means of the projection of the boundary condition onto the space spanned by the corresponding B-spline basis functions.
The discretization and numerical approximation of the weak form yield an algebraic linear system of equations for and , and hence, the approximated fields and are obtained.
B. Immersed boundary B-spline approximation
Since the approximation space is unfitted to the geometry, it is not interpolant on , even if B-splines of degree or any other family of basis functions are considered. Therefore, in the context of B-spline basis functions, knot multiplicity does not provide any benefit, and uniform knot vectors are typically considered instead.
The physical boundary is allowed to intersect the cells of the embedding mesh arbitrarily, leading to a unfitted discretization of . The cells with a nonempty intersection with are labeled as active, whereas outer (or inactive) cells are neglected. Active cells can be either inner [ , see Fig. 7(b)] or cut [ , see Fig. 7(c)].
4. Cell classification
Labeling a cell depending on its intersection with is usually accomplished by checking whether all vertices of each cell (and possibly more points within the cell) lie within the domain (inner cell), only part of them (cut cell), or none of them (outer cell). In the case of implicit boundary representation (e.g., level set approaches), it is enough to evaluate the level set function on the evaluation points (see Fries and Omerović,100 Fries,101 Kudela et al.,102 Legrain et al.103). For explicit boundary representation (e.g., CAD descriptions), ray-tracing procedures are required, as explained in Marco et al.104,105
5. Enforcement of Dirichlet boundary conditions
Immersed boundary methods permit considering arbitrary geometries and involve trivial mesh generation, at the cost of having to deal with a non-conforming discretization. Since the approximation space is defined independently of , it is not interpolant on its boundary , hence, Dirichlet boundary conditions must be weakly enforced, e.g., by means of Nitsche’s method (cf. Sec. II C 3).
6. Numerical integration
An additional challenge, associated with the presence of cut cells, consists on defining a good-enough numerical integration on cut cells. This implies defining special quadrature rules for each cell since only their intersection with , i.e., , is to be considered. The standard approach consists on dividing into several non-overlapping sub-domains, each of them being easily integrated with predefined quadrature rules.
On the one hand, conforming subdivision approaches,101,102,104 most of them based on the marching cubes algorithm,106 attempt to subdivide cut cells into smaller conforming entities (e.g., squares or triangles in 2D, see Fig. 8), even allowing in some cases curved faces or edges.107–109 On the other hand, non-conforming subdivision schemes110,111 are based on quadtree/octree approaches, where each cell is recursively subdivided into pieces until some predefined recursion level (see Fig. 9). Inner pieces are completely integrated, whereas at those that remain cut in the finest subdivision level, only the Gauss points that lie within are considered.
Conforming approaches are typically more accurate but harder to implement, whereas non-conforming approaches are typically easier to implement and as accurate as required by increasing the recursivity.
Numerical integration on the faces and edges of cut cells requires an explicit boundary representation, available only if a CAD (or other explicit) description of is considered. In the case of implicit boundary representations such as level sets, an explicit approximation of must be precomputed at each cell.
7. Cut-cell stabilization
Another challenge related with cut cells consists on alleviating ill-conditioning produced by degrees of freedom with small intersection with . The resulting algebraic system of equations suffers from ill-conditioning, i.e., large condition number of the system matrix, in the presence of cut cells with a small portion in the domain, i.e., when the volume fraction . In particular, for the boundary value problems of flexoelectricity considered in this paper, the condition number scales with the minimum volume fraction in meshes of fixed size at a rate of ,62 which implies that ill-conditioning is more severe for high-order basis (see Fig. 10).
