This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Null Lagrangians in linear theories of micropolar type and few other generalizations of elasticity

Nirupam Basak1, Basant Lal Sharma2
1Department of Mathematics, Indian Institute of Technology Kanpur
Kanpur, U. P. 208016, India; [email protected]
2Department of Mechanical Engineering, Indian Institute of Technology Kanpur
Kanpur, U. P. 208016, India; [email protected]
Corresponding author email: [email protected]
Abstract

In the context of linear theories of generalized elasticity including those for homogeneous micropolar media, quasicrystals, piezoelectric and piezomagnetic media, we explore the concept of null Lagrangians. For obtaining the family of null Lagrangians we employ the sufficient conditions of H. Rund. In some cases a non-zero null Lagrangian is found and the stored energy admits a split into a null Lagrangian and a remainder. However, the null Lagrangian vanishes whenever the relevant elasticity tensor obeys certain symmetry conditions which can be construed as an analogue of the Cauchy relations.

Introduction

In the three-dimensional theory of classical linear elasticity, the function

Ψ(ϵ)=12ϵ[ϵ]{\Psi}(\boldsymbol{\epsilon})=\frac{1}{2}\boldsymbol{\epsilon}{\cdot}{{\mathbb{C}}}[\boldsymbol{\epsilon}]

represents the stored energy, where {{\mathbb{C}}} is the elasticity tensor [2] and the symmetric second order tensor ϵ\boldsymbol{\epsilon} depends on displacement vector field. It is known [36] that, for no choice of {\mathbb{C}} other than 𝟘\mathbb{0}, with the symmetries typical of elasticity tensors can it correspond to a null Lagrangian. From a physical point of view, here, we recall that null Lagrangian refers to an energy density that does not contribute to the equilibrium equation for a given energy function [7] but rather accounts for environmental boundary effects. As an illustrative and significant example in the history of this subject, the topic of Frank constant k24k_{24} in the classical theory of liquid crystals is well known one [35]; indeed, Frank constant (in the role of material modulus) has been shown to be associated with a term that is a null Lagrangian [7]. As null Lagrangians cannot be interpreted as stored energy functionals in the linear theory of elasticity, i.e., borrowing a term from [7], nilpotent energies are absent. However, as found out by [34] (and [36]) a given elasticity tensor {\mathbb{C}} can be decomposed so that the stored energy functional admits a split into a null Lagrangian [6, 7] and a remainder which further reveals interesting connections with the six Cauchy’s relations [34, 2, 36, 40, 47]. Exploring along these lines, we naturally attempt to widen the scope of these investigations to generalized theories of elasticity. We recall that the classical theory of elasticity considers material particles as simple points having three translational degrees of freedom. The constitutive relations of such materials are characterized by symmetric Cauchy stress as a consequence of the Cauchy hypothesis where surface loading is determined by force vector, while neglecting surface couples [2, 23]. Occasionally, the classical model is found to be insufficient for some deformable solids such as cellular materials, foam-like structures (e.g., bones), periodic lattices, etc [32]. In this context, we invoke a three dimensional generalized elastic body which also has a rotational degree of freedom [1, 5, 13, 11, 15]. In the case of linear micropolar material [37], the stored energy is represented by

Ψ(ϵ,κ)=12ϵ𝔸[ϵ]+12κ𝔹[κ]+ϵ𝔻[κ],{{\Psi}}({{\upepsilon}},{{\upkappa}})=\frac{1}{2}{{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]+\frac{1}{2}{{\upkappa}}{\cdot}{{\mathbb{B}}}[{{\upkappa}}]+{{\upepsilon}}{\cdot}{{\mathbb{D}}}[{{\upkappa}}],

where 𝔸,{{\mathbb{A}}}, 𝔹{{\mathbb{B}}}, 𝔻{{\mathbb{D}}} are the fourth order tensors of micropolar elasticity while ϵ{{\upepsilon}} is the linearized strain tensor and κ{{\upkappa}} is the linearized wryness tensor depending on the displacement vector field and the rotation vector field. In the first part of this document we discuss this energy functional from the point of view of null Lagrangians. When 𝔻{{\mathbb{D}}} is non-zero it is commonly referred as a chiral (non-centrosymmetric) micropolar material [26]. In addition to the linear micropolar elasticity, there have been also several developments towards generalization of elasticity for quasi-crystals [19, 31], piezoelectric and piezomagnetic media [12, 31]. We also discuss these theories from the viewpoint of nilpotent energies albeit briefly. Such generalizations of the elasticity based models have important role in description of mechanical behaviour of the body in the presence of additional effects due to the presence of microstructure and other physical phenomenon. In fact, the micropolar kinematics itself is also deemed to be relevant for the piezoelectric or piezomagnetic materials since electromagnetic fields produce forces and moments [30, 28]. Based on the pioneering idea of [1], several variants of the theory [5] of media with microstructure have been proposed [13, 14, 37]. Overall there have been a tremendous amount of researches within these classes of models, over the last century as well last couple of decades, which we do not intend to survey in this document.

We restrict our study to those Lagrangian functions Ψ{\Psi} that belong to the null-space of the Euler–Lagrange operator; i.e., those functions for which the corresponding Euler–Lagrange equations are identically satisfied (see [17, 18, 29]; and also [7], [8], [9], [10], [24]). The characterization of the null Lagrangians goes back to the works of Caratheodory [3] (so called ‘theory of equivalent integrals’) and given in [4, 6]. The reader is also referred to the treatments given in [17] and [22]. It is also pertinent to mention the contribution of [27], who have given a representation for null Lagrangians through the use of differential forms and suggested a way to map certain traction data to null Lagrangians.

Notation

Let Ω\Omega denote an open, bounded region in three-dimensional space, 𝒏\boldsymbol{n} be the outward unit normal to the boundary Ω{\operatorname{\partial}}\Omega. Also let 𝒖{{\boldsymbol{u}}} denote the displacement vector field and ϕ{\boldsymbol{\phi}} be the rotation vector field over Ω\Omega. Let {𝒆1,𝒆2,𝒆3}\{{\boldsymbol{e}}_{1},{\boldsymbol{e}}_{2},{\boldsymbol{e}}_{3}\} be the standard basis for 3{{\mathbb{R}}}^{3}. Let SO(3)\mathrm{SO(3)} denote the set of all rotation tensors in three dimensions, i.e., for 𝐐SO(3)\mathbf{Q}\in\mathrm{SO(3)} 𝐐𝐐=𝐈,det𝐐=+1\mathbf{Q}^{\top}\mathbf{Q}=\mathbf{I},\det\mathbf{Q}=+1. For a skew tensor 𝐖\mathbf{W} (i.e., 𝐖=𝐖\mathbf{W}^{\top}=-\mathbf{W}) the axial vector 𝒘=axl𝐖,\boldsymbol{w}=\text{axl}\mathbf{W}, is defined by 𝐖𝒂=𝒘𝒂,𝒂3.\mathbf{W}\boldsymbol{a}=\boldsymbol{w}\wedge\boldsymbol{a},\forall\boldsymbol{a}\in\mathbb{R}^{3}. The differentiation of a function ff with respect to the xjx_{j} coordinate is written as f,j.f_{,j}. Note that 𝐐𝐐,j\mathbf{Q}^{\top}\mathbf{Q}_{,j} is a skew tensor for a rotation tensor field 𝐐\mathbf{Q} on Ω\Omega. A fourth order tensor, with components AijklA_{ijkl}, is denoted by 𝔸\mathbb{A} and its action on the second order tensors is expressed as 𝔸[𝐀]\mathbb{A}[\mathbf{A}] with its ijij component given by AijklAkl.A_{ijkl}A_{kl}. The inner product between two second tensors 𝐀\mathbf{A} and 𝐁\mathbf{B} is denoted by 𝐀𝐁\mathbf{A}\cdot\mathbf{B} and equals AijBijA_{ij}B_{ij}. The Einstein summation convention is used unless specified otherwise. For convenience as well as additional clarity the notation for indices is changed at some places in the manuscript as specified later. Also, the notation and symbols for other relevant physical entities are defined as appropriate.

