Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/349164616

A Collocation Mixed Finite Element Method for the Analysis of Flexoelectric


Solids

Article in International Journal of Solids and Structures · February 2021


DOI: 10.1016/j.ijsolstr.2021.01.031

CITATIONS READS

14 328

5 authors, including:

Xinpeng Tian Jan Sladek


Huazhong University of Science and Technology Slovak Academy of Sciences
9 PUBLICATIONS 69 CITATIONS 475 PUBLICATIONS 8,435 CITATIONS

SEE PROFILE SEE PROFILE

Qian Deng
Huazhong University of Science and Technology
63 PUBLICATIONS 1,401 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Analysis of functionally-graded quantum dot (FGQD) systems View project

Flexoelectricity based energy harvesting: theory and experiments View project

All content following this page was uploaded by Xinpeng Tian on 22 March 2021.

The user has requested enhancement of the downloaded file.


International Journal of Solids and Structures 217-218 (2021) 27–39

Contents lists available at ScienceDirect

International Journal of Solids and Structures


journal homepage: www.elsevier.com/locate/ijsolstr

A collocation mixed finite element method for the analysis of


flexoelectric solids
Xinpeng Tian a,b, Jan Sladek a,⇑, Vladimir Sladek a, Qian Deng b, Qun Li b
a
Institute of Construction and Architecture, Slovak Academy of Sciences, 84503 Bratislava, Slovakia
b
State Key Laboratory for Strength and Vibration of Mechanical Structures, School of Aerospace Engineering, Xi’an Jiaotong University, Xi’an 710049, China

a r t i c l e i n f o a b s t r a c t

Article history: A collocation mixed finite element method (MFEM) for direct and converse flexoelectricity in piezoelec-
Received 23 September 2020 tric materials is developed for 2D problems. The size-effect phenomenon in micro/nano structures is con-
Received in revised form 14 December 2020 sidered by the strain- and electric intensity vector-gradient effects. C0 continuous finite element method
Accepted 27 January 2021
is inadequate to treat flexoelectricity problems involving the size-effect. To this end, the MFEM with
Available online 09 February 2021
Lagrangian multipliers to treat these solids has been reported recently. With existing MFEM, the compu-
tational efficiency is low due to the additional nodal degrees of freedom (DOFs) for the Lagrangian mul-
Keywords:
tipliers. In this study, a new collocation MFEM is proposed, in which the number of the DOFs, when
Flexoelectricity
collocation MFEM
compared to the traditional Lagrangian approach, can be reduced. At the same time, the kinematic con-
Gradients of strain and electric intensity straints between the displacement and strain are guaranteed. These kinematic constraints are satisfied by
vector the collocation method at some internal points in the finite elements. The present collocation MFEM can
Higher efficiency be used to solve flexoelectricity problems with higher efficiency. Its accuracy is verified by comparing the
Geometric dependence numerical results with available analytical solutions for the bending of a cantilever beam and the com-
pression of a truncated pyramid, respectively. The results indicate that flexoelctricity is strongly related
to the geometry of the physical problem. It is shown that flexoelectricity increases significantly with the
decrease of the sample size. The same occurs when, for the beam problem, the ratio of the length to depth
dimensions increases; similarly, for the truncated pyramid problem, when the ratio of the width of the
bottom and top surfaces increases.
Ó 2021 Elsevier Ltd. All rights reserved.

1. Introduction tricity was only performed a few decades later in the 2000s, see,
e.g. Ma and Cross (2001), Ma and Cross (2006). Perspectives on
Recent progress in microelectronics is, in large part, driven by the future directions for research on flexoelectricity are given in
continuous miniaturization of devices and the use of nanotech- some review papers (Krichen and Sharma, 2016; Wang et al.,
nologies in which size-dependent effects cannot be ignored. The 2019; Zhuang et al., 2020; Deng et al., 2020).
direct flexoelectric effect describes the coupling between the strain To utilize the flexoelectric effect, the strain gradients have to be
gradients and the electric polarization (Kogan, 1964; Meyer, 1969; relatively large; they are therefore more easily generated in nano-
Sharma et al., 2006). A non-uniform strain, i.e. the presence of scale structures. The dimensions of such solids are of the same
strain gradients, may potentially break the inversion symmetry order of the material length scale parameter used in generalized
thereby inducing electric polarization even in centrosymmetric theories of continua. As has been shown experimentally, the stiff-
crystals (Tagantsev, 1986; Tagantsev et al., 2009; Maranganti ness of such a structure increases with a decrease in its size. The
et al., 2006). Flexoelectricity is found to be a universal electrome- presence of strain gradients can also be realized by differences in
chanical coupling that exists in all dielectric materials even with material properties at the interfaces of these materials even under
a centrosymmetric crystal structure (Yudin and Tagantsev, 2013; a uniform stress (Deng et al., 2014a, 2014b).
Deng et al., 2014a, 2014b, 2020). The existence of flexoelectric In order to perform good design of devices with flexoelectric
effect in solids has been observed experimentally as first reported properties, it is necessary to analyze general boundary value prob-
by Harris (1965). However, a systematic measurement of flexoelec- lems (BVPs) of these components. It is well-known that classical
continuum mechanics neglect the influence of the material
⇑ Corresponding author. microstructure and the results are size-independent. To overcome
E-mail address: jan.sladek@savba.sk (J. Sladek). intrinsic limitations of classical elasticity, atomistic models have

https://doi.org/10.1016/j.ijsolstr.2021.01.031
0020-7683/Ó 2021 Elsevier Ltd. All rights reserved.
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