Several strategies have been proposed to alleviate ill-conditioning of trimmed cells, such as the ghost penalty method,112 the artificial stiffness approach,110,111 the extended B-spline method,113–116 or special preconditioning techniques specifically designed for immersed boundary methods,117 among others. We recommend here the extended B-spline approach by Höllig et al.113–116 due to its simplicity, ease of implementation, and good performance, cf. Fig. 10. The main idea consists on removing the critical basis functions (the ones with smaller intersection with ) from the approximation space and extrapolating well-behaved basis functions from neighboring cells toward the cut cell. Figure 11 illustrates this process for the univariate case. The modified basis has less degrees of freedom, but the condition number and error converge rates are independent of , and equivalent to those of body-fitted (untrimmed) methods.113 The extended B-spline basis stabilization can be easily implemented as a linear constraint on the approximation space of cut cells based on the uniform Cartesian structure of the discretization.62,113–116
C. Beyond uniform meshes: Local h-refinement via hierarchical B-splines
So far, uniform meshes have been considered. However, in practice, spatial -adaptivity is required to increase the resolution of the approximation where needed. In the context of multivariate B-spline bases, refining the knot vectors around a precise location is not effective since the tensor product structure of the basis yields a delocalized refinement that spans along the whole mesh.
Alternatively, one can resort to hierarchical B-spline refinement,118 where the approximation space is locally enriched by replacing selected coarse B-spline basis functions (parents) with finer ones (children), see Fig. 12. This process can be performed recursively, leading to a parent-children hierarchy spanning several levels of refinement, where each refinement level is defined by its own uniform knot vector set.
In practical computations, the goal is to refine some cells in the mesh. In a hierarchical B-spline context, it corresponds to refining some of the non-vanishing basis functions on those cells. There exist different hierarchical refinement strategies, depending on the relation between cells and basis to be refined,120–124 yielding to more or less localized refinement.
D. Nondimensionalization of the equations
A subtle implementation detail of major practical relevance is considering the nondimensionalization of all the variables and parameters in the equations, specially since one is dealing with a coupled problem with multiple physics. To understand its relevance, think of the and blocks in the diagonal of the system matrix . If one considers the International System of Units (SI), their magnitudes are expressed in Pa and F/m, corresponding to Young’s modulus and electric permittivity of the material. In a typical dielectric, they are in the order of (or higher) and (or lower), respectively, which makes at least orders of magnitude difference! This would preclude the solution of the numerical problem due to a very large condition number of .
the electric potential ,
Young’s modulus ,
the flexoelectric coefficients ,
the electric permittivity ,
and so on. Typically, suitable values for the normalizing factors are , , and , where is a characteristic length of the geometry.
The normalized flexoelectric coefficients scale inversely proportional to the size of the geometry, whereas the normalized Young modulus and electric permittivity remain constant. This fact reveals the well-known size-dependent nature of the flexoelectric coupling, relevant only at sub-micrometer scales.
IV. COMPUTATIONAL EXPERIMENTS OF CANTILEVER BEAM BENDING WITH DIFFERENT FORMULATIONS
In this section, we make use of the numerical methods described in Sec. III to solve different boundary values problems (BVP) stated in Sec. II C. We focus on cantilever beam bending, the most well-known benchmark for flexoelectricity, widely used by experimentalists to characterize the transversal flexoelectric effect.85,87,125–127 It has also been studied numerically58,59,63 and analytically.128,129 In the experiments in Secs. IV A and IV B, we analyze the electromechanical response of microscopic cantilevers under applied force and applied electric potential in closed circuit, respectively, considering the Direct and Lifshitz-invariant flexoelectricity formulations.
A. Cantilever beam bending
We consider a 2D (plane strain) cantilever beam of length and thickness . The material properties are simple enough to isolate the transversal flexoelectric effect, i.e., a Young modulus GPa, electric permittivity nC/V m, and transversal flexoelectric coefficient . The other material parameters are set to . For a complete description of material tensors, we refer to Appendix A.
The left tip is clamped and a vertical force is applied on the top right corner. The right tip is electrically grounded and the other boundaries are free, which corresponds to open-circuit electrical boundary conditions [Fig. 14(a)]. The transversal flexoelectric effect is triggered due to the mechanically-induced gradient of axial strains along the beam’s cross section.