1 Linearized micropolar model

In what follows we consider a micropolar elastic body where each point (particle) of the body is considered as infinitesimal rigid body with six degrees of freedom. The deformation of the body is described by a mapping from a fixed reference configuration into an current (deformed) configuration including rotations of the particles. The reference configuration of such body is a bounded domain denoted by Ω3\Omega\subset{\mathbb{R}}^{3}. We denote the microdeformation (describing the placement), and the microrotation tensor (describing the orientation of each particle) with

χ:Ω3,𝐑:ΩSO(3),\displaystyle\chi:\Omega\rightarrow\mathbb{R}^{3},\quad{\mathbf{R}}:\Omega\rightarrow\mathrm{SO(3)}, (1.1)

respectively. It is assumed that Ω\Omega represents the reference configuration of the micropolar body. Consider any part PP with 𝒏\boldsymbol{n} outward unit normal vector field on the surface P\operatorname{\partial}P then

𝒕=σ[𝒏],𝒎=μ[𝒏],\boldsymbol{t}={{\upsigma}}[\boldsymbol{n}],\quad\boldsymbol{m}={{\upmu}}[\boldsymbol{n}], (1.2)

where 𝒕\boldsymbol{t} is the traction vector field, 𝒎\boldsymbol{m} is the moment vector field, σ{{\upsigma}} is the stress tensor field and μ{{\upmu}} is the couple stress tensor field. In components, ti=σijnj,mi=μijnj.t_{i}={{\upsigma}}_{ij}n_{j},m_{i}={{\upmu}}_{ij}n_{j}. For the micropolar body to be in equilibrium,

P𝒃𝑑c+P𝒕𝑑s=𝟎, and P((𝒙𝒐)𝒃+𝒄)𝑑v+P((𝒙𝒐)𝒕+𝒎)𝑑s=𝟎,\displaystyle\int_{P}\boldsymbol{b}dc+\int_{{\operatorname{\partial}}P}\boldsymbol{t}ds=\boldsymbol{0},\text{ and }\int_{P}((\boldsymbol{x}-\boldsymbol{o})\wedge\boldsymbol{b}+\boldsymbol{c})dv+\int_{{\operatorname{\partial}}P}((\boldsymbol{x}-\boldsymbol{o})\wedge\boldsymbol{t}+\boldsymbol{m})ds=\boldsymbol{0}, (1.3)

(with 𝒐\boldsymbol{o} as the reference point) for every body part PΩP\subset\Omega. The equations of equilibrium in presence of an external body force (with components bib_{i}) and an external body couple (with components cic_{i}) are given by

σijxj+bi=0,μijxj𝙴ijkσjk+ci=0.\displaystyle\frac{{\operatorname{\partial}}{{\upsigma}}_{ij}}{{\operatorname{\partial}}x_{j}}+b_{i}=0,\quad\frac{{\operatorname{\partial}}{{\upmu}}_{ij}}{{\operatorname{\partial}}x_{j}}-{{\mathtt{E}}}_{ijk}{{\upsigma}}_{jk}+c_{i}=0. (1.4)

Using (1.1), the deformation gradient and the nonsymmetric right stretch tensor are [44] defined by

𝐅=χ=χ,i𝒆i,𝐔=𝐑𝐅,\displaystyle\mathbf{F}=\nabla\chi=\chi_{,i}\otimes{\boldsymbol{e}}_{i},\quad\mathbf{U}={\mathbf{R}}^{\top}\mathbf{F}, (1.5)

respectively. In combination with the material frame indifference, this leads to the relative Lagrangian stretch tensor, as strain measure, and the wryness tensor, as a Lagrangian measure for curvature, respectively defined by

𝐄=𝐔𝐈,𝐊=axl(𝐑𝐑,j)𝒆j,{\mathbf{E}}={\mathbf{U}}-{\mathbf{I}},\quad{\mathbf{K}}=\text{axl}\big{(}{\mathbf{R}}^{\top}{\mathbf{R}}_{,j}\big{)}\otimes{{\boldsymbol{e}}}_{j}, (1.6)

In the nonlinear theory of micropolar elasticity, the strain energy density function [44] is (𝐄,𝐊).{\mathcal{E}}({\mathbf{E}},{\mathbf{K}}). Let us consider the case when the translations 𝒖{{\boldsymbol{u}}} and microrotations ϕ{\boldsymbol{\phi}} are very small, where

𝒖:=χ(𝒙)𝒙,ϕ:=φ𝒆\displaystyle{{\boldsymbol{u}}}:=\chi(\boldsymbol{x})-\boldsymbol{x},\quad{\boldsymbol{\phi}}:=\varphi{\boldsymbol{e}} (1.7)

is the (small) translation vector and the (small) rotation vector (i.e., 𝐑𝐈𝙴[ϕ]\mathbf{R}\approx\mathbf{I}-{{\mathtt{E}}}[{\boldsymbol{\phi}}]), respectively, so that

𝐄𝒖+𝙴[ϕ],𝐊ϕ,\displaystyle{\mathbf{E}}\approx\nabla{{\boldsymbol{u}}}+{{\mathtt{E}}}[{\boldsymbol{\phi}}],\quad{\mathbf{K}}\approx\nabla{\boldsymbol{\phi}}, (1.8)

where, 𝙴{\mathtt{E}} is the Levi-Civita (alternating) tensor in three dimensions (with components defined by 𝙴ijk=𝒆i𝒆j𝒆k{{\mathtt{E}}}_{ijk}={\boldsymbol{e}}_{i}\cdot{\boldsymbol{e}}_{j}\wedge{\boldsymbol{e}}_{k}). Above is essentially the definition of an asymmetric strain and an asymmetric torsion obtained by the linearization around the natural state [37]. Indeed, the tensors ϵ{{\upepsilon}}, called the linear stretch tensor and κ{{\upkappa}}, called the linear wryness tensor, are defined by

ϵ=𝒖+𝙴[ϕ], and κ=ϕ,{{\upepsilon}}=\nabla{{\boldsymbol{u}}}+{{\mathtt{E}}}[{\boldsymbol{\phi}}],\quad\text{ and }{{\upkappa}}=\nabla{\boldsymbol{\phi}}, (1.9)

respectively. Thus, these are kinematic relations between the ‘small’ displacement field 𝒖{{\boldsymbol{u}}} and the ‘small’ rotation vector ϕ{\boldsymbol{\phi}} vs the strain tensor (interpreted as linearized stretch tensor also) and torsion tensor (interpreted as linearized wryness tensor also); in components ϵij=ui,j+𝙴kijϕk{{\upepsilon}}_{ij}={u}_{i,j}+{{\mathtt{E}}}_{kij}{\phi}_{k} and κij=ϕi,j.{{\upkappa}}_{ij}={\phi}_{i,j}. We consider the case of homogeneous, linear, anisotropic, micropolar elasticity [37]. The micropolar strain energy for a body of such a material is

𝒱(𝒖,ϕ)=Ω(ϵ,κ)𝑑v,\displaystyle{{\mathcal{V}}}({{{\boldsymbol{u}}}},{{\boldsymbol{\phi}}})=\int_{\Omega}{{\mathcal{E}}}({{\upepsilon}},{{\upkappa}})dv, (1.10)

with the energy density

(ϵ,κ)=12Aijklϵijϵkl+12Bijklκijκkl+Dijklϵijκkl.\displaystyle{{\mathcal{E}}}({{\upepsilon}},{{\upkappa}})=\frac{1}{2}{{A}}_{ijkl}{{\upepsilon}}_{ij}{{\upepsilon}}_{kl}+\frac{1}{2}{{B}}_{ijkl}{{\upkappa}}_{ij}{{\upkappa}}_{kl}+{{D}}_{ijkl}{{\upepsilon}}_{ij}{{\upkappa}}_{kl}. (1.11)

The needed constitutive relations for the stress tensor σij{{\upsigma}}_{ij} and the couple stress tensor μij{{\upmu}}_{ij}, supplementing the equations of equilibrium (1.4), are

σij=ϵij=Aijklϵkl+Dijklκkl,μij=κij=Bijklκkl+Dklijϵkl,\displaystyle{{\upsigma}}_{ij}=\frac{{\operatorname{\partial}}{{\mathcal{E}}}}{{\operatorname{\partial}}{{\upepsilon}}_{ij}}={{A}}_{ijkl}{{\upepsilon}}_{kl}+{{D}}_{ijkl}{{\upkappa}}_{kl},\quad{{\upmu}}_{ij}=\frac{{\operatorname{\partial}}{{\mathcal{E}}}}{{\operatorname{\partial}}{{\upkappa}}_{ij}}={{B}}_{ijkl}{{\upkappa}}_{kl}+{{D}}_{klij}{{\upepsilon}}_{kl}, (1.12)

where Aijkl{{A}}_{ijkl}, Dijkl{{D}}_{ijkl} and Bijkl{{B}}_{ijkl} are the (constant) elastic tensors of micropolar elasticity with the symmetries of 𝔸{{\mathbb{A}}} and 𝔹{{\mathbb{B}}} given by