been developed to describe the micro-scale phenomena in materi- dient is enforced by Lagrangian multipliers. In the formulation, the
als. Extremely high requirements on computer memory in these displacement gradient and Lagrangian multipliers are set as addi-
models have led to the development of multiscale approaches tional nodal DOFs to displacements, electric potential and polariza-
where atomistic and continuum subdomains are bridged. Due to tion (Amanatidou and Aravas, 2002; Mao et al., 2016). Therefore,
intrinsic heterogeneities in the atomistic-continuum coupling, extra DOFs are introduced thereby leading to lower computational
some physically unrealistic phenomena have been observed espe- efficiency. For example, there are 87 DOFs for a 2D quadrilateral
cially for time-dependent problems. Another approach to treat this element in formulation by Mao et al. (2016); and the correspond-
problem is based on the phenomenological theory of flexoelectric- ing number is 47 in Deng et al’s formulation. This is a significant
ity within the generalized thermodynamics of a dielectric. This drawback for analysing large scale models. A new finite element
involves introducing a flexoelectric contribution into the total free method with greater efficiency would therefore be highly desir-
energy (Tagantsev, 1986). The flexoelectricity can be realized as able; this is the subject of the present paper.
the direct flexoelectric effect (Sharma et al., 2006; Gharbi et al., In this study, a collocation MFEM for direct and converse flex-
2009), and converse flexoelectricity. The former refers to the linear oelectricity in piezoelectric materials is developed. The size-effect
coupling of electric fields and strain gradients, and it is frequently is considered by including the strain gradients, electric field gra-
investigated in the literature. The converse flexoelectricity is the dients, and their coupling in the constitutive equations of the
coupling between the stress and applied electric intensity vector piezoelectric materials. The C0 continuous approximation is
gradients (Yang et al., 2004). Maugin (1980) showed the duality applied independently for displacements and strains. The kine-
between the theory of electric field gradient and the theory of flex- matic constraints between strains and displacements are satisfied
oelectricity. Hu and Shen (2009) and Shen and Hu (2010) have by collocation method at judiciously chosen internal points of the
extended the general flexoelectric theory by the surface effects elements (Dong and Atluri, 2011; Bishay et al., 2012). In contrast,
for nano-sized elastic dielectrics. They have developed the varia- the Lagrange multipliers are used to enforce these kinematic con-
tional principle for these problems. Recently, flexoelectricity in straints in the traditional MFEM (Amanatidou and Aravas, 2002;
biological membranes and soft electrets has attracted the attention Mao et al., 2016). The electric potential and electric intensity vec-
of several researchers. The corresponding flexoelectric membrane tor are approximated by C0 continuity in the same manner. The
theory and nonlinear flexoelectric theory in electret soft materials corresponding constraint between the electric intensity vector
have been proposed (Mohammadi et al., 2014; Deng et al., 2014a, and the electric potential are also satisfied by the same colloca-
2014b; Rahmati et al., 2019). tion method. No extra DOFs are introduced in this collocation
For nano-sized flexoelectric structures it is necessary to apply MFEM. Thus, there are just 12 DOFs for a 2D quadrilateral ele-
gradient theory in which the governing equations are partial differ- ment (eight displacements and four potentials). In the case of iso-
ential equations (PDE) of the fourth order. To solve flexoelectric geometric analysis, the non-uniform rational B-spline functions
problems with strain- and electric intensity vector-gradient effects, with higher order continuity can be employed; no extra DOFs
several numerical methods have been developed, namely, moving are therefore needed (Thai et al., 2018), like in the present
least square (MLS) (Sladek et al., 2013), meshfree formuation approach.
method (Abdollahi et al., 2014, 2015), isogeometric analysis (IGA) A MFEM code based on this formulation is developed in this
(Thai et al., 2018; Nguyen et al., 2018, 2019; Liu et al., 2019), hier- study and its veracity is demonstrated with two example prob-
archical B-spline method (Codony et al., 2019), and finite element lems, The first is a simple cantilever beam problem for which the
method (FEM) (Yvonnet and Liu, 2017; Sladek et al., 2018; analytical solution is available. The second example is a truncated
Amanatidou and Aravas, 2002; Mao et al., 2016; Deng et al., pyramid under a compressive load. Before discussing these numer-
2017, 2018). ical examples, a review of the governing equations in flexoelectric-
Among the numerical methods mentioned above, the FEM has ity is perhaps in order. This will be followed by a presentation of
been well established as a powerful computational tool for analys- the formulation of the proposed collocation MFEM.
ing general BVPs with complex geometries. However, the conven-
tional FEM cannot be used to study flexoelectricity due to the 2. Direct and converse flexoelectricity
strain- and electric intensity vector-gradient effects, where second
derivatives of the primary fields (displacement and electric poten- The electric enthalphy density for piezoelectric solids can be
tial) are required. Two modified FEMs to resolve this problem have written as (Maranganti et al., 2006; Hu and Shen, 2009)
been reported in the literature. The first approach involves the use
of C1 continuity elements. For example, Yvonnet and Liu (2017)
1 1 1
H¼ cijkl eij ekl  aij Ei Ej  ekji eij Ek þ g jklmni gjkl gmni  f ijkl Ei gjkl
have applied C1 Argyris triangular elements for soft flexoelectric 2 2 2
solids at finite strains. Sladek et al. (2018) have developed conform- 1
 bklij eij Ek;l  hijkl Ei;j Ek;l ð1Þ
ing elements with C1 continuity, where each node has 9 degrees of 2
freedom (6 mechanical quantities, electric potential, and two where symbols a and c are used for the second-order permittiv-
potential gradients) for 2D flexoelectric problems. The C1 continu- ity and the fourth-order elastic constant tensors, respectively. The
ous element is established by using higher order shape functions. piezoelectric coefficient is denoted by e and f is the direct flexo-
It is, however, difficult to develop C1 elements for 3D problems. electric coefficient. The tensor g is used for higher order elastic
Another way to resolve the gradient problem is to use mixed finite coefficients representing the strain-gradient elasticity. The sym-
element methods (MFEMs) which are relatively more convenient to bols b and h are used for the converse flexoelectric coefficients
develop. Following the works of Amanatidou and Aravas (2002) for and higher-order electric parameters, respectively. The strain ten-
the MFEM in gradient theory of elasticity, Mao et al. (2016) con- sor eij and the electric field vector Ej are defined as (Parton and
structed a MFEM formulation for flexoelectricity with extra nodal Kudryavtev, 1988)
degrees-of-freedom (DOFs) for polarizations and developed a 2D
1 
element to solve general BVPs. Deng et al. (2017), Deng et al. eij ¼ ui;j þ uj;i ; Ej ¼ /;j ð2Þ
(2018) have also developed a MFEM with strain gradient and flex- 2
oelectricity and extended it to 3D flexoelectricity problems. where ui and / are displacements and electric potential,
In traditional MFEM (Mao et al., 2016; Deng et al., 2017, 2018), respectively.
the kinematic relationship between displacement field and its gra- The strain-gradient tensor g is given by
28
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

1  The constitutive Eq. (4) for orthotropic materials


gijk ¼ eij;k ¼ ui;jk þ uj;ik ð3Þ
2 (aij ¼ a1 di1 dj1 þ a2 di3 dj3 , c ¼ di1 dj1 ðc11 dk1 dl1 þ c13 dk3 dl3 Þ þ di3 dj3
 ijkl 
Under the infinitesimal deformations, the constitutive equa- ðc13 dk1 dl1 þ c33 dk3 dl3 Þ þ c44 di1 dj3 þ di3 dj1 ðdk1 dl3 þ dk3 dl1 Þ)
tions can be obtained from the electric enthalphy density expres- can be rewritten into a matrix form as (Lekhnitskii, 1963)
2
sion (1) (Hu and Shen, 2009; Shen and Hu, 2010) r11 3 2 c11 c13 0 32 e11 3 2 0 e31 3
@H 6 7 6 76 7 6 7
rij ¼ ¼ cijkl ekl  ekij Ek  bklij Ek;l 4 r33 5 ¼ 4 c13 c33 0 54 e33 5  4 0 e33 5
@ eij
r13 0 0 c44 2e13 e15 0
2 3
@H 2 3 E1;1
sjkl ¼ ¼ f ijkl Ei þ g jklmni gnmi " # b1 0 0 b1 6 7
@ gjkl E1 6 76 E3;1 7
 4 b1 0 0 b1 56
6E 7 ¼
7
@H
E3 4 1;3 5
Di ¼  ¼ aij Ej þ eijk ejk þ f ijkl gjkl 0 b3 b2 0
@Ei E3;3
2
@H e11 3 " # " #
Q ij ¼  ¼ bijkl ekl þ hijkl Ek;l ð4Þ E1 E1;3
@Ei;j 6 7
¼ C4 e33 5  K U ð11Þ
E3 E3;3
Where rij , Di , sjkl and Q ij are the stress tensor, electric displace- 2e13
ments, higher order stress and electric quadrupole, respectively.
2
The size scale of higher-order elastic parameters g jklmni is " # " e11 3 " # #" #
expressed by a proportionality of the conventional elastic stiffness
D1 0 0 7 a1 0
e15 6 E1
¼ e33 5 þ 4
coefficients cklmn and the internal length material parameter l D3 e31 e33 0 0 a2 E3
(Gitman et al., 2010; Yaghoubi et al., 2017) 2e13
2
2 g111 3
g jklmni ¼ l cjkmn dli ð5Þ 6 7
6 g331 7
" #66 7
with dli being the Kronecker delta. 7
f 1 þ 2f 2 f 1 0 0 0 f 2 6 2g131 7
Similarly, the higher-order electric parameters h is expressed in þ 6 7
6 7
terms of the dielectric constants akl and another length-scale 0 0 f 2 f 1 f 1 þ 2f 2 0 6 g113 7
6 7
parameter q as 6 g 7 ð12Þ
4 333 5
hijkl ¼ q2 aik djl ð6Þ 2g133
2 3
Deng et al. (2017) considered two independent components f 1 g111
6
and f2 for

the direct

flexoelectric coefficient f ijkl ,
2 6 g 7 7
f ijkl ¼ f 1 djk dil þ f 2 dij dkl þ dik djl . Then, the electric enthalphy density e11 3 " # 6 331 7
6 7
6 7 E1 6 2g131 7
has the following form ¼¼ KT 4 e33 5 þ P þ F66 g
7;
7
E3 6 113 7
1 1 l
2
2e13 6 7
H¼ cijkl eij ekl  aij Ei Ej  ekji eij Ek þ cjkmn gjkl gmnl  f 1 Ei gkki 6 g 7
2 2 2 4 333 5
  q2 2g133
 f 2 Ei gikk þ gjij þ bklij eij Ek;l  aik Ei;j Ek;j ð7Þ
2 2
s111 3 2
f 1 þ 2f 2 0
3
The number of independent converse flexoelectric coefficients 6 7 6 7
bijkl may be reduced as follows. If in the poling direction is along 6 s331 7 6 f1 0 7
6 7 6 7" #
the x3 -axis in the piezoelectric material, the stresses induced by 6 7 6 7
6 s131 7 6 0 f2 7 E1
6 7 ¼ 6 7
electric intensity vector can be written as 6 7 6 7
6 s113 7 6 0 f1 7 E3
r11 ¼ e31 E3 ; r33 ¼ e33 E3 ; r13 ¼ e15 E1 ð8Þ 6 7 6 7
6s 7 6 f 1 þ 2f 2 7
    4 333 5 4 0 5
with ekij ¼ e31 di1 dj1 þ e33 di3 dj3 dk3 þ e15 di1 dj3 þ di3 dj1 dk1 , where s133 f2 0
2 32
g111 3
standard Voight notation is applied for piezoelectric coefficients
(Sladek et al., 2018).
c11 c13 0 0 0 0
6 76 7
Analogously, consider a similar form for induced stresses by the 6 c13 c33 0 0 0 0 76 g331 7
6 76 7
converse flexoelectricity 6 76 7
26
0 0 c44 0 0 0 76 2g131 7
rij ¼ dij b1 ðE1;1 þ E3;3 Þ; r13 ¼ r31 ¼ b2 E1;3 þ b3 E3;1 ð9Þ þl 66
76
76
7
7 ð13Þ
6 0 0 0 c11 c13 0 76 g113 7
6 76 7
with the converse flexoelectric coefficients reduced into three 6 0 0 c13 c33 0 7 6 7
4 0 54 g333 5
independent coefficients b1 , b2 and b3 by
  0 0 0 0 0 c44 2g133
bklij ¼ b1 dij dkl þ di1 dj3 þ di3 dj1 ðb2 dk1 dl3 þ b3 dk3 dl1 Þ. With this reduc- 2 3
tion, the electric enthalphy has the following form g111
6 7
1 1 6 g331 7
H¼ cijkl eij ekl  aij Ei Ej  e31 e11 E3  e33 e33 E3  e15 ðe13 þ e31 ÞE1 " # 6 7
6 7
2 6
2 2 E1 2g131 7
  ¼¼ F T
þ l G66 7:
l
2
7
þ cjkmn gjkl gmnl  f 1 Ei gkki  f 2 Ei gikk þ gjij þ b1 ekk Ei;i E3 6 g113 7
2 6 7
6 g 7
q2 4 333 5
þ ðb2 E1;3 þ b3 E3;1 Þðe13 þ e31 Þ  aik Ei;j Ek;j ð10Þ
2 2g133