The problem is numerically solved by means of the body-fitted B-spline method in Sec. III. The results are shown in Fig. 14. As pointed out in Sec. II A, the Lifshitz-invariant and direct flexoelectricity boundary value problems are not equivalent, even though their associated Euler–Lagrange Eqs. (53) and (65) coincide. The difference is due to the different boundary terms (e.g., tractions and surface charges) in Eqs. (54) and (66) given the different expressions in (49) and (64). In particular, traction-free or charge-free boundary conditions have different meaning in each formulation and thus the resulting fields are different.
Two main differences are pointed out next. First, the mechanical results are quite similar. The axial strains vary linearly along the cross sections of the bent beam in both cases [Fig. 14(c)], as expected. However, the flexoelectricity-induced stiffening of the beam41 is different in each case. Comparing the maximum deflection of the direct-flexoelectric beam ( ) and the Lifshitz-invariant-flexoelectric beam ( ) with respect to a standard elastic one ( ), it becomes apparent that the effective stiffness is increased around in the former and a in the latter.
Nevertheless, the most interesting difference arises in the electrical response, as shown in Figs. 14(b) and 14(d). While the direct flexoelectricity form presents an electric potential varying linearly along a cross section of the beam, the Lifshitz-invariant form features boundary layers with opposite sign than that of the bulk. This phenomenon is highlighted in Fig. 14(d), which depicts the transversal electric fields along the middle cross section.
The electric response in the Lifshitz-invariant model resembles the predicted theoretical response of flexoelectricity in the presence of surface piezoelectricity.35 Since the fundamental difference between the two studied BVPs is just the null Lagrangian in Eq. (1), which is also acting on the surface, a relation may exist between the modeling of null Lagrangians and surface effects in this context.
As discussed later on in Appendix B, the appearance of a boundary layer may lead to mesh-dependent spurious oscillations in the numerical solutions. Hence, some stabilization technique is required in that case. The simulations presented in Secs. IV A and IV B are properly stabilized according to the approach described in Appendix B.
B. Cantilever beam actuator
In this experiment, we explore the transversal flexoelectric effect triggered by electrical actuation. This device was first used by Bursian and Zaikovskii3 to experimentally demonstrate for the first time the flexoelectric effect, which had been predicted theoretically by Mashkevich and Tolpygo.11 Computational studies are also present in Abdollahi et al.58 and Zhuang et al.59
We consider the same geometry and material properties as in previous experiment. Here, an electric field across the beam thickness is enforced by attaching an electrode on the top boundary at prescribed voltage V, while the bottom boundary is grounded, which corresponds to closed-circuit electrical boundary conditions [Fig. 15(a)]. Mechanically, the left tip is clamped, and no force is applied. Due to the transversal flexoelectric effect, the electric field will yield an axial strain gradient along the thickness of the beam, inducing a constant curvature.
The results are shown in Fig. 15 and are quite similar to the ones reported in the cantilever bending case. The two differences between the direct and Lifshitz-invariant flexoelectricity models are also present here. The latter presents more stiffening, in view of the maximum deflections obtained: for the direct case and only for the Lifshitz-invariant one. Boundary layers in the electric field distribution are also obtained here for the Lifshitz-invariant form [Fig. 15(d)]. However, in this case the bulk electric field is much larger than the boundary layer effect, and hence the electric potential distributions are much more alike [Fig. 15(b)].
V. COMPUTATIONAL ANALYSIS OF A SOFT CANTILEVER ROD ACTUATOR
In this section, we illustrate the richer physics in the regime of finite deformations by modeling a soft cantilever rod actuator.41 This device was considered by Bursian and Zaikovskii3 back in 1968 to experimentally demonstrate for the first time the flexoelectric effect.