Aijkl=Aklij,Bijkl=Bklij.\displaystyle{{A}}_{ijkl}={{A}}_{klij},\quad{{B}}_{ijkl}={{B}}_{klij}. (1.13)

It is easy to see that the Euler–Lagrange equations for the functional (1.10) are

Aijklxj(uk,l+𝙴mklϕm)+Dijklxjϕk,l=0,\displaystyle{{A}}_{ijkl}\frac{{\operatorname{\partial}}}{{\operatorname{\partial}}x_{j}}({u}_{k,l}+{{\mathtt{E}}}_{mkl}{\phi}_{m})+{{D}}_{ijkl}\frac{{\operatorname{\partial}}}{{\operatorname{\partial}}x_{j}}{\phi}_{k,l}=0, (1.14)

and

Bijklxjϕk,l+Dklijxj(uk,l+𝙴mklϕm)𝙴ijk(Ajkmn(um,n+𝙴pmnϕp)+Djkmnϕm,n)=0,\displaystyle{{B}}_{ijkl}\frac{{\operatorname{\partial}}}{{\operatorname{\partial}}x_{j}}{\phi}_{k,l}+{{D}}_{klij}\frac{{\operatorname{\partial}}}{{\operatorname{\partial}}x_{j}}({u}_{k,l}+{{\mathtt{E}}}_{mkl}{\phi}_{m})-{{\mathtt{E}}}_{ijk}({{A}}_{jkmn}({u}_{m,n}+{{\mathtt{E}}}_{pmn}{\phi}_{p})+{{D}}_{jkmn}{\phi}_{m,n})=0, (1.15)

which naturally coincide with the equations of equilibrium (1.4) in the absence of external forces and couples. Expanding further,

Aijkluk,lj+Dijklϕk,lj+Aijkl𝙴mklϕm,j=0,\displaystyle{{A}}_{ijkl}{u}_{k,lj}+{{D}}_{ijkl}{\phi}_{k,lj}+{{A}}_{ijkl}{{\mathtt{E}}}_{mkl}{\phi}_{m,j}=0, (1.16)

and

Bijklϕk,lj+Dklijuk,lj+Dklin𝙴mklϕm,n𝙴ijk(Ajkmn(um,n+𝙴pmnϕp)+Djkmnϕm,n)=0.\displaystyle{{B}}_{ijkl}{\phi}_{k,lj}+{{D}}_{klij}{u}_{k,lj}+{{D}}_{klin}{{\mathtt{E}}}_{mkl}{\phi}_{m,n}-{{\mathtt{E}}}_{ijk}({{A}}_{jkmn}({u}_{m,n}+{{\mathtt{E}}}_{pmn}{\phi}_{p})+{{D}}_{jkmn}{\phi}_{m,n})=0. (1.17)

We intend to look for null Lagrangian of the form of (1.11). For the conditions on such a null Lagrangian, we insist on the vanishing of the highest order derivative term (second order) in above equation. In other words, a set of sufficient conditions for the null Lagrangian are

Aijkl+Ailkj=0,Bijkl+Bilkj=0,Dijkl+Dilkj=0,\displaystyle{{A}}_{ijkl}+{{A}}_{ilkj}=0,\quad{{B}}_{ijkl}+{{B}}_{ilkj}=0,\quad{{D}}_{ijkl}+{{D}}_{ilkj}=0, (1.18)

and

𝙴mklDklin𝙴ijkDjkmn=0,\displaystyle{{\mathtt{E}}}_{mkl}{{D}}_{klin}-{{\mathtt{E}}}_{ijk}{{D}}_{jkmn}=0, (1.19)

as well as,

𝙴mklAijkl=0,𝙴ijkAjkmn=0,𝙴ijkAjkmn𝙴pmn=0.\displaystyle{{\mathtt{E}}}_{mkl}{{A}}_{ijkl}=0,\quad{{\mathtt{E}}}_{ijk}{{A}}_{jkmn}=0,\quad{{\mathtt{E}}}_{ijk}{{A}}_{jkmn}{{\mathtt{E}}}_{pmn}=0. (1.20)

In view of (1.13)1, it is clear that the last three conditions are implied by 𝙴mklAijkl=0{{\mathtt{E}}}_{mkl}{{A}}_{ijkl}=0. This implies, for klk\neq l,

AijklAijlk=0,\displaystyle{{A}}_{ijkl}-{{A}}_{ijlk}=0, (1.21)

however, due to (1.13)1, for iji\neq j: AijklAjikl=0{{A}}_{ijkl}-{{A}}_{jikl}=0 and Aijkl+Akjil=0{{A}}_{ijkl}+{{A}}_{kjil}=0. This leads to

Aijkl=Akjil=Ajkil=Aikjl.\displaystyle{{A}}_{ijkl}=-{{A}}_{kjil}=-{{A}}_{jkil}={{A}}_{ikjl}. (1.22)
Aijkl=Akjil=Ajkil=Ajikl=Aijkl\displaystyle\therefore\quad{{A}}_{ijkl}=-{{A}}_{kjil}=-{{A}}_{jkil}=-{{A}}_{jikl}=-{{A}}_{ijkl} (1.23)
Aijkl=0𝔸=0.\displaystyle\implies{{A}}_{ijkl}=0\implies{\mathbb{A}}=0. (1.24)

Concerning the conditions for 𝔻,\mathbb{D}, we make the following claim:

Claim 1.

The conditions (1.18)3 and (1.19) are equivalent to the following

Dijkl\displaystyle{{D}}_{ijkl} =Dilkj\displaystyle=-{{D}}_{ilkj} (1.25a)
Dijji\displaystyle{{D}}_{ijji} =Dikki for ijki\displaystyle=-{{D}}_{ikki}\text{ for }i\neq j\neq k\neq i (1.25b)
Dijjk\displaystyle{{D}}_{ijjk} =Dkikk+Djijk for ijki.\displaystyle={{D}}_{kikk}+{{D}}_{jijk}\text{ for }i\neq j\neq k\neq i. (1.25c)
Proof.

First assume that (1.18)3 and (1.19) hold. By (1.18)3,

Driki=0\displaystyle{{D}}_{riki}=0 (1.26)

Put (h,s,t)=π(1,2,3)(h,s,t)=\pi(1,2,3). Then (1.19) gives (no sum)

DthhtDhtht=DststDtsst\displaystyle{{D}}_{thht}-{{D}}_{htht}={{D}}_{stst}-{{D}}_{tsst}
Dthht=Dtsstby (1.26)\displaystyle\implies\;{{D}}_{thht}=-{{D}}_{tsst}\quad\text{by \eqref{E:pex}} (1.27)

Also if (h,s,t)=π(1,2,3)(h,s,t)=-\pi(1,2,3) then by same calculation (1.27) holds.
We now put s=ths=t\neq h, and assume that (s,h,k)=π(1,2,3)(s,h,k)=\pi(1,2,3). Then (1.19) gives (no sum)

DhkhsDkhhs=DksssDskss\displaystyle{{D}}_{hkhs}-{{D}}_{khhs}={{D}}_{ksss}-{{D}}_{skss}
Dkhhs=Dhkhs+Dskssby (1.26)\displaystyle\implies\;{{D}}_{khhs}={{D}}_{hkhs}+{{D}}_{skss}\quad\text{by \eqref{E:pex}} (1.28)

Also if (s,h,k)=π(1,2,3)(s,h,k)=-\pi(1,2,3) then by same calculation (1.28) holds. Also for h=tsh=t\neq s, by similar argument (1.28) holds.