29
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

2 3 2 3
Q 11 b1 b1 0 2 3 set as independent variables with C0 continuity. However, the
6Q 7 6 0 e11
6 31 7 6 0 b3 7
76 strains should also satisfy the geometric relationship with the
6 7¼6 74 e33 7
5 displacements. To this end, the strain values from these two con-
4 Q 13 5 4 0 0 b2 5
2e13 siderations are made to be equal at the Gauss points in the ele-
Q 33 b1 b1 0
2 32 3 2 3 ð14Þ ments (Dong and Atluri, 2011; Bishay et al., 2012). The
a1 0 0 0 E1;1 2 3 E1;1 kinematic constraints between strains and displacements are
60 e11
6 a2 0 0 7 6E 7
76 3;1 7 6
6E 7
e33 7 6 3;1 7 thus satisfied by the collocation method at Gauss points. This
þ q2 6 76 7 ¼ UT 4 5 þ q2 H 6 7
40 0 a1 0 54 E1;3 5 4 E1;3 5 is similarly applied to the electric potential and electric intensity
2e13 vector. By using this collocation scheme, each node has only
0 0 0 a2 E3;3 E3;3
three DOFs (two displacements and one electric potential) which
Recently, the authors have derived the governing equations for is significantly smaller in number as compared to the traditional
piezoelectric solid with direct and converse flexoelectric effects MFEM via Lagrangian multipliers (Mao et al., 2016; Deng et al.,
(Sladek et al., 2018) 2017).
rij;j ðxÞ  sijk;jk ðxÞ ¼ 0 The variational formulation of the FEM in gradient theory can
be derived from the principle of virtual work as (Hu and Shen
Di;i ðxÞ  Q ij;ji ðxÞ ¼ 0 ð15Þ 2009)
Z  
The essential and natural boundary conditions (b.c.) can be
prescribed:
rij deij þ sijk dgijk þ Dk dEk þ Q ij dEi;j dX
V
Z 
Z 
Z 
Z 
 ¼ t i dui dC þ Ri dsi dC þ S d/dC þ Z dpdC ð23Þ
1) Essential b.c.: ui ðxÞ ¼ ui ðxÞon Cu , Cu  C Ct CR CS CZ


si ðxÞ ¼ s i on Cs ; Cs  C ð16Þ    
Where t i , Ri , S, and Z are prescribed values corresponding to the
 external work on the right hand side of (23).
/ðxÞ ¼ /ðxÞ on C/ ; C/  C The mechnanical displacements and electric potential in each
element as shown in Fig. 1 are expressed in terms of nodal values
@/  and shape functions
pðxÞ ¼ ¼ pðxÞ on Cp ; Cp  C
@n !
  ðeaÞ
u1 ðxÞ  P4 u1
¼ N a ðn1 ; n2 Þ
u ðxÞ 
3 V e
a¼1
u3
ðeaÞ
 !
2) Natural b.c.:t i ðxÞ ¼ t i ðxÞ on Ct Ct [ Cu ¼ C; Ct \ Cu ¼ £ P4 u1
ðeaÞ
ð24Þ
ðeaÞ a ðeaÞ
¼ a¼1 q N ðn1 ; n2 Þ; q :¼ ðeaÞ
 u3
Ri ðxÞ ¼ Ri ðxÞ on CR ; CR [ Cs ¼ C; CR \ Cs ¼ £ ð17Þ P4 ðeaÞ a
/ðxÞjV e ¼ a¼1 / N ðn1 ; n2 Þ

SðxÞ ¼ SðxÞ on CS ; CS [ C/ ¼ C; CS \ C/ ¼ £
where qðeaÞ and /ðeaÞ are nodal displacements and electric
 potential, respectively.
ZðxÞ ¼ Z ðxÞ on CZ ; CZ [ Cp ¼ C; CZ \ Cp ¼ £ The gradient in global coordinates can be expressed within the
Where finite element V e in terms of deivatives in the local (intrinsic) coor-
dinates as
@ui @/
si :¼ ; p :¼ ; Ri :¼ nk nj sijk ; Z :¼ ni nj Q ij ð18Þ    
@n @n @=@x1  @=@n1 1
¼ ½Y e  ; ½Y e  ¼ ½J e 
and the traction vector, and the electric charge are defined as @=@x  3 Ve @=@n2
  @ qi X
t i :¼ nj rij  sijk;k  þ k qi ðxc Þ kdðx  xc Þ ð19Þ   !
@p c @x1 =@n1 @x3 =@n1  X ðecÞ
x1 Nc;1
ðecÞ
x3 Nc;1
½J e  ¼ ¼ ð25Þ
@x1 =@n2 @x =@n  Ve
ðecÞ
x1 Nc;2
ðecÞ
x3 Nc;2
 @a X
3 2
 c
S :¼ nk Dk  Q kj;j  þ k aðxc Þ kdðx  xc Þ ð20Þ
@p c Hence
 P ðeaÞ ea
with qi :¼ nk pj sijk ; a :¼ ni pj Q ij , d(x) being the Dirac delta func- @f 
@xi  e
¼ a f bi ðn1 ; n2 Þ;
V
tion and pi is the Cartesian component of the unit tangent vector ea
b1 ðnÞ ¼ Y e11 ðnÞNa;1 ðnÞ þ Y e12 ðnÞNa;2 ðnÞ; ð26Þ
on C.
ea
The jump at a corner (xc) on the oriented boundary contour C is b3 ðnÞ ¼ Y e21 ðnÞNa;1 ðnÞ þ Y e22 ðnÞNa;2 ðnÞ
defined as
The expression for the approximation of the electric intensity
k qi ðxc Þ k :¼ qi ðxc þ 0Þ  qi ðxc  0Þ ð21Þ vector within V e , from the Maxwell equation, is given by
  !
k aðxc Þ k :¼ aðxc þ 0Þ  aðxc  0Þ ð22Þ E1  /;1 
fEgjV e ¼ ¼ 
E3 V e /;3 
Ve
! ð27Þ
3. The mixed finite element using collocation method X n ea o ðeaÞ n ea o ea
b1 ðnÞ
¼ B/ ðnÞ / B/ ðnÞ ¼ ea
a b3 ðnÞ
To solve the strain- and electric intensity vector-gradient
effects in flexoelectric materials, a collocation MFEM is devel- The electric intensity vector is treated as independent variable;
oped. In the present scheme, the displacements and strains are thus,
30
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Fig. 1. Global Cartesian coordinates x1  x2  x3 , curvilinear coordinates n1  n2 , for the 4-node quadrilateral finite element.