Let us consider the direct flexoelectricity boundary value problem stated in Sec. II C 4 c in a compliant slender rod of length and thickness , see Fig. 16(a). The left tip of the rod is clamped, while all other boundaries are traction-free. The electric potential at the top boundary is set to a certain value , and the bottom boundary is grounded ( ). The voltage drop induces a transverse electric field across the rod thickness which triggers (i) the flexoelectric effect, thereby generating a non-uniform axial strain that bends the rod, and (ii) the electrostrictive effect, yielding a uniform axial strain that elongates the rod. The induced curvature and axial stretch are constant throughout the neutral axis of the rod. Since the flexoelectric effect is reversible in sign, the cantilever will bend upwards or downwards depending on the direction of the applied electric field. However, since the electrostrictive effect is not, the beam length will always increase regardless of the electric field direction.
To solve the resulting non-linear system of equations, the voltage is applied incrementally in a sequence of load steps. We resort to a modified-step monolithic Newton–Raphson algorithm presented in Codony et al.,41 which is robust against the nonlinearities in the formulation, including the eventual presence of mechanical (geometrical) instabilities such as wrinkling, creasing, or buckling.
VI. COMPUTATIONAL ANALYSIS OF FLEXOELECTRIC DEVICES AND METAMATERIALS
We analyze next several devices and metamaterials exploiting the flexoelectric effect through the effective accumulation of the response to gradients in the mechanical and electric fields in the limit of infinitesimal deformation. The example in Sec. VI A shows a scalable flexoelectricity-based deformation sensor which works upon collective beam-bending, resulting in a potential difference that accumulates from beam to beam. In Secs. VI B and VI C, we simulate the RVE corresponding to different engineered, geometrically-polarized architected materials and show that they feature apparent piezoelectric behavior without piezoelectric constituents, allowing its use in sensing or actuation applications.
A. Collective-beam bending scalable flexoelectric device
Previous works have shown that bending of thin structural elements is the most efficient way to mobilize flexoelectricity.9,75 We present here a setup achieving collective beam bending in such a way that the flexoelectrically-generated electric potential at the thin structural elements can be effectively accumulated throughout the structure and collected at the electrodes. The structure is composed of several thin beams connected by vertical elements on one end and to other beams through insulating connectors at the other end. The insulating connectors are placed as to achieve a non-centrosymmetric system and thus avoid internal cancelation of flexoelectrically-generated electric potential of opposite signs at the beams.83 This unit is appropriately repeated in series, producing a scalable device, here shown in 2D in Fig. 17(a). We analyze such a structure, with beam thicknesses of nm and lengths of nm, and beam spacings of nm, under three point displacement as shown in Fig. 17(a), with nm. The direct formulation for flexoelectricity in Sec. II C 1 is used, with the interface conditions in Sec. II C 4 2 at the dielectric-insulator interfaces. The material parameters are reported in Table I.
Material | Y | ν | ℓmech | ɛ | μL | μT | μS |
(GPa) | — | (nm) | (nC/V m) | (μC/m) | (μC/m) | (μC/m) | |
Matrix | 152 | 0.33 | 1 | 141 | 150 | 110 | 110 |
Insulators | 152 | 0.33 | 1 | 0 | 0 | 0 | 0 |
Material | Y | ν | ℓmech | ɛ | μL | μT | μS |
(GPa) | — | (nm) | (nC/V m) | (μC/m) | (μC/m) | (μC/m) | |
Matrix | 152 | 0.33 | 1 | 141 | 150 | 110 | 110 |
Insulators | 152 | 0.33 | 1 | 0 | 0 | 0 | 0 |
Figure 17(b) shows the electric potential, , distribution. The accumulation of electric potential from the grounded (bottom-left) toward the top-right edge is apparent. This electric potential accumulation scales linearly with the number of connected beams.