For the proof in the other direction, we assume that (1.25) holds. To obtain (1.19), we fix h{1,2,3}h\in\{1,2,3\}. If k=hk=h then (1.19) trivially holds, so that we consider khk\neq h. Then

i,j=13Dijkleijh=(DijklDjikl)eijh(no sum)\sum_{i,j=1}^{3}{{D}}_{ijkl}e_{ijh}=\left({{D}}_{ijkl}-{{D}}_{jikl}\right)e_{ijh}\;\text{(no\;sum)}

where hh is fixed, ii and jj are also fixed considering non-zero contributions. Then six cases arises for kk and ll (kk has two values and ll has three). Let us divide these six possibilities in the following three cases:
Case 1. [hlkh\neq l\neq k]
This case consists of two sub-cases (h,k,l)=π(1,2,3)(h,k,l)=\pi(1,2,3) and (h,k,l)=π(1,2,3)(h,k,l)=-\pi(1,2,3). Then for nonzero term, {i,j}={k,l}\{i,j\}=\{k,l\}. Without any loss of generality, we assume that i=li=l. Therefore i,j=13Dijkleijh=Dijkleijh=Dlkklelkh\sum_{i,j=1}^{3}{{D}}_{ijkl}e_{ijh}={{D}}_{ijkl}e_{ijh}={{D}}_{lkkl}e_{lkh}. Similarly, i,j=13Dijhleijk=Dlhhlelhk.\sum_{i,j=1}^{3}{{D}}_{ijhl}e_{ijk}={{D}}_{lhhl}e_{lhk}. But Dlkklelkh=Dlhhlelkh=Dlhhlelhk,{{D}}_{lkkl}e_{lkh}=-{{D}}_{lhhl}e_{lkh}={{D}}_{lhhl}e_{lhk}, so that (1.19) holds.
Case 2. [k=lk=l]
This case consists of two sub-cases (p,l,h)=π(1,2,3)(p,l,h)=\pi(1,2,3) and (p,l,h)=π(1,2,3)(p,l,h)=-\pi(1,2,3), where pp is the only element of the set {1,2,3}\{h,k}\{1,2,3\}\backslash\{h,k\}. Then for nonzero contributions of the relevant term, either i=k=li=k=l or j=k=lj=k=l. Without any loss of generality, we assume that i=k=li=k=l. Therefore i,j=13Dijkleijh=Dijkleijh=Dljlleljh=(DjhhlDhjhl)eljh=(DhjhkDjhhk)ehjk\sum_{i,j=1}^{3}{{D}}_{ijkl}e_{ijh}={{D}}_{ijkl}e_{ijh}={{D}}_{ljll}e_{ljh}=\left({{D}}_{jhhl}-{{D}}_{hjhl}\right)e_{ljh}=\left({{D}}_{hjhk}-{{D}}_{jhhk}\right)e_{hjk}. Thus, for k=lk=l,

i,j=13Dijhleijk=(DhjhlDjhhl)ehjk(WLOG)=(DhjhkDjhhk)ehjk=i,j=13Dijkleijh.\sum_{i,j=1}^{3}{{D}}_{ijhl}e_{ijk}=\left({{D}}_{hjhl}-{{D}}_{jhhl}\right)e_{hjk}(WLOG)=\left({{D}}_{hjhk}-{{D}}_{jhhk}\right)e_{hjk}=\sum_{i,j=1}^{3}{{D}}_{ijkl}e_{ijh}.

Therefore, (1.19) holds.
Case 3. [h=l]h=l]
This case consists of two sub-cases (p,k,l)=π(1,2,3)(p,k,l)=\pi(1,2,3) and (p,k,l)=π(1,2,3)(p,k,l)=-\pi(1,2,3), where pp is the only element of the set {1,2,3}\{h,k}\{1,2,3\}\backslash\{h,k\}. Similar calculation as case 2 shows that (1.19) holds.

Remark 1.

According to above, it is noteworthy that (1.18)1, (1.18)2, and (1.25) are sufficient conditions on the fourth order tensors 𝔸{{\mathbb{A}}}, 𝔹{{\mathbb{B}}}, and 𝔻{{\mathbb{D}}} satisfying the symmetries (1.13) for an energy function to be a null Lagrangian in the linear micropolar elastostatics.

1.1 Centrosymmetric model

In the absence of chiral term involving 𝔻{\mathbb{D}}, for the general anisotropic case of the so called centrosymmetric model, the constitutive relations (1.12) become σij=Aijklϵkl,μij=Bijklκkl.{{\upsigma}}_{ij}={{A}}_{ijkl}{{\upepsilon}}_{kl},{{\upmu}}_{ij}={{B}}_{ijkl}{{\upkappa}}_{kl}. The suitable expression of the stored energy function for centrosymmetric micropolar material is

(ϵ,κ)=12ϵ𝔸[ϵ]+12κ𝔹[κ].{{\mathcal{E}}}({{\upepsilon}},{{\upkappa}})=\frac{1}{2}{{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]+\frac{1}{2}{{\upkappa}}{\cdot}{{\mathbb{B}}}[{{\upkappa}}]. (1.29)

Thus the total stored energy functional is

𝒱(𝒖,ϕ)=12Ω(ϵ𝔸[ϵ]+κ𝔹[κ])𝑑v.{{\mathcal{V}}}({{{\boldsymbol{u}}}},{{\boldsymbol{\phi}}})=\frac{1}{2}\int_{\Omega}({{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]+{{\upkappa}}{\cdot}{{\mathbb{B}}}[{{\upkappa}}])dv. (1.30)

By (1.9), the Lagrangian

Ψ(𝒖,ϕ)=12(ui,j+𝙴ijsϕs)Aijkl(uk,l+𝙴klrϕr)+12ϕi,jBijklϕk,l\displaystyle{\Psi}(\nabla{{{\boldsymbol{u}}}},\nabla{{\boldsymbol{\phi}}})=\frac{1}{2}({u}_{i,j}+{{\mathtt{E}}}_{ijs}{\phi}_{s}){{A}}_{ijkl}({u}_{k,l}+{{\mathtt{E}}}_{klr}{\phi}_{r})+\frac{1}{2}{\phi}_{i,j}{{B}}_{ijkl}{\phi}_{k,l} (1.31)

is a polynomial of degree 22. It is of interest to characterize the possible null Lagrangians within the family stated by H. Rund [17].

As a departure from the indicial notation used so far, we employ a different notation in the following. This closely follows the covariant/contravariant notation, but it should be remembered that we are not dealing with a curvilinear basis here; in fact we stick to the standard basis {𝒆1,𝒆2,𝒆3}\{{\boldsymbol{e}}_{1},{\boldsymbol{e}}_{2},{\boldsymbol{e}}_{3}\} of 3{{\mathbb{R}}}^{3} (and {𝒆1,𝒆2,,𝒆N}\{{\boldsymbol{e}}_{1},{\boldsymbol{e}}_{2},\dotsc,{\boldsymbol{e}}_{N}\} of N{{\mathbb{R}}}^{N}) and employ {𝒆1,𝒆2,𝒆3}\{{\boldsymbol{e}}^{1},{\boldsymbol{e}}^{2},{\boldsymbol{e}}^{3}\} as its own reciprocal copy (resp. {𝒆1,𝒆2,,𝒆N}\{{\boldsymbol{e}}^{1},{\boldsymbol{e}}^{2},\dotsc,{\boldsymbol{e}}^{N}\} of N{{\mathbb{R}}}^{N}). In general, the Greek indices range over 1,2,3,1,2,3, while the Latin indices range over 1,,N1,\dotsc,N unless specified otherwise. Accordingly, we denote the components of 𝒙3\boldsymbol{x}\in{{\mathbb{R}}}^{3} as xαx^{\alpha}. In view of the analysis of other models in the sequel, it is useful to consider 𝒚N\boldsymbol{y}\in{{\mathbb{R}}}^{N}; for instance, in the present section where we discuss linear micropolar media, we have N=6N=6 with the identification

𝒚=(𝒖,ϕ),D𝒚=(𝒖,ϕ).\displaystyle\boldsymbol{y}=({{\boldsymbol{u}}},{\boldsymbol{\phi}}),\quad D{\boldsymbol{y}}=(\nabla{{{\boldsymbol{u}}}},\nabla{{\boldsymbol{\phi}}}). (1.32)