0 1 0 1
b1i b1i Substituting Eqs. (31) into (28), one can obtain
 B 2C B 2C  X
 B bi C B C 
EIn ¼ ð 1 n1 n2 n1 n2 Þ B C ¼ fPðnÞgT B bi C EIn
i ðxÞ ¼ fLðnÞgT xea /ðeaÞ ; fLðnÞgT :¼ fPðnÞgT ½A1 ð32Þ
i ðxÞ ð28Þ
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} B 3C B 3C i
Ve
Ve @ bi A @ bi A a
fPðnÞgT
b4i b4i or
! !
or n o EIn  X fLðnÞgT xea
 1 ðxÞ 
/ðeaÞ
1
n oT   EIn ðxÞ  ¼  ¼
  Ve E ðxÞ 
In T
fLðnÞg xea
 EIn ðxÞ  ¼  EIn  a
e
1 ðxÞ EIn
3 ðxÞ V e
3 V 3
V e X  mea
1 ðnÞ

0 1 1 ¼ /ðeaÞ
b1 b13 a 3 ðnÞ
mea
B 2 C ð29Þ
B b1 b23 C X
¼ ð 1 n1 n2 n1 n2 ÞB
B 3
C
3C
T ea
@ b1 b3 A i ðnÞ :¼ fLðnÞg
mea xea
i ¼ fLðnÞgc bi ðnc Þ ð33Þ
c
b41 b43
and
with the coefficients bci being determined from equating the n oT    X
two approximations at collocation points xc , selected as the Gauss
   EIn ðxÞ  ¼  EIn
1 ðxÞ EIn  ¼
3 ðxÞ V e 1 ðnÞ
ð mea 3 ðnÞ Þ/
mea ðeaÞ

points, in the element V e with intrinsic coordinates nc1 ; nc2 , i.e.


e
V a

EIn ð34Þ
i ðx Þ ¼ Ei ðx Þ. Thus,
c c

0 1 0 10 1 1  
  
EIn
i ðx Þ
1
 1 n11 n12 n11 n12 bi Since @x@1  ¼ Y e11 @n@1 þ Y e12 @n@2 ; @x@3  ¼ Y e21 @n@1 þ Y e22 @n@2 , the electric
Ve Ve
B In 2 C  B CB C
B Ei ðx Þ C B 1 n21 n22 n21 n22 C B b2i C intensity vector gradient can be derived as
B C B CB C
B In 3 C ¼ B 3 3 3 3 CB 3 C

@ Ei ðx Þ A @ 1 n1 n2 n1 n2 A @ bi A   
 @EIn
 e i ðxÞ @ @
EIn
i ðx Þ
4
1 n41 n41 n41 n42 b4i   ¼ Y e11 þ Y e12 fPðnÞgT fbi g
V
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} @x1  @n1 @n2
½A ð30Þ Ve |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
0 1 fP 1 ðnÞgT
0 1 Bea 1

Ei ðx1 Þ /i ðn Þ X
 B ea 2 C ¼ fS1 ðnÞgT xea /ðeaÞ
B E ðx2 Þ C XB C i
B i C B B/i ðn Þ C ðeaÞ ea c ea c a
¼ B C ¼ B ea 3 C/ ; B/i ðn Þ ¼ bi ðn Þ;
@ Ei ðx3 Þ A B C
a @ B/i ðn Þ A
   
E ðx4 Þ  Bea 4 @EIn 
i ðxÞ @ @
Ve /i ðn Þ
i
  ¼ Y e21 þ Y e22 fPðnÞgT fbi g
@x3  @n1 @n2
Hence, Ve |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
0 1 0 1 fP 3 ðnÞgT
b1i
ea
bi ðn1 Þ X
B 2C ¼ fS3 ðnÞg T
xea /ðeaÞ
B bi C XBB bea 2 C
C
i

fbi g :¼ B C ¼ ½A1 B i ðn Þ C/ðeaÞ a


B 3C B ea 3 C
@ bi A a @ bi ðn Þ A
½A1
T T
ea Sj ðnÞ :¼ Pj ðnÞ ð35Þ
b4i bi ðn4 Þ
0 ea 1 1 ð31Þ
bi ðn Þ or
X B ea 2 C  X X
B bi ðn Þ C 
1
/ðeaÞ ; xea :¼ B C ðeaÞ ea
¼ ½A xea
i i B ea 3 C EIn
i;j ðxÞ ¼ ij ðnÞ/
lea ; lea
ij ðnÞ :¼ Sj ðnÞ c bi ðnc Þ ð36Þ
@ bi ðn Þ A
e
V
a a c
ea
bi ðn4 Þ
and its matrix form is
31
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

0 1 oT 1
0n
EIn  01
0n o 1 1;1 ðxÞ  11 ðnÞ
lea Bea 1
 B In C 0 1B eij ðn Þ C
EIn ðxÞ  B E ðxÞ C X B l31 ðnÞ C
B a 1
Bn C
B ea 2 oT C
ea
B ;1 C B 3;1 C C ðeaÞ ij
@ n o A ¼ B In C ¼ B ea C/ ð37Þ B 2C X B B ðn Þ C
 B E ðxÞ C @ l ðnÞ A B ij C
a B eij C
EIn ðxÞ  @ 1;3 A aij :¼ B C 1
B a3 C ¼ ½A Bn C
a 13
;3 e  B o T C
V
EIn ðxÞ  lea
33 ðnÞ @ ij A a B B ðn Þea 3
C
3;3 Ve B eij C
4
aij @ n oT A
A similar derivation can be repeated for the approximation of ea 4
Beij ðn Þ
the strain tensor. The expression for the strain tensor obtained 0n oT n o1 ð44Þ
ðeaÞ
from the geometric relationship within V e is given by Bea ðn1 Þ qu
0 1 B eij C
Bn C
ea
b1 ðnÞ 0 B ea 2 oT n ðeaÞ o C
X B C XB Beij ðn Þ qu C Xn o
fegjV e ¼ Bea ðeaÞ
; Bea
ea B C
e ðnÞ qu e ðnÞ ¼ @ b3 ðnÞ A; 1 1
0 qðeaÞ ¼ ½ A Bn o n o C ¼ ½ A cea
1 ea ea
u
B
a B Bea ðn3 Þ
T
ðeaÞ C
ij
a
b ðnÞ 1
b ðnÞ ð38Þ q C a
! 2 3 2 1 B eij u
C
ðeaÞ @n oT n oA
u1 ðeaÞ
qðeaÞ
u ¼ Bea 4
eij ðn Þ qu
ðeaÞ
u3
in which
The strain tensor is also set as independent variable; thus, 0n oT n o1
ðeaÞ
0 Bea ðn1 Þ qu
11 0 11 B e ij C
aij aij Bn C
B 2C B 2C B ea 2 oT n ðeaÞ o C
 B aij C Ba C n o B Beij ðn Þ qu C
 B C
^eIn
ij ðxÞ ¼ ð 1 n1 n1 n2 ÞB C ¼ fPðnÞgT B ij C; cea :¼ B n C ð45Þ
B ea 3 oT n ðeaÞ o C
n2 B a3 C B a3 C ij
Ve @ ij A @ ij A B Beij ðn Þ qu C
ð39Þ B C
a4ij a4ij @n oT n oA
( ea 4
Beij ðn Þ
ðeaÞ
qu
eInij ; if i ¼ j
^eIn
ij :¼
ij ; if i–j
2eIn Then,
0 1
a1ij
or B 2C
  Ba C Xn o Xn o
 T B ij C
eIn11 ðxÞ eIn33 ðxÞ 2eIn13 ðxÞ V e ^eIn
ij ðxÞ e ¼ fPðnÞg B 3 C ¼ fPðnÞg ½A
T 1
cea ¼ fLðnÞgT cea
@ aij A |fflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflffl}
ij ij
V
0 1 1 a a
a11 a133 a113 a4ij
fLðnÞgT
B 2 C
B a11 a233 a213 C ð40Þ
¼ ð 1 n1 n2 n1 n2 ÞB B 3
C ð46Þ
@ a11 a333 a313 C
A or
a411 a433 a413 0 1 0 1
eIn11 ðxÞ 
T
fLðnÞg cea
n o XB 11
C
where the coefficents acij (c ¼ 1; 2; 3; 4) are determined from the ^eIn ðxÞ  B In C
¼ @ e33 ðxÞ A ¼ B fLðnÞgT cea C
@ 33 A
ð47Þ
Ve 
collocation of ^eij ðxc Þ ¼ ^eij ðxc Þ at the Gauss points xc in the finite ele-
In  e a
2eIn 13 ðxÞ
T
V fLðnÞg cea
13
ment V e with Since
 X n ea c oT 0n oT 1
eij ; if i ¼ j 
^eij :¼ ; ^eij ðxc ÞV e ¼ Beij ðn Þ quðeaÞ ð41Þ Bea ðn1 Þ
B eij C
2eij ; if i–j Bn C
B ea 2 oT C
a
n o  n o n o B Beij ðn Þ C
where B C
fLðnÞg
T
cea ¼ fLðnÞgT bea qðeaÞ ea
; bij :¼ B n C
  ij ij u
B ea 3 oT C
Bea c T ea B Beij ðn Þ C
e11 ðn Þ ¼ 0 ;
b1 ðnc Þ B C
T   T   ð42Þ @n oT A
Bea c
e33 ðn Þ ¼ 0 bea c
3 ðn Þ
; Bea c
e13 ðn Þ ¼ bea c
3 ðn Þ
ea
b1 ðnc Þ ea 4
Beij ðn Þ
Thus, ð48Þ
0 1 0 10
^In  11 The following is finally obtained,
ij ðx Þ
1
e  1 n11 n12 n11 n2 aij 1
B C B CB C 0 In 1 0 1
B In 2 C
B ^eij ðx Þ C B 2 2 CB 2 C ^e11 ðxÞ  fLðnÞgT bea
B 1 n1 n2 n1 n2 C B aij C
2 2
B C  B C  XB 11
C
B C ¼ BB
CB C eIn ðxÞ V e ¼ B In C
@ ^e33 ðxÞ A ¼
B fLðnÞgT bea C
B ^In 3 C 3 3 CB 3 C @ 33 A
e
B ij ðx Þ C B 1 n 3
n 3
n n CBa C 
@ A @ 1 2 1 2 A @ ij A
2^e13 ðxÞ V e
In a T
fLðnÞg bea
 0 1
13
ð49Þ
^eIn  4
1 n1 n2 n1 n2 4 4 4
a4ij
ij ðx Þ
4 T
Ve |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} Mea 11 ðnÞ
½A XB C X ea
qðeaÞ ¼ B Mea ðnÞ T C qðeaÞ ¼ M quðeaÞ
0n oT 1 ð43Þ u @ 33 A u
a a
Bea ðn1 Þ Mea
T
B eij C 0 1 13 ðnÞ
B C ^eij ðx1 Þ 
Bn oT C 
B Bea ðn2 Þ C B C where
XB eij C B ^eij ðx2 Þ C
B C ðeaÞ B C  ea 
¼ Bn C q ¼B C Mea
T
:¼ fLðnÞgT bea
B ea 3 oT C u B ^eij ðx3 Þ C 11 ðnÞ 11 ¼ k1 ðnÞ 0
a B
B ðn Þ C @ A
B eij C 
B C ^e ðx4
Þ  e  
@n oT A Mea
33 ðnÞ
T
:¼ fLðnÞgT bea
33 ¼ 0 kea
3 ðnÞ
ij V
ea 4
Beij ðn Þ
T  ea 
Mea
13 ðnÞ :¼ fLðnÞgT bea
13 ¼ k3 ðnÞ kea
1 ðnÞ
ð50Þ
Hence
32
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