B. Geometrically-polarized periodic inclusions
According to Sharma et al.75 and Mocci et al.,83 geometrically-polarized inclusions embedded in a non-piezoelectric matrix produce an apparently piezoelectric response upon macroscopic homogeneous deformation. This response results from the effective accumulation of the flexoelectric response to the localized strain gradients. With this idea, one can endow any dielectric with apparent piezoelectricity.83
We illustrate this idea by simulating the electromechanical response of flexoelectric metamaterial with triangular voids under homogeneous macroscopic vertical compression, Fig. 18. We consider the Lifshitz-invariant flexoelectric formulation, and barium strontium titanate (BST) in its paraelectric phase as base material, see Table II for material parameters. Paraelectric BST is a non-piezoelectric dielectric with high permittivity, and hence, a good flexoelectric.35
Y | ν | ℓmech | ɛ | ℓelec | μL | μT | μS | |
(GPa) | — | (nm) | (nC/V m) | (nm) | (μC/m) | (μC/m) | (μC/m) | |
152 | 0.33 | 50 | 8 | 300 | 1.2 | 1.1 | 0.05 |
Y | ν | ℓmech | ɛ | ℓelec | μL | μT | μS | |
(GPa) | — | (nm) | (nC/V m) | (nm) | (μC/m) | (μC/m) | (μC/m) | |
152 | 0.33 | 50 | 8 | 300 | 1.2 | 1.1 | 0.05 |
To efficiently evaluate the performance of the periodic structure, high-order generalized periodic conditions in Eq. (76) are enforced in a single unit cell. Standard periodicity for the solution fields along the horizontal direction is applied ( ), whereas the jump in electric potential along the vertical direction ( ) is resolved as a result of a prescribed vertical jump in the displacement field ( ).
Figure 18 (middle) shows the resulting electric potential distribution on the deformed sensor. The plot on the right represents the trend of the electric potential along the vertical dashed line by computing a portion of the whole domain (thin black line) and a single unit cell with a height of m with generalized periodicity conditions (thick magenta line). Both simulations produce exactly the same result on a unit cell far away from the top and bottom boundaries. A non-vanishing net potential difference between the top and bottom layers of the device is achieved upon homogeneous deformation. The observed apparent piezoelectricity is the result of the accumulation of the flexoelectric response to strain gradients generated around the voids. This accumulation is in turn possible because the triangular voids produce a geometrical polarization of the microstructure. Such a material exhibits also inverse apparent piezoelectricity under the application of a homogeneous electrical bias.83
C. Bending-dominated lattice metamaterial
Building on the ideas discussed in Secs. VI A and VI B, a new class of flexoelectricity-based metamaterials have been recently proposed by Mocci et al.83 These metamaterials are bending dominated, low area fraction lattices, with a non-centrosymmetric arrangement of small non-piezoelectric dielectric beams. Upon homogeneous macroscopic deformation or electric bias, the metamaterial behaves as an apparent piezoelectric due to the accumulation of the flexoelectric response of the lattice beam components. Mocci et al.83 show that it can reach performances comparable in some situations to those of very common piezoelectric materials, such as quartz or PZT.
Figure 19 shows a bending-dominated BST lattice metamaterial in actuation mode, i.e., deforming due to an applied macroscopic electric bias ( ), modeled by the Lifshitz-invariant flexoelectric formulation. A RVE of the lattice is analyzed using generalized periodicity conditions on the top and bottom faces and standard periodicity in the horizontal direction. The lattice unit cell is has dimensions m and m and the beam thickness is nm. The resulting vertical displacement ( ) and the deformed configuration are also depicted in a reduced portion of the whole domain. The metamaterial behaves again as an apparent piezoelectric despite its base material is not.
VII. CONCLUSIONS
We have reviewed the mathematical modeling of the flexoelectric effect in the regime of infinitesimal deformations, including different flexoelectric couplings, variational principles, and boundary value problems and presented its natural extension to the regime of finite deformations. We have also described in detail an efficient computational approach based on immersed-boundary hierarchical B-splines, enabling the computation on arbitrarily-shaped geometries with attractive numerical properties such as trivial mesh generation and optimal convergence rates. The provided numerical examples illustrate the mechanisms behind the flexoelectric effect and show the feasibility of engineered devices and materials designed for a variety of electromechanical applications. More specifically, we have shown that the Lifshitz-invariant formulation leads to boundary layers, and we have proposed material architectures capable of mobilizing the flexoelectric effect at a microstructural level and upscaling this effect to a mesoscale.