The components of 𝒚\boldsymbol{y} are denoted by yiy^{i}, and the components of its derivative D𝒚N×3D{\boldsymbol{y}}\in{{\mathbb{R}}}^{N\times 3} as yαiy^{i}_{\alpha}. Using this notation, the Euler operator for an arbitary Lagrangian Ψ{\Psi} is found to be [17]

k(Ψ)=2Ψxγyγk+2Ψyjyγkyγj+2Ψyβjyγk2yjxβxγΨyk,{\mathscr{E}}_{k}({\Psi})=\frac{{\operatorname{\partial}}^{2}{\Psi}}{{\operatorname{\partial}}x^{{{\gamma}}}{\operatorname{\partial}}y^{k}_{{{\gamma}}}}+\frac{{\operatorname{\partial}}^{2}{\Psi}}{{\operatorname{\partial}}y^{j}{\operatorname{\partial}}y^{k}_{{{\gamma}}}}y^{j}_{{{\gamma}}}+\frac{{\operatorname{\partial}}^{2}{\Psi}}{{\operatorname{\partial}}y^{j}_{{{\beta}}}{\operatorname{\partial}}y^{k}_{{{\gamma}}}}\frac{{\operatorname{\partial}}^{2}y^{j}}{{\operatorname{\partial}}x^{{{\beta}}}{\operatorname{\partial}}x^{{{\gamma}}}}-\frac{{\operatorname{\partial}}{\Psi}}{{\operatorname{\partial}}y^{k}}, (1.33)

where k=1,,Nk=1,\dotsc,N. Here we recall that a Lagrangian is called null Lagrangian if and only if

k(Ψ)0.{\mathscr{E}}_{k}({\Psi})\equiv 0. (1.34)

The characterization theorem for null Lagrangians of the form of Ψ{\Psi} (:=Ψ(D𝒚):={\Psi}(D{\boldsymbol{y}})) is stated as Theorem 1 in Appendix A; In fact, Ψ{\Psi} is a null Lagrangian of polynomial degree 22 in D𝒚D{\boldsymbol{y}} if

Ψ=12𝔇(α1,α2;i1,i2)yα1i1yα2i2+𝔇(α1;i1)yα1i1+12𝔇(0;0),{\Psi}=\frac{1}{2}\mathfrak{D}(\alpha_{1},\alpha_{2};i_{1},i_{2})y^{i_{1}}_{\alpha_{1}}y^{i_{2}}_{\alpha_{2}}+\mathfrak{D}(\alpha_{1};i_{1})y^{i_{1}}_{\alpha_{1}}+\frac{1}{2}\mathfrak{D}(0;0), (1.35)

where Sα(𝒙,𝒚)S^{\alpha}(\boldsymbol{x},\boldsymbol{y}) are 𝒞2\mathcal{C}^{2} functions, while

𝔇(α1,α2;i1i2)\displaystyle\mathfrak{D}(\alpha_{1},\alpha_{2};i_{1}i_{2}) :=|Si1α1Si1α2Si2α1Si2α2|,𝔇(α1;i1)\displaystyle:=\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix},\quad\mathfrak{D}(\alpha_{1};i_{1}) :=|Si1α1Si1α2S|α2α1S|α2α2|,𝔇(0;0)\displaystyle:=\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix},\quad\mathfrak{D}(0;0) :=|S|α1α1S|α1α2S|α2α1S|α2α2|.\displaystyle:=\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{1}}&S^{\alpha_{2}}_{|\alpha_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix}. (1.36)

For the purpose of comparison with (1.31), it is useful to consider above in terms of splitting of 𝒚\boldsymbol{y} in 𝒖{{\boldsymbol{u}}} and ϕ{\boldsymbol{\phi}}. Thus, (1.35) can be expressed as

Ψ(𝒖,ϕ)\displaystyle{\Psi}(\nabla{{{\boldsymbol{u}}}},\nabla{{\boldsymbol{\phi}}}) =12𝔇1(α1,α2;i1,i2)ui1,α1ui2,α2+12𝔇2(α1,α2;i1,i2)ϕi1,α1ϕi2,α2+\displaystyle=\frac{1}{2}\mathfrak{D}_{1}(\alpha_{1},\alpha_{2};i_{1},i_{2}){u}_{i_{1},\alpha_{1}}{u}_{i_{2},\alpha_{2}}+\frac{1}{2}\mathfrak{D}_{2}(\alpha_{1},\alpha_{2};i_{1},i_{2}){\phi}_{i_{1}^{\prime},\alpha_{1}}{\phi}_{i_{2}^{\prime},\alpha_{2}}+
12𝔇3(α1,α2;i1,i2)ui1,α1ϕi2,α2+𝔇1(α1;i1)ui1,α1+𝔇2(α1;i1)ϕi1,α1\displaystyle\frac{1}{2}\mathfrak{D}_{3}(\alpha_{1},\alpha_{2};i_{1},i_{2}){u}_{i_{1},\alpha_{1}}{\phi}_{i_{2}^{\prime},\alpha_{2}}+\mathfrak{D}_{1}(\alpha_{1};i_{1}){u}_{i_{1},\alpha_{1}}+\mathfrak{D}_{2}(\alpha_{1};i_{1}){\phi}_{i_{1}^{\prime},\alpha_{1}}
+12𝔇(0;0)\displaystyle+\frac{1}{2}\mathfrak{D}(0;0) (1.37)

where 𝔇1\mathfrak{D}_{1}’s are restriction of corresponding 𝔇\mathfrak{D}’s (1.36) for 1i1,i231\leq i_{1},i_{2}\leq 3, 𝔇2\mathfrak{D}_{2}’s for 4i1,i264\leq i_{1},i_{2}\leq 6, 𝔇3\mathfrak{D}_{3} for 1i13, 4i261\leq i_{1}\leq 3,\,4\leq i_{2}\leq 6, while the primed indices are given by

i1=i13,i2=i23.i_{1}^{\prime}=i_{1}-3,\,i_{2}^{\prime}=i_{2}-3.

The representation for 𝔇3\mathfrak{D}_{3} in (1.37) is valid since 𝔇3(α1,α2;i1,i2)=𝔇3(α2,α1;i2,i1).\mathfrak{D}_{3}(\alpha_{1},\alpha_{2};i_{1},i_{2})=\mathfrak{D}_{3}(\alpha_{2},\alpha_{1};i_{2},i_{1}).

With the expanded form (1.37) of the expression of stored energy (1.35), it is easier to compare it with the form prescribed by (1.31). Thus, (1.31) is a null Lagrangian provided there exist functions SαS^{\alpha} such that

|Si1α1Si1α2Si2α1Si2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix} =𝔇1(α1,α2;i1,i2)=Ai1α1i2α2,\displaystyle=\mathfrak{D}_{1}(\alpha_{1},\alpha_{2};i_{1},i_{2})={{A}}_{i_{1}\alpha_{1}i_{2}\alpha_{2}},\quad |Si1α1Si1α2Si2α1Si2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix} =𝔇2(α1,α2;i1,i2)=Bi1α1i2α2,\displaystyle=\mathfrak{D}_{2}(\alpha_{1},\alpha_{2};i_{1},i_{2})={{B}}_{i_{1}^{\prime}\alpha_{1}i_{2}^{\prime}\alpha_{2}}, (1.38)
|Si1α1Si1α2Si2α1Si2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix} =𝔇3(α1,α2;i1,i2)=0,\displaystyle=\mathfrak{D}_{3}(\alpha_{1},\alpha_{2};i_{1},i_{2})=0,\quad |Si1α1Si1α2S|α2α1S|α2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix} =𝔇1(α1;i1)=eijsϕsAiji1α1,\displaystyle=\mathfrak{D}_{1}(\alpha_{1};i_{1})=e_{ijs}{\phi}_{s}{{A}}_{iji_{1}\alpha_{1}}, (1.39)
|Si1α1Si1α2S|α2α1S|α2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix} =𝔇2(α1;i1)=0,\displaystyle=\mathfrak{D}_{2}(\alpha_{1};i_{1})=0,\quad |S|α1α1S|α1α2S|α2α1S|α2α2|\displaystyle\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{1}}&S^{\alpha_{2}}_{|\alpha_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix} =𝔇(0;0)=eijseklrϕsϕrAijkl.\displaystyle=\mathfrak{D}(0;0)=e_{ijs}e_{klr}{\phi}_{s}{\phi}_{r}{{A}}_{ijkl}. (1.40)

We obtain further explicit form of necessary conditions on 𝔸{{\mathbb{A}}} and 𝔹{\mathbb{B}} from these conditions (1.38)–(1.40). From (1.38), using the properties of the determinant, we get the following restrictions for 𝔸{\mathbb{A}} and 𝔹{\mathbb{B}}:

Aijkl=Aklij,\displaystyle{{A}}_{ijkl}={{A}}_{klij},\quad Bijkl=Bklij,\displaystyle{{B}}_{ijkl}={{B}}_{klij}, (1.41)
Ailkj=Aijkl,\displaystyle{{A}}_{ilkj}=-{{A}}_{ijkl},\quad Bilkj=Bijkl,\displaystyle{{B}}_{ilkj}=-{{B}}_{ijkl}, (1.42)
Akjil=Aijkl,\displaystyle{{A}}_{kjil}=-{{A}}_{ijkl},\quad Bkjil=Bijkl,\displaystyle{{B}}_{kjil}=-{{B}}_{ijkl}, (1.43)
Aijkl=0,\displaystyle{{A}}_{ijkl}=0,\quad Bijkl=0,\displaystyle{{B}}_{ijkl}=0, if i=k or j=l.\displaystyle\text{if $i=k$ or $j=l$}. (1.44)