and and the variational condition (23) becomes


X
4 X
4
kea fLðnÞgc b1 ðnc Þ; kea
ea
fLðnÞgc b3 ðnc Þ
ea
PP R n ðecÞ oT n ec n o
1 ðnÞ :¼ 3 ðnÞ :¼ ð51Þ ðeaÞ
T 2 T
e Ve
dqu
c;a ½M ðnÞ ½C M ea ðnÞ þ l Wec ðnÞ ½G Wea ðnÞ qu
c¼1 c¼1
  o
T T
Thus,  ½M ec ðnÞ ð½Kfmea ðnÞg þ ½Uflea ðnÞgÞ þ Wec ðnÞ ½FT fmea ðnÞg /ðeaÞ dV
n oT    PP R n  
^eIn ðxÞ  þ e c;a V e d/ðecÞ fmec ðnÞgT ½K Mea ðnÞ þ ½F Wea ðnÞ
¼ eIn11 ðxÞ eIn33 ðxÞ 2eIn13 ðxÞ V e n o
e
X
V
ð52Þ þflec ðnÞgT ½UT Mea ðnÞ
ðeaÞ
T  qu
¼ quðeaÞ Mea
11 ðnÞ Mea
33 ðnÞ Mea
13 ðnÞ   o
a þ fmec ðnÞgT ½PT fmea ðnÞg þ q2 flec ðnÞgT ½Hflea ðnÞg /ðeaÞ dV

with P P n ðecÞ oT R R ec

Ce fTðnÞgN ðnÞdC þ Ce fRðnÞgnj ðnÞbj ðnÞdC
c
 ea  ¼ e c dqu
T
  k1 ðnÞ
t R

qðeaÞ ¼ ; Mea ðnÞðeaÞ


u1¼ ; ðeaÞ
u3 R  R  
u 11 ðecÞ ec
0 þd/ CeS SðnÞN ðnÞdC þ
c
CeZ Z ðnÞnj ðnÞbj ðnÞdC ;
  ! ð53Þ
0 kea
3 ðnÞ ð59Þ
Mea
33 ðnÞ ¼ ea ; Mea
13 ðnÞ ¼
k3 ðnÞ kea
1 ðnÞ where Cet ¼ V e \ Ct , CeR ¼ V e \ CR , CeS ¼ V e \ CS , CeZ ¼ V e \ CZ ,

!  !
For derivatives of strains, the approximation of these gradients  
Ce ¼ @V \ V e , fT g ¼ t 1 , fRg ¼ R 
1
, dVjV e ¼ det½J e dn1 dn2 .
can be written as t3 R3
 Xn o
 Let xðesÞ and xðef Þ be the starting and final point, respectively, on
^eIn
ij;k ðxÞ ¼ fPk ðnÞgT aij ¼ fPk ðnÞgT ½A1 cea  
V e . Then, dCjCe ¼ xðef Þ  xðesÞ =2 dnv ¼ h ðnÞdnv with nv 2 ½1; 1
e ij e
V
a
Xn o
being the intrinsic variable along the side Ce of the finite element
¼ fSk ðnÞgT cea
ij ð54Þ
V e.
a  
Now, let eg cg be the global number of the node xg ,
Since
ðg ¼ 1; 2; :::; nÞ. Since the variations of the independent field vari-
n o  n o n oT
T
fSk ðnÞg cea ¼ fSk ðnÞgT bea qðeaÞ ¼ Xea quðeaÞ ables are arbitrary inside the analyzed domain and on its boundary,
ij ij u ijk ðnÞ ð55Þ
the last variational equation yields the system of algebraic
with equations
T T
Xea
111 ðnÞ 1 ðnÞ
:¼ ð vea 0 Þ Xea
331 ðnÞ :¼ ð 0 3 ðnÞ Þ
vea XXZ 1 Z 1 n
T 2 T

T ½Meg cg ðnÞ ½C Meg a ðnÞ þ l Weg cg ðnÞ ½G Weg a ðnÞ
X ea
131 ðnÞ :¼ ð v 3 ðnÞ
ea
v 1 ðnÞ Þ
ea
eg a 1 1
n o 
ðe aÞ T
T T qu g  ½Meg cg ðnÞ ð½Kfmeg a ðnÞg þ ½Ufleg a ðnÞgÞ
Xea
113 ðnÞ 1 ðnÞ
:¼ ð jea 0 Þ Xea
333 ðnÞ :¼ ð 0 3 ðnÞ Þ
jea
T X Z 1
Xea
133 ðnÞ 3 ðnÞ
:¼ ð jea 1 ðnÞ Þ
jea T
þ Weg cg ðnÞ ½FT fmeg a ðnÞg ¼
e
fTðnÞgNcg ðnÞh g ðnÞdnv
1
X ea
X ea
eg
Ceg  Cet
v j ðnÞ
ea
:¼ fS1 ðnÞgc bj ðnc Þ; j j ðnÞ
ea
:¼ fS3 ðnÞgc bj ðnc Þ ð56Þ  
X Z 1  
fRðnÞgnj ðnÞbj g g ðnÞh g ðnÞdnv /ðeg aÞ det½J eg dn1 dn2 ð60Þ
c c e c e
þ
1
this yields eg

0 1 Ceg  CeR
ð ea1 ðnÞ
v 0Þ
0 n In o 1 B C
ð0 3 ðnÞ Þ
v
ea
 B C P P R 1 R 1 n eg cg T 
 B
^e ðxÞ 
;1 C
XBBðv 3 ðnÞ
ea
1 ðnÞ Þ C
ea
v
C
C eg a 1 1 fm ðnÞg ½K Meg a ðnÞ þ ½F Weg a ðnÞ
gIn ðxÞ V e ¼ @n o A ¼ B
B C
n o
^eIn ðxÞ  a B ð eaj ðnÞ 0 Þ C þfleg cg ðnÞgT ½UT Meg a ðnÞ
ðe aÞ
qu g
 B
1
C ð57Þ   o
;3 Ve @ ð0 ea
ðnÞ
j Þ A 
3 þ fmeg cg ðnÞgT ½PT fmeg a ðnÞg þ q2 fleg cg ðnÞgT ½Hfleg a ðnÞg /ðeg aÞ det½J eg dn1 dn2
ðj ea
ðnÞ ea
ðnÞ
j Þ P R1 
X 3 1
¼ e
1
e
S ðnÞNcg ðnÞh g ðnÞdnv
g

qðeaÞ
u ¼ W ea
qðeaÞ
u P
Ceg  CeS
R1
Z  ðnÞnj ðnÞbj g g ðnÞh g ðnÞdnv
e c e
a þ eg
1
Ceg  CeZ
The above approximations further results in the following ð61Þ
expression being obtained,
 X T  The system of algebraic equations (60) - (61) can be applied for
rij deij V e ¼
T
dqðecÞ
u ½Mec ðnÞ ½C M ea ðnÞ quðeaÞ  ð½Kfmea ðnÞg þ ½Uflea ðnÞgÞ/ðeaÞ a general BVP described by the gradient theory with direct and
c;a
converse flexoelectricity.
 X  
sijk dgijk  e ¼
T T
½FT fmea ðnÞg/ðeaÞ þ l ½G Wea ðnÞ qðeaÞ
2
dqðecÞ
u Wec ðnÞ u
V
c;a
4. Numerical examples
X    
ðecÞ ðeaÞ
qðeaÞ
T
Dk dEk jV e ¼ d/ fm ðnÞg
ec
½P fm ðnÞg/
T ea
þ ½K M ðnÞ þ ½F W ðnÞ
ea ea

c;a
u A computer code for collocation MFEM based on the formula-
tions presented in the previous sections has been developed in this
 X ðecÞ   study. To demonstrate its veracity, two example problems are pre-
Q ij dEi;j V e ¼ d/ flec ðnÞg ½U Mea ðnÞ qðeaÞ þ q2 ½Hflea ðnÞg/ðeaÞ
T T
u sented here, namely, (a) a simple cantilever beam with a transverse
c;a
end load, and (b) a truncated pyramid under compression.
ð58Þ