This paper exemplifies how accurate computational methods can enable new material designs that exploit flexoelectricity in engineering applications, for instance, to yield apparent piezoelectricity in metamaterials made of non-piezoelectric base materials,83 given the inherent complexity of setups mobilizing field gradients. With this background, we expect that materials designs can be optimized geometrically and topologically,44,130,131 and that ultimately the fundamental design principles of flexoelectric metamaterials can be distilled. There are some aspects that still hinder the quantitative evaluation of flexoelectricity. A clear physical interpretation of the high-order fields is still missing, which would allow us to solve the ambiguity regarding the high-order boundary conditions on physical grounds. Continuum models would benefit from a more precise connection with first principles. This would allow us not only to identify parameters for the continuum descriptions but also learn proper enrichments in order to capture additional physics, e.g., finite sample effects. These important effects, such as surface piezoelectricity and surface flexoelectricity, have not been included here.35 Finally, a wider and more precise experimental quantification of flexoelectricity in an ample catalog of materials would also be desirable.
Avenues for future research in the mathematical and computational modeling of flexoelectricity are vast. One of the most appealing ones is the complete understanding and exploitation of flexoelectricity in soft materials. Despite larger strain gradients (hence, larger flexoelectric response) are easily possible together with functional deformation modes, the fundamental mechanisms leading to flexoelectricity in polymers remain unclear.36 Other interesting research lines involve the understanding and exploitation of flexoelectricity in other important multifunctional materials, such as magnetoelectrics, liquid crystal elastomers, ferroelectrics, or semiconductors, with multiple potential applications.
ACKNOWLEDGMENTS
This work was supported by the European Research Council (No. StG-679451 to I.A.), the Spanish Ministry of Economy and Competitiveness (No. RTI2018-101662-B-I00), and the Generalitat de Catalunya (ICREA Academia award for excellence in research to I.A. and Grant No. 2017-SGR-1278 to I.A.). CIMNE is a Severo Ochoa Centre of Excellence (2019-2023) under the grant CEX2018-000797-S funded by MCIN/AEI/10.13039/501100011033.
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.
APPENDIX A: MATERIAL CHARACTERIZATION
The material enthalpy density in an infinitesimal deformation framework is characterized by the tensors of elasticity , strain gradient elasticity , dielectricity , flexoelectricity , and eventually gradient dielectricity .
a. Elasticity tensor c
The Lamé parameters and are related to the Young modulus and Poisson’s ratio as and .
b. Strain gradient elasticity tensor h
c. Dielectricity tensor κ
d. Gradient dielectricity tensor M
e. Flexoelectricity tensor μ
The isotropic flexoelectric tensor is a particular case of the cubic one with only two independent parameters, where .
APPENDIX B: RESIDUAL-BASED WEAK FORM STABILIZATION
As discussed in Remark 8, depending on the geometry of , the boundary conditions, the material parameters, and the mesh size, the numerical solution to the flexoelectricity problem features spurious oscillations that completely spoil the quality of the results. Figure 20 shows an example, with spurious oscillations arising in a Lifshitz-form flexoelectric cantilever beam with open circuit boundary conditions undergoing bending (i.e., a similar setting to Sec. IV A). The beam is long by thick. The dielectric material has Young’s modulus GPa, Poisson’s ratio , dielectric permittivity , and longitudinal and transversal flexoelectric coefficients . Other material coefficients are . A force of is applied vertically on the top of the right end.
In view of the results in Secs. IV A and IV B, and in agreement with the experience of this paper’s authors, the numerical instabilities are associated with the presence of boundary layers in the electric field, which may frequently appear when considering the Lifshitz-invariant flexoelectricity formulation.
In the literature of computational flexoelectricity, the direct flexoelectricity form is much more popular than others, specially for the computational examples. Hence, it is difficult to say whether numerical instabilities are also present in other numerical schemes rather than smooth B-spline approximations.