In order to proceed towards more simplification, let us introduce an alternate working notation for the functions of the form Si1α1S^{\alpha_{1}}_{i_{1}} and bring in more explicit form of the partial derivative involved. In view of the partition in the range of index i1i_{1}, etc, for clearer exposition, it is convenient to denote

Sαui as Suiα and Sαϕi as Sϕiα,\frac{\partial S^{\alpha}}{\partial{u}_{i}}\text{ as }S^{\alpha}_{{u}_{i}}\text{ and }\frac{\partial S^{\alpha}}{\partial{\phi}_{i}}\text{ as }S^{\alpha}_{{\phi}_{i}},

where α\alpha and ii both range over 1,2,3.1,2,3. Let us consider the following cases:
Case 1. Let Suiα0Sϕiαα,i{1,2,3}.S^{\alpha}_{{u}_{i}}\neq 0\neq S^{\alpha}_{{\phi}_{i}}\;\forall\alpha,i\in\{1,2,3\}. Then as (1.39)1 holds, i.e., |Si1α1Si1α2Si2α1Si2α2|=0|Suiα1Suiα2Sϕjα1Sϕjα2|=0\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix}=0\Leftrightarrow\begin{vmatrix}S^{\alpha_{1}}_{{u}_{i}}&S^{\alpha_{2}}_{{u}_{i}}\\ S^{\alpha_{1}}_{{\phi}_{j}}&S^{\alpha_{2}}_{{\phi}_{j}}\end{vmatrix}=0. Hence, cij\exists_{c_{ij}} such that 0|cij|<0\neq|c_{ij}|<\infty and

Suiα=cijSϕjα(no sum over j).S^{\alpha}_{{u}_{i}}=c_{ij}S^{\alpha}_{{\phi}_{j}}\quad\text{(no sum over $j$).}

This implies

Aiαjβ\displaystyle{{A}}_{i\alpha j\beta} =|SuiαSuiβSujαSujβ|=cjj|SuiαSuiβSϕjαSϕjβ|=cjj𝔇3(α,β;i,j)=0\displaystyle=\begin{vmatrix}S^{\alpha}_{{u}_{i}}&S^{\beta}_{{u}_{i}}\\ S^{\alpha}_{{u}_{j}}&S^{\beta}_{{u}_{j}}\end{vmatrix}=c_{jj}\begin{vmatrix}S^{\alpha}_{{u}_{i}}&S^{\beta}_{{u}_{i}}\\ S^{\alpha}_{{\phi}_{j}}&S^{\beta}_{{\phi}_{j}}\end{vmatrix}=c_{jj}\mathfrak{D}_{3}(\alpha,\beta;i,j)=0
and Biαjβ\displaystyle\text{and }{{B}}_{i\alpha j\beta} =|SϕiαSϕiβSϕjαSϕjβ|=1cii|SuiαSuiβSϕjαSϕjβ|=1cii𝔇3(α,β;i,j)=0\displaystyle=\begin{vmatrix}S^{\alpha}_{{\phi}_{i}}&S^{\beta}_{{\phi}_{i}}\\ S^{\alpha}_{{\phi}_{j}}&S^{\beta}_{{\phi}_{j}}\end{vmatrix}=\frac{1}{c_{ii}}\begin{vmatrix}S^{\alpha}_{{u}_{i}}&S^{\beta}_{{u}_{i}}\\ S^{\alpha}_{{\phi}_{j}}&S^{\beta}_{{\phi}_{j}}\end{vmatrix}=\frac{1}{c_{ii}}\mathfrak{D}_{3}(\alpha,\beta;i,j)=0

Therefore in this case no non-trivial null Lagrangian exists.
Case 2. Let for some α~,i~{1,2,3},Sϕi~α~=0\tilde{\alpha},\tilde{i}\in\{1,2,3\},\;S^{\tilde{\alpha}}_{{\phi}_{\tilde{i}}}=0. Then as (1.39)1 holds, it implies ai~,aj\exists_{a_{\tilde{i}},a_{j}} such that not both are zero, |ai~|,|aj|<|a_{\tilde{i}}|,|a_{j}|<\infty and

{ai~Sϕi~α~=ajSujα~ai~Sϕi~β=ajSujβ(no sum over i~,j).\begin{cases}a_{\tilde{i}}S^{\tilde{\alpha}}_{{\phi}_{\tilde{i}}}=a_{j}S^{\tilde{\alpha}}_{{u}_{j}}\\ a_{\tilde{i}}S^{\beta}_{{\phi}_{\tilde{i}}}=a_{j}S^{\beta}_{{u}_{j}}\end{cases}\quad\text{(no sum over ${\tilde{i}},j$).}

If ai~=0a_{\tilde{i}}=0, then aj0a_{j}\neq 0 which implies Sujα~=0=SujβS^{\tilde{\alpha}}_{{u}_{j}}=0=S^{\beta}_{{u}_{j}}. If ai~0a_{\tilde{i}}\neq 0 then aj=0a_{j}=0 implies Sϕi~β=0S^{\beta}_{{\phi}_{\tilde{i}}}=0 and Sujα~=0S^{\tilde{\alpha}}_{{u}_{j}}=0. Therefore, 𝔇3(α~,β;j,i~)=0Sujα~=0orSϕi~β=0\mathfrak{D}_{3}({\tilde{\alpha}},\beta;j,{\tilde{i}})=0\implies S^{\tilde{\alpha}}_{{u}_{j}}=0\;or\;S^{\beta}_{{\phi}_{\tilde{i}}}=0.

Consider the following subcases:

subcase 2.1. Let Sϕi~β0S^{\beta}_{{\phi}_{\tilde{i}}}\neq 0 for some βα~\beta\neq{\tilde{\alpha}}. Then 𝔇3(α~,β;j,i~)=0\mathfrak{D}_{3}({\tilde{\alpha}},\beta;j,{\tilde{i}})=0 implies Sujα~=0j{1,2,3}S^{\tilde{\alpha}}_{{u}_{j}}=0\;\forall j\in\{1,2,3\}. i.e., Sα~S^{\tilde{\alpha}} is independent of 𝒖{{\boldsymbol{u}}}. Therefore Akγlη=0{{A}}_{k\gamma l\eta}=0 if any one of γ,η\gamma,\eta equals α~{\tilde{\alpha}}.

subcase 2.2. Let Sujα~0S^{\tilde{\alpha}}_{{u}_{j}}\neq 0 for some α~,j{1,2,3}{\tilde{\alpha}},j\in\{1,2,3\}. Then 𝔇3(α~,β;j,i~)=0\mathfrak{D}_{3}({\tilde{\alpha}},\beta;j,{\tilde{i}})=0 implies Sϕi~β=0β{1,2,3}.S^{\beta}_{{\phi}_{\tilde{i}}}=0\;\forall\beta\in\{1,2,3\}. i.e., SβS^{\beta} is independent of ϕi~{\phi}_{\tilde{i}} for all β\beta. Therefore Bkγlη=0{{B}}_{k\gamma l\eta}=0 if any one of k,lk,l equals i~{\tilde{i}}.

Thus we conclude the following:

1. If all SαS^{\alpha} depends on all components of 𝒖{{\boldsymbol{u}}} and ϕi{\phi}_{i} then 𝔸=0=𝔹.{\mathbb{A}}=0={\mathbb{B}}.

2. If Sϕiα=0S^{\alpha}_{{\phi}_{i}}=0 for some α,i{1,2,3}{\alpha},i\in\{1,2,3\} then SαS^{\alpha} is independent of 𝒖{{\boldsymbol{u}}} or SβS^{\beta} is independent of ϕi{\phi}_{i} for all β\beta.

3. By interchanging the role of 𝒖{{\boldsymbol{u}}} and ϕ{\boldsymbol{\phi}} in case 2 thus far, if Suiα=0S^{\alpha}_{{u}_{i}}=0 for some α,i{1,2,3}{\alpha},i\in\{1,2,3\} then SαS^{\alpha} is independent of ϕ{\boldsymbol{\phi}} or SβS^{\beta} is independent of ui{u}_{i} for all β\beta.