33
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

4.1. A cantilever beam where

Fig. 2 shows a cantilever beam subjected to a tansverse end Z 1=2


load. With direct flexoelectricity, q ¼ 0, b1 ¼ b2 ¼ b3 ¼ 0. For the k X kL2 ¼ XT XdX ð65Þ
X
analysis, it is treated as a 2D problem. A similar problem has been
analyzed by Liang and Shen (2013) using the deformation asump- where xh is the present collocation MFEM result, and x is the
tions of the Bernoulli-Euler beam theory. exact analytic solution. The relative errors for the beam deflection
The length of the beam is L = 500 nm, the width is B = 0.5H and u3 and the electric field E3 are studied. There is, however, no exact
the thickness is H = 20 nm. The following boundary conditions are analytical solution for the present flexoelectric beam problem.
prescribed: Thus, the approximate analytical solution of flexoelectric
Z   Bernoulli-Euler beam will be used here as x, where the shear stres-
  
u3 jx1 ¼0 ¼ 0; u3;1 x ¼ 0; s111 x3 dA ¼ MH  ¼ 0; M H  ¼0 ses are neglected and all of the solutions are just a function of x1. It
1 ¼0 x1 ¼L
A x1 ¼0 should be reminded that the present 2D numerical solutions are
  d   functions of x1 and x3, which is closer to the actual physical

M þ P  M H;1  ¼ 0; M þ P  M H;1  ¼Q ð62Þ situation.
x1 ¼L dx1 x1 ¼L The convergence of the present collocation MFEM is investi-
R R gated using three different meshes with increasing refinement;
where M ¼ A r11 x3 dA and P ¼ A s113 dA.
they containing 320 elements, 900 elements and 1600 elements,
In this test problem, only direct flexoelectricity is considered.
respectively. Fig. 3 shows the relative errors for the deflection u3
Analytical results are available only for this special case, derived
and electric field E3 obtained using the three meshes for f1 = 1.0 
by Liang and Shen (2013). The deflection of the cantilever beam
107C/m and Q = 1nN, where h is the average element length. The
with flexoelectricity is given by
high convergence rates for both relative errors are clearly evident.
 
kx1 In what follows, the most refined mesh mentioned above, with
u3 ðx1 Þ ¼ C 1 þ C 2 x1 þ C 3 ðx1 Þ2 þ C 4 ðx1 Þ3 þ C 5 exp
L 1600 elements, is used for further numerical simulations of this
  beam problem.
kx1
þ C 6 exp  ð63Þ The distribution of beam deflection u3 along the cantilever
L
beam for the load Q = 1nN is presented in Fig. 4 with parameters
h   i f1 = 1.0  107C/m and f2 = 0. It is noted that the deflection
2
c11 I þ c11 l þ f 1 f 1 =a33 A L2 increases gradually from the clamped end to the free end. In the
k2 ¼ 2
c11 l I present 2D numerical simulation, the deflection u3 also vary along
x3, while it is just a function of x1 in the analytical solution. A com-
with I as the second moment of cross-section area. The expres- parision of the deflections between the numerical solutions and
sion for C 4 in the work of Liang and Shen (2013) should be replaced the analytical solutions is presented in Fig. 5 for different flexoelec-
2
by the correct one C 4 ¼ 6kQL
2
g I
.
11
tric coefficients (f1 = 0, f1 = 1.0  107C/m, f1 = 2.0  107C/m, and
The corresponding theoretical solution for the flexoelectric field f1 = 3.0  107C/m). Excellent agreement between the numerical
has also been given, and it is and analytical results is observed for various flexoelectric coeffi-
cients; the very small differences between them are due to
2
f1 f d u3 assumptions in the analytical solution which yields a linear varia-
E3 ¼  g ¼ 1 ð64Þ
a33 113 a33 dx21 tion of normal stresses r11 along x3 and vanishing shear stresses. In
Fig. 5, it is also observed that the deflection of the beam decreases
The piezoelectric material, PZT-5H is considered in this study..
gradually as the flexoelectric coefficients increase in value. The
The analytical solution for the cantilever beam is proposed for iso-
classical beam theory predicts the largest deflection of the can-
tropic materials. To verify the results of the present collocation
tilever beam.
MFEM, the following isotropic material parameters are chosen:
The distributions of electric responses inside the cantilever
Young’s modulus E = 12.6  1010Pa, Poisson’s ratio v = 0.2, permit-
beam are presented in Fig. 6 for f1 = 1.0  107C/m. The free-end
tivity of the dielectric a33 = 13.0  109 C2/N/m2, internal material
of the beam is electrically grounded in the numerical simulations.
length l = 2 nm.
Fig. 6(a) shows a non-uniform electric potential distribution inside
First, the convergence of the present collocation MFEM is inves-
the cantilever beam induced by direct flexoelectricity. The value of
tigated, using the relative error e (x) for the L2-norm:
the electric potential decreases along the upper surface from the
k Xh  X kL2 clamped-end to the free-end, while the inverse trend for it is
eðXÞ ¼ ; shown on the lower surface of the beam. The distributions of the
k X kL2
electric field E3 are presented in Fig. 6(b). The electric field reaches
its maximum at the clamped-end, where the strain gradient is the
largest, thus resulting in maximum flexoelectricity. At the free-end
of the cantilever beam, the electric field is zero, satisfying the elec-
tric boundary condition.
A comparision of the electric field E3 between the numerical
solutions and the analytical solutions for different flexoelectric
coefficients are shown in Fig. 7. In the 2D numerical results, as
expected, the electric field E3 is not constant across the beam thick-
ness; however, the average electric field value across the beam
thicknes at each station along the beam is used for comparision
with the analytical solution. It can be seen in Fig. 7 that the numer-
ical solutions are in good agreement with the analytical results for
different flexoelectric coefficients. There is no electro-mechanical
Fig. 2. Cantilever beam with shear force load. transformation in classical beam theory, and vanishing electric
34
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Fig. 3. Relative errors and convergence rates for the present collocation MFEM. Fig. 5. Deflections of the cantilever beam with different flexoelectric coefficients.

Fig. 4. Distributions of beam deflection u3 (m) for the cantilever beam.

field is also seen here. From the present numerical simulation, it is


also observed that the electric fields of the cantilever beam for f1 =
2.0  107C/m, and f1 = 3.0  107C/m are more significant when
compared with f1 = 1.0  107C/m. From equation (64), it is noted Fig. 6. Distributions of electric responses inside the cantilever beam. (a) the electric
potential /(V), (b) the electric field E3 (V/m).
that the electric field E3 is determined by the flexoelectric coeffi-
cient and the strain gradient g113 together. A linear growth of the
electric intensity vector with the flexoelectric coefficient is moder- beam. The magnitude of electric field is smaller in value for the lar-
ated due to a rapid reduction of the strain gradient in (64). ger sample size, and larger electric field is produced for reduced
sample size. For example, the maximum value of electric field for
4.1.1. Dependence of flexoelectricity on the geometry of the cantilever R = 1 is about 2.5 times larger of that for R = 5. The apparent size
beam effect of flexoelectricity is clearly evident in this cantilever beam
In this section, the size dependence of flexoelectricity is first example problem, demonstrating the well-known phenomenon
investigated by using the same orthotropic material, PZT-5H of stronger flexoelectricity as the geometric size of the component
(Sladek et al., 2017, 2018) for the cantilever beam. decreases.
Another factor that may significantly affect the magnitude and
c11 ¼ 12:6  1010 Pa;c13 ¼ 5:3  1010 Pa; c33 ¼ 11:7  1010 Pa; c44 distribution of the electric field is the shape of the cantilever beam.
¼ 3:53  1010 Pa; For given beam length L = 500 nm, five beam thichnesses are con-
sidered to study the effect on flexoelectricity. The numerical results
are shown in Fig. 9. The beam thichness H are characterized by
a11 ¼ 15:1  109 C2 =N=m2 ; a33 ¼ 13:0  109 C2 =N=m2 ;
using the normalized thichness H* where H = H*  20 nm, and
the corresponding load is Q = H*  1nN. From Fig. 9, it can be
f 1 ¼ 1  107 C=m;f 2 ¼ 1  107 C=m ð66Þ observed that the electric field distributions monotonously
Fig. 8 shows the average electric field distributions of the can- decrease from the clamped-end to the free-end for the cantilever
tilever beam with five geometric sizes, where R is the size factor beam with different thicknesses. It is also noted that smaller H*
R = L/500 nm = H/20 nm = Q/1nN. Here, the material length param- results in larger electric field. For example, the maximum value
eter is taken to be l = 2 nm and only the direct flexoelectric effect of electric field for H* = 1 is about 10 times larger of that for
(q ¼ 0, b1 ¼ b2 ¼ b3 ¼ 0) is analysed. It is noted that the electric H* = 5. The results indicate that the electric fields increase with
field distributions of the beam have a similar trend for all of the decreasing H*. This is consistent with the fact that thinner beam
five geometric sizes, and it is seen that the electric field gradually leads to larger strain gradients, thus, stronger flexoelectricity.
decreases from the clamped-end to the free-end. It indicates that
larger strain gradients and stronger electro-mechanical coupling 4.2. A truncated pyramid
exist at the clamped-end of the cantilever beam. Additionally, it
can also be observed that the the electric fields induced by flexo- A truncated pyramid under a compression is another good flex-
electricity strongly depends on the geometric size of the cantilever oelectric example that has been extensively studied experimen-
35
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Fig. 7. Electric field of the cantilever beam with different flexoelectric coefficients. Fig. 9. Electric field distributions of the cantilever beam with five various
thicknesses.

Fig. 10. Schematic of a truncated pyramid.

cation MFEM analysis. The material constants of PZT-5H in Eq. (66)


are also used here with q ¼ 0, b1 ¼ b2 ¼ b3 ¼ 0, being interested in
the direct flexoelectric effect. In the numerical simulations, the
Fig. 8. Electric field distributions of the cantilever beam with five geometric sizes.
electric potential V on the top surface is unknown a priori but is
a constant. With the resulting V, the effective piezoelectric con-
tally and theoretically (Cross, 2006). Numerical simulations can stant deff can be calculated as (Abdollahi et al., 2014)
also be found in some works (Abdollahi et al., 2014; Codony a33 E2 a33 Vs2 c11
et al., 2019; Zhuang et al., 2020). Fig. 10 shows the geometry of a deff ¼ ¼ ð68Þ
e22 hF
truncated pyramid with top surface s1, bottom surface s2, and
height h. The electric boundary conditions are prescribed as fol- Fig. 11 shows the effective piezoelectric constant deff of the
lows: the bottom surface is grounded with zero electric potential truncated pyramid as a function of the normalization parameters
and an electrode is attached to the top surface that results in an h/h0, keeping s1 = h = s2/2 and h0 = 20 lm. For this model, the mate-
equipotential value there. The mechanical boundary conditions rial length parameter is taken to be l = 2 lm. Good agreement is
are: a uniform total load F is applied on the top surface with the obtained between the theoretical estimates using Eq. (67) and
bottom surface rigidly supported against vertical displacement, the present collocation MFEM results. As discussed in the can-
i.e., u2 = 0. Due to the different lengths of the top surface and bot- tilever beam problem earlier, the small difference between the the-
tom surface, a longitudinal strain gradient is produced resulting in oretical and the numerical results is due to the 2D effect, which is
flexoelectric polarization. In Cross’ study (Cross, 2006), the effec- simplified to 1D in the analytical solution. From Fig. 11, it can be
tive piezoelectric constant deff of the truncated pyramid is simpli- seen that the effective piezoelectric constant deff increases with
fied as decreasing values of the the normalization parameter- h/h0. An
apparent size-dependent effective piezoelectric constant deff is
  observed for the present truncated pyramid model without piezo-
s2  s1
deff ¼ f 1 ð67Þ electricity. The distributions of mechanical fields (longitudinal dis-
hs1
placement u2 and longitudinal strain e22) and flexoelectric
Note that the effective piezoelectric constant deff in Eq. (67) is responses (electric potential / and longitudinal electric field E2)
size-dependent. Plane strain condition is assumed in the 2D collo- are shown in Fig. 12. The results shown are obtained for
36
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Fig. 11. Effective piezoelectric constant deff as a function of the normalization


parameters h/h0, where the aspect ratio of the truncated pyramid is s1 = h = s2/2 and
Fig. 13. Effective electric field Eeff as a function of the normalization parameters h/
h0 = 20 lm.
h0, where the aspect ratio of the truncated pyramid is s1 = h = s2/2 and h0 = 20 lm.

h = 20 lm and F = 200 N. In can be seen in Fig. 12(b) that the lon- DV


Eeff ¼ ð69Þ
gitudinal strain distributions are non-uniform, and significant h
strain gradients exist inside the truncated pyramid. The corre- where DV represents the voltage difference between the the top
sponding electric potential distributions are shown in Fig. 12(c); surface and the bottom surface of the truncated pyramid, and h is
the electric boundary conditions are clearly satisfied, namely, the its height.
electric potential is zero on the bottom surface and the equipoten- For a given aspect ratio of the truncated pyramid model
tial value is 0.22 mV on the top surface. The longitudinal electric s1 = h = s2/2, changing h would lead to the change of s1 and s2 at
field distributions are plotted in Fig. 12(d), and the electric field the same time. Fig. 13 shows the variation of the effective electric
E2 is larger near the pyramid corners due to the strong stress con- field Eeff with the normalization parameter h/h0, where h0 = 20 lm,
centration accompanied with larger strain gradients. and a total load 10 MPa is uniformly applied on the top surface. It
can be observed that the effective electric field Eeff is larger for a
4.2.1. Dependence of flexoelectricity on the geometry of the truncated reduced sample size. This is in agreement with the results reported
pyramid in the literature (see, e.g., Deng, 2017; Codony et al., 2019). Similar
In this section, the effects of geometric size and shape factor of to the cantilever beam problem above, size-dependent flexoelec-
the truncated pyramid model on flexoelectricity are studied. To tricity is evident in this truncated pyramid model as well.
evaluate the flexoelectric response, the effective electric field Eeff To further understand the effect of shape factor on flexoelectric-
is defined as follows. ity, the aspect ratio s2 : h : s1 = 1 : 0.5 : R, is kept the same. Thus, for

Fig. 12. Distributions of mechanical fields and flexoelectric responses under uniform load on the top surface and rigid support on the bottom surface, the bottom surface is
grounded and an electrode is attached on the top surface. (a) longitudinal displacement u2 (m), (b) longitudinal strain e22, (c) the electric potential /(V), (d) the electric field E2
(V/m).

37
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Declaration of Competing Interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared
to influence the work reported in this paper.

Acknowledgements

This work was supported by the National Key R&D Program of


China (2017YFE0119800). The authors acknowledge the supports
by the Slovak Science and Technology Assistance Agency registered
under number SK-CN-RD-18-0005 and VEGA-2/0061/20.

References

Abdollahi, A., Millan, D., Peco, C., Arroyo, M., Arias, I., 2015. Revisiting pyramid
compression to quantify flexoelectricity: A three-dimensional simulation study.
Phys. Rev. B 91, 104103.
Abdollahi, A., Peco, C., Millan, D., Arroyo, M., Arias, I., 2014. Computational
Fig. 14. Effective electric field Eeff of the truncated pyramid with different aspect evaluation of the flexoelectric effect in dielectric solids. J. Appl. Phys. 116,
ratios, where s2 : h : s1 = 1 : 0.5 : R. 093502.
Amanatidou, E., Aravas, N., 2002. Mixed finite element formulations of strain-
gradient elasticity problems. Comput. Method. Appl. M. 191, 1723–1751.
Bishay, P.L., Sladek, J., Sladek, V., Atluri, S.N., 2012. Analysis of functionally graded
a given heigh h, the bottom length s2 is confirmed. Changing R multiferroic composites Using Hybrid/Mixed Finite Elements and Node-Wise
would lead to the change of the top length s1. As shown in Material Properties. CMC 29, 213–262.
Fig. 14, three values of h = 20 lm, 40 lm, and 60 lm, respectively Codony, D., Marco, O., Fernández-Méndez, S., Arias, I., 2019. An immersed boundary
hierarchical B-spline method for flexoelectricity. Comput. Method. Appl. M.
are chosen to study the shape effects on the effective electric field 354, 750–782.
Eeff of the truncated pyramid with a uniform load 10 MPa on the Cross, L.E., 2006. Flexoelectric effects: Charge separation in insulating solids
top surface. The effective electric field Eeff is found to increase with subjected to elastic strain gradients. J. Mater. Sci. 41, 53–63.
Deng, F., Deng, Q., Yu, W., Shen, S., 2017. Mixed finite elements for flexoelectric
decreasing value of R for all these three cases. As a smaller R value
solids. J. Appl. Mech. 84, 081004.
result in a larger difference between s1 and s2; this leads to larger Deng, F., Deng, Q., Shen, S.P., 2018. A three-dimensional mixed finite element for
strain gradients which produce a stronger flexoelectric effect flexoelectricity. J. Appl. Mech. 85, 031009–31011.
inside the truncated pyramid. Deng, Q., 2017. Size-dependent flexoelectric response of a truncated cone and the
consequent experimental measurement of flexoelectric properties. J. Appl.
Mech. 84, 101007–101011.
Deng, Q., Liu, L., Sharma, P., 2014a. Flexoelectricity in soft materials and biological
membranes. J. Mech. Phys. Solids 62, 209–227.
5. Conclusions Deng, Q., Liu, L., Sharma, P., 2014b. Electrets in soft materials: Nonlinearity, size
effects, and giant electromechanical coupling. Phy. Rev. E. 90, 012603.
In this paper, the governing equations with direct and con- Deng, Q., Lv, S., Li, Z., Tan, K., Shen, S., 2020. The impact of flexoelectricity on
materials, devices, and physics. J. Appl. Phys. 128, 080902.
verse flexoelectricity have been derived. A collocation MFEM for- Dong, L., Atluri, S.N., 2011. A simple procedure to develop efficient & stable hybrid/
mulation was presented and a 2D element also developed to mixed elements, and Voronoi cell finite elements for macro- & micromechanics.
solve standard BVPs in flexoelectricity. In this collocation MFEM CMC 24, 61–104.
Gharbi, M., Sun, Z.H., Sharma, P., White, K., 2009. The origins of electromechanical
formulation, C0 continuous approximation was applied for inde- indentation size effect in ferroelectrics. Appl. Phys. Lett. 95, 142901.
pendent displacement, strain, electric potential and electric inten- Gitman, I., Askes, H., Kuhl, E., Aifantis, E., 2010. Stress concentrations in fractured
sity vector variables. The kinematic constraints between primary compact bone simulated with a special class of anisotropic gradient elasticity.
Int. J. Solids Struct. 47, 1099–1107.
fields (displacements, potential) and secondary fields (strain, elec-
Harris, P., 1965. Mechanism for the shock polarization of dielectrics. J. Appl. Phys.
tric intensity) are satisfied by a collocation method at judiciously 36, 739–741.
chosen internal points in the elements. This scheme significantly Hu, S.L., Shen, S.P., 2009. Electric field gradient theory with surface effect for nano-
reduces the number of the DOFs when compared to the tradi- dielectrics. CMC 13, 63–87.
Krichen, S., Sharma, P., 2016. Flexoelectricity: A perspective on an unusual
tional Lagrangian approach, and at the same time, the kinematic electromechanical coupling. J. Appl. Mech. 83, 030801–30811.
constraints between the displacement and strain are guaranteed. Kogan, S.M., 1964. Piezoelectric effect during inhomogeneous deformation and
The corresponding 2D collocation MFEM code has also been acoustic scattering of carriers in crystals. Sov. Phys. Solid State 5, 2069–2070.
Lekhnitskii, S.G., 1963. Theory of Elasticity of an Anisotrophic Elastic Body. Holden-
developed. Two example problems have also been presented to Day.
demonstrate its veracity for obatining accurate and convergent Liang, X., Shen, S., 2013. Size-dependent piezoelectricity and elasticity due to the
solutions. In the first example, namely, a cantilever beam subject electric field-strain gradient coupling and strain gradient elasticity. Int. J. Appl.
Mech. 5, 1350015.
to an end load, both the deflection and the electric response Liu, C., Wang, J., Xu, G., Kamlah, M., Zhang, T.Y., 2019. An isogeometric approach to
obtained were in in good agreement with the results from theo- flexoelectric effect in ferroelectric materials. Int. J. Solids Struct. 162, 198–210.
retical solutions. In the second example of a truncated pyramid, Ma, W., Cross, L.E., 2001. Large flexoelectric polarization in ceramic lead magnesium
niobate. Appl. Phys. Lett. 79, 4420–4422.
good agreement of the effective piezoelectric constant with avail- Ma, W., Cross, L.E., 2006. Flexoelectricity of barium titanate. Appl. Phys. Lett. 88,
able solution in the literature was also obtained. The size- 2004–2007.
dependence of flexoelectricity was successfully predicted for both Mao, S., Purohit, P.K., Aravas, N., 2016. Mixed finite-element formulations in
piezoelectricity and flexoelectricity. P. R. Soc. A Math. Phy. 472, 20150879.
these two example problems using the collocation MFEM devel-
Maranganti, R., Sharma, N.D., Sharma, P., 2006. Electromechanical coupling in
oped in this study. Larger aspect ratios of the physical dimensions nonpiezoelectric materials due to nanoscale nonlocal size effects: Green’s
of the solid body lead to larger strain gradients and flexoelectric- function solutions and embedded inclusions. Phys. Rev. B 74, 014110.
ity. The collocation MFEM formulation developed in this paper Maugin, G.A., 1980. The method of virtual power in continuum mechanics:
Applications to coupled fields. Acta Mech. 35, 1–80.
can be applied to, for example, structural health monitoring of Meyer, R.B., 1969. Piezoelectric effects in liquid crystals. Phys. Rev. Lett. 22, 918–
flexoelectric sensors or actuators. 921.

38
X. Tian, J. Sladek, V. Sladek et al. International Journal of Solids and Structures 217-218 (2021) 27–39

Mohammadi, P., Liu, L., Sharma, P., 2014. A theory of flexoelectric membranes and Sladek, J., Stanak, P., Han, Z.D., Sladek, V., Atluri, S.N., 2013. Applications of the MLPG
effective properties of heterogeneous membranes. J. Appl. Mech. 81, 011007– method in engineering & Sciences: A review. CMES-Comp. Model. Eng. 92, 423–
11012. 475.
Nguyen, B., Zhuang, X., Rabczuk, T., 2018. Numerical model for the characterization Tagantsev, A.K., 1986. Piezoelectricity and flexoelectricity in crystalline dielectrics.
of Maxwell-Wagner relaxation in piezoelectric and flexoelectric composite Phys. Rev. B 34, 5883–5889.
material. Comput. Struct. 208, 75–91. Tagantsev, A.K., Meunier, V., Sharma, P., 2009. Novel electromechanical phenomena
Nguyen, B., Zhuang, X., Rabczuk, T., 2019. NURBS-based formulation for nonlinear at the nanoscale: phenomenological theory and atomistic modelling. MRS
electro-gradient elasticity in semiconductors. Comput. Method. Appl. M. 346, Bullet. 34, 643–647.
1074–1095. Thai, T.Q., Rabczuk, T., Zhuang, X., 2018. A large deformation isogeometric approach
Parton, V.Z., Kudryavtsev, B,A., 1988. Electromagnetoelasticity: piezoelectrics and for flexoelectricity and soft materials. Comput. Method. Appl. M. 341, 718–739.
electrically conductive solids, Taylor & Francis. Wang, B., Gu, Y., Zhang, S., Chen, L.Q., 2019. Flexoelectricity in solids: Progress,
Rahmati, A.H., Yang, S., Bauer, S., Sharma, P., 2019. Nonlinear bending deformation challenges, and perspectives. Prog. Mater. Sci. 106, 100570.
of soft electrets and prospects for engineering flexoelectricity and transverse Yaghoubi, S.T., Mousavi, S.M., Paavola, J., 2017. Buckling of centrosymmetric
(d31) piezoelectricity. Soft. Matter. 15, 127–148. anisotropic beam structures within strain gradient elasticity. Int. J. Solids Struct.
Sharma, P., Maranganti, R., Sharma, N.D., 2006. Electromechanical coupling in 109, 84–92.
nanopiezoelectric materials due to nanoscale nonlocal size effects: Green Yang, X.M., Hu, Y.T., Yang, J.S., 2004. Electric field gradient effects in anti-plane
function solution and embedded inclustions. Phys. Rev. B 74, 014110. problems of polarized ceramics. Int. J. Solids Struct. 41, 6801–6811.
Shen, S.P., Hu, S.L., 2010. A theory of flexoelectricity with surface effect for elastic Yudin, P., Tagantsev, A., 2013. Fundamentals of flexoelectricity in solids.
dielectrics. J. Mech. Phys. Solids 58, 665–677. Nanotechnology 24, 432001.
Sladek, J., Sladek, V., Stanak, P., Zhang, Ch., Tan, C.L., 2017. Fracture mechanics Yvonnet, J., Liu, L., 2017. A numerical framework for modeling flexoelectricity and
analysis of size-dependent piezoelectric solids. Int. J. Solids Struct. 113, 1–9. maxwell stress in soft dielectrics at finite strains. Comput. Method. Appl. M.
Sladek, J., Sladek, V., Wunsche, M., Zhang, C., 2018. Effects of electric field and 313, 450–482.
strain gradients on cracks in piezoelectric solids. Eur. J. Mech. A-Solid. 71, Zhuang, X., Nguyen, B.H., Nanthakumar, S.S., Thai, T.Q., Alajlan, N., Rabczuk, T., 2020.
187–198. Computational modeling of flexoelectricity – A review. Energies 13, 1326.

39

View publication stats

You might also like