Under subcase 2.1, consider two subcases as follows:

subcase 2.1.1. All other partial derivatives are non-zero. Then By first conclusion, Aiγjη=0{{A}}_{i\gamma j\eta}=0 whenever γα~η\gamma\neq{\tilde{\alpha}}\neq\eta. Therefore by the result in subcase 2.1, 𝔸=0{\mathbb{A}}=0.

subcase 2.1.2. Any one of other partial derivatives is zero. Then we can apply second or third conclusion. Thus applying the conclusions repeatedly, after a finite number of steps we find that either 𝔸=0=𝔹{\mathbb{A}}=0={\mathbb{B}} or SαS^{\alpha} is independent of 𝒖{{\boldsymbol{u}}} for all α{\alpha} or SαS^{\alpha} is independent of ϕ{\boldsymbol{\phi}} for all α{\alpha}. When all SαS^{\alpha}’s are independent of 𝒖{{\boldsymbol{u}}}, by (1.38)1, 𝔸=0{\mathbb{A}}=0. And when all Sα~S^{\tilde{\alpha}}’s are independent of ϕ{\boldsymbol{\phi}}, by (1.38), 𝔹=0{\mathbb{B}}=0. Our claim is in this case also 𝔸=0{\mathbb{A}}=0. To prove the claim, let us choose ϕ0{\boldsymbol{\phi}}\equiv 0. Then RHS (right hand side) of (1.39)2 becomes zero and hence LHS (left hand side). Since LHS determinants do not depend on ϕ{\boldsymbol{\phi}}, they are zero for all ϕ{\boldsymbol{\phi}}. Therefore we get

eijsϕsAijkl\displaystyle e_{ijs}{\phi}_{s}{{A}}_{ijkl} =0\displaystyle=0 (1.45)

We choose ϕ(𝒙)=(ϕ1(𝒙),ϕ2(𝒙),ϕ3(𝒙))=(1,0,0){\boldsymbol{\phi}}(\boldsymbol{x})=({\phi}_{1}(\boldsymbol{x}),{\phi}_{2}(\boldsymbol{x}),{\phi}_{3}(\boldsymbol{x}))=(1,0,0). Then from (1.45),

e321A32kl+e231A23kl=0\displaystyle e_{321}{{A}}_{32kl}+e_{231}{{A}}_{23kl}=0 1k,l3\displaystyle\quad 1\leq k,l\leq 3
A23kl=A32kl\displaystyle\implies{{A}}_{23kl}={{A}}_{32kl}

Similarly taking different ϕ{\boldsymbol{\phi}}’s we get Aijkl=Ajikl,{{A}}_{ijkl}={{A}}_{jikl}, and Aijkl=Akjil=Ajkil=Aikjl.{{A}}_{ijkl}=-{{A}}_{kjil}=-{{A}}_{jkil}={{A}}_{ikjl}. Therefore Aijkl=Akjil=Akijl=Aikjl=Aijkl,-{{A}}_{ijkl}={{A}}_{kjil}={{A}}_{kijl}={{A}}_{ikjl}={{A}}_{ijkl}, Aijkl=0,i.e.,𝔸=0.\implies{{A}}_{ijkl}=0,\text{i.e.,}{\mathbb{A}}=0.

Thus for all cases it is necessary that

𝔸=0.\displaystyle{\mathbb{A}}=0. (1.46)

Clearly, it is not necessary that 𝔹{\mathbb{B}} is 0. Thus the necessary conditions for null Lagrangian of the form (1.31), in centrosymmetric case (note that we assume Bklij=Bijkl{{B}}_{klij}={{B}}_{ijkl}) and within the family stated by H. Rund [17], are Bilkj=Bijkl,Bkjil=Bijkl,i,j,k,l,{{B}}_{ilkj}=-{{B}}_{ijkl},{{B}}_{kjil}=-{{B}}_{ijkl},\forall i,j,k,l, and Bijkl=0,{{B}}_{ijkl}=0, if i=ki=k or j=lj=l, and, most importantly, 𝔸=0.{{\mathbb{A}}}=0. It is emphasized that the last condition completely removes any contribution from the displacement field in a possible expression for a null Lagrangian.

Remark 2.

Although we briefly the same in the context of Claim 1.25, the sufficiency of (1.46) and (1.42)2, so that (1.31) is a null Lagrangian, can be easily verified by a direct calculation. Let

F\displaystyle F :=12ϕi,jBijklϕk,l.\displaystyle:=\frac{1}{2}{\phi}_{i,j}{{B}}_{ijkl}{\phi}_{k,l}.

Let i\mathscr{E}_{i} be the components of the Euler operator for i=1,,6.i=1,\dots,6. Then i(F)0\mathscr{E}_{i}(F)\equiv 0 for i=1,2,3i=1,2,3 since FF is independent of 𝐮{{\boldsymbol{u}}} and 𝐮\nabla{{\boldsymbol{u}}}. For i=4,5,6i=4,5,6,

i(F)=xj(Fϕi,j),where Fϕi,j=Bijklϕk,l.\mathscr{E}_{i}(F)=\frac{\partial}{\partial x_{j}}\left(F_{{\phi}_{i,j}}\right),\text{where }F_{{\phi}_{i,j}}={{B}}_{ijkl}{\phi}_{k,l}.

Therefore, by (1.42)2,

xj(Fϕi,j)=Bijklϕk,lj=12(Bijkl+Bilkj)ϕk,lj=0.\frac{\partial}{\partial x_{j}}\left(F_{{\phi}_{i,j}}\right)={{B}}_{ijkl}{\phi}_{k,lj}=\frac{1}{2}({{B}}_{ijkl}+{{B}}_{ilkj}){\phi}_{k,lj}=0.

As k(F)0\mathscr{E}_{k}(F)\equiv 0 for k=1,,6k=1,\dots,6, hence FF is a null Lagrangian.

Remark 3.

In contrast to the case of linear elasticity above null Lagrangian can be interpreted as stored energy functionals in the linear micropolar theory.

1.2 Splitting of stored energy using null terms

Let us decompose the fourth order elastic tensor 𝔹{{\mathbb{B}}} as

𝔹=𝔹^+𝔹~+𝔹̊{{\mathbb{B}}}=\hat{{\mathbb{B}}}+\tilde{{\mathbb{B}}}+\mathring{{\mathbb{B}}}
where B^ijkl:=14(Bijkl+Bklij+Bilkj+Bkjil),B~ijkl:=14(Bijkl+BklijBilkjBkjil),\displaystyle\text{where }\hat{{B}}_{ijkl}:=\frac{1}{4}({{B}}_{ijkl}+{{B}}_{klij}+{{B}}_{ilkj}+{{B}}_{kjil}),\quad\tilde{{B}}_{ijkl}:=\frac{1}{4}({{B}}_{ijkl}+{{B}}_{klij}-{{B}}_{ilkj}-{{B}}_{kjil}), (1.47)

and

B̊ijkl=12(BijklBklij).\displaystyle\mathring{{B}}_{ijkl}=\frac{1}{2}({{B}}_{ijkl}-{{B}}_{klij}). (1.48)

Then clearly 𝔹~\tilde{{\mathbb{B}}} satisfies (1.41)–(1.44). Due to (1.13)2, we have 𝔹̊=𝟘\mathring{{\mathbb{B}}}=\mathbb{0} (which are total 45 conditions). The remaining are 36 arbitrary constants in 𝔹^+𝔹~\hat{{\mathbb{B}}}+\tilde{{\mathbb{B}}} and are split between 𝔹^\hat{{\mathbb{B}}} and 𝔹~.\tilde{{\mathbb{B}}}. But 𝔹^\hat{{\mathbb{B}}} and 𝔹~\tilde{{\mathbb{B}}} individually satisfy the relation that swaps the indices (ij)(ij) and (kl)(kl). By symmetry, there are 9+18×2=459+18\times 2=45 zero entries in 𝔹~\tilde{{\mathbb{B}}}. Note that there are 1818 independent, possibly non-zero, entries in 𝔹~\tilde{{\mathbb{B}}}. With this decomposition, the energy functional (1.30) is split as follows:

𝒱\displaystyle{{\mathcal{V}}} =𝒱^+𝒱~,𝒱^(𝒖,ϕ):=12Ω(ϵ𝔸[ϵ]+κ𝔹^[κ])𝑑v,𝒱~(𝒖,ϕ):=12Ωκ𝔹~[κ]𝑑v,\displaystyle=\widehat{{{\mathcal{V}}}}+\widetilde{{{\mathcal{V}}}},\quad\widehat{{{\mathcal{V}}}}({{{\boldsymbol{u}}}},{{\boldsymbol{\phi}}}):=\frac{1}{2}\int_{\Omega}({{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]+{{\upkappa}}{\cdot}\hat{{\mathbb{B}}}[{{\upkappa}}])dv,\quad\widetilde{{{\mathcal{V}}}}({{{\boldsymbol{u}}}},{{\boldsymbol{\phi}}}):=\frac{1}{2}\int_{\Omega}{{\upkappa}}{\cdot}\tilde{{\mathbb{B}}}[{{\upkappa}}]dv, (1.49)

where 𝒱^\widehat{{{\mathcal{V}}}} cannot be a null functional (i.e. with its Lagrangian as null) unless 𝔸=𝟘=𝔹^{\mathbb{A}}=\mathbb{0}=\hat{{\mathbb{B}}}. In this context, it is noted that the Lagrangian in (1.49)2 can be null only if ϵ𝔸[ϵ]=0=κ𝔹^[κ]{{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]=0={{\upkappa}}{\cdot}\hat{{\mathbb{B}}}[{{\upkappa}}], since if they are non-zero, Ailkj=Aijkl,B^ilkj=B^ijkl{{A}}_{ilkj}={{A}}_{ijkl},\hat{{B}}_{ilkj}=\hat{{B}}_{ijkl} whch violate (1.42). Following [36] (see also [41]), the part 𝔹~\tilde{{\mathbb{B}}} of the elasticity tensor affects equilibrium through boundary term only and the equilibrium equation on the boundary takes the form (recall (1.2))

[ϕ]=𝒎0 in Ω, where [ϕ]\displaystyle\mathcal{M}[{\boldsymbol{\phi}}]=\boldsymbol{m}_{0}\text{ in $\partial\Omega$,}\quad\text{ where }\mathcal{M}[{\boldsymbol{\phi}}] :=𝔹[ϕ]𝒏=𝔹^[ϕ]𝒏+𝒎~[ϕ],𝒎~[ϕ]:=𝔹~[ϕ]𝒏,\displaystyle:={{\mathbb{B}}}[\nabla{\boldsymbol{\phi}}]\boldsymbol{n}=\hat{{\mathbb{B}}}[\nabla{\boldsymbol{\phi}}]\boldsymbol{n}+\tilde{\boldsymbol{m}}[{\boldsymbol{\phi}}],\quad\tilde{\boldsymbol{m}}[{\boldsymbol{\phi}}]:=\tilde{{\mathbb{B}}}[\nabla{\boldsymbol{\phi}}]\boldsymbol{n},

where 𝒏\boldsymbol{n} is the outward unit normal on the boundary of the body Ω\partial\Omega while 𝒎0\boldsymbol{m}_{0} is the externally applied surface moment (per unit area). This leads to the existence of a surface potential [36, 41] (for moments) corresponding to 𝒎~[ϕ]\tilde{\boldsymbol{m}}[{\boldsymbol{\phi}}], that is,

~[ϕ]=12Ω𝔹~ijklϕi,jϕknl𝑑σ.\tilde{\mathcal{M}}[{\boldsymbol{\phi}}]=\frac{1}{2}\int_{\partial\Omega}\tilde{{\mathbb{B}}}_{ijkl}{\phi}_{i,j}{\phi}_{k}n_{l}d\sigma. (1.50)

In order that 𝔹~\tilde{{\mathbb{B}}} (that satisfies the symmetry (1.13)2) vanishes we require that B~ijkl=0;\tilde{{B}}_{ijkl}=0; In other words,

[B~2233B~1233B~1322B~2133B~1133B~2311B~3122B~3211B~1122]=0,[B~3113B~3123B~2132B~3213B~1221B~1231B~2312B~1321B~2332]=0,\begin{bmatrix}\tilde{{B}}_{2233}&\tilde{{B}}_{1233}&\tilde{{B}}_{1322}\\ \tilde{{B}}_{2133}&\tilde{{B}}_{1133}&\tilde{{B}}_{2311}\\ \tilde{{B}}_{3122}&\tilde{{B}}_{3211}&\tilde{{B}}_{1122}\\ \end{bmatrix}=0,\quad\begin{bmatrix}\tilde{{B}}_{3113}&\tilde{{B}}_{3123}&\tilde{{B}}_{2132}\\ \tilde{{B}}_{3213}&\tilde{{B}}_{1221}&\tilde{{B}}_{1231}\\ \tilde{{B}}_{2312}&\tilde{{B}}_{1321}&\tilde{{B}}_{2332}\\ \end{bmatrix}=0,

provide a set of 1818 additional symmetry relations. These can be interpreted as analogues to the six Cauchy’s relations in the theory of linear elasticity [34, 2, 36, 40, 47].

Remark 4.

In the above scenario 𝔹{\mathbb{B}} lacks any minor symmetries unless the space of rotation vectors is restricted to gradients of scalar functions (say, curl-free vector fields on simply connected domains so that ϕi,j=ϕj,i{\phi}_{i,j}={\phi}_{j,i}). This is reflected in the fact that the number of conditions is 18 rather than the classical case of six [34]. The restriction to curl-free vector fields can be written as that when ψ=0\psi=0 or χ=0\chi=0 in the Clebsch representation of ϕ{\boldsymbol{\phi}}.

1.3 Non-chiral isotropic model

In the case of an isotropic material, 𝔸{{\mathbb{A}}} and 𝔹{{\mathbb{B}}} take the form,

{Aijkl=λδijδkl+(μ+κ)δikδjl+μδilδjk=AklijBijkl=β1δijδkl+β2δikδjl+β3δilδjk=Bklij\begin{cases}{{A}}_{ijkl}={{\lambda}}\delta_{ij}\delta_{kl}+({{\mu}}+{{\kappa}})\delta_{ik}\delta_{jl}+{{\mu}}\delta_{il}\delta_{jk}~{}&={{A}}_{klij}\\ {{B}}_{ijkl}={{\upbeta_{1}}}\delta_{ij}\delta_{kl}+{{\upbeta_{2}}}\delta_{ik}\delta_{jl}+{{\upbeta_{3}}}\delta_{il}\delta_{jk}&={{B}}_{klij}\end{cases}

where μ,κ,λ,β1,β2,β3{{\mu}},{{\kappa}},{{\lambda}},{{\upbeta_{1}}},{{\upbeta_{2}}},{{\upbeta_{3}}} are constants. Therefore,

σij\displaystyle\quad{{\upsigma}}_{ij} =Aijklϵkl=(μ+κ)ϵij+μϵji+λδijϵll,\displaystyle={{A}}_{ijkl}{{\upepsilon}}_{kl}=({{\mu}}+{{\kappa}}){{\upepsilon}}_{ij}+{{\mu}}{{\upepsilon}}_{ji}+{{\lambda}}\delta_{ij}{{\upepsilon}}_{ll},
and μij\displaystyle\text{and }\quad{{\upmu}}_{ij} =Bijklκkl=β2κij+β3κji+β1δijκll\displaystyle={{B}}_{ijkl}{{\upkappa}}_{kl}={{\upbeta_{2}}}{{\upkappa}}_{ij}+{{\upbeta_{3}}}{{\upkappa}}_{ji}+{{\upbeta_{1}}}\delta_{ij}{{\upkappa}}_{ll}

Using the fact that

ϵ𝔸[ϵ]=(μ+κ)ϵϵ+μϵϵ+λ(trϵ)2{{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]=({{\mu}}+{{\kappa}}){{\upepsilon}}{\cdot}{{\upepsilon}}+{{\mu}}{{\upepsilon}}{\cdot}{{\upepsilon}}^{\top}+{{\lambda}}(\operatorname{tr}{{\upepsilon}})^{2} (1.51)

is positive, and similar condition for κ𝔹[κ]{\upkappa}\cdot{{\mathbb{B}}}[{\upkappa}], we get the well known [16, 20, 37] necessary and sufficient conditions for positive definite stored energy are

κ>0,2μ+κ>0,3λ+2μ+κ>0,3β1+β2+β3>0,β2+β3>0,β2β3>0.{{\kappa}}>0,2{{\mu}}+{{\kappa}}>0,3{{\lambda}}+2{{\mu}}+{{\kappa}}>0,\quad 3{{\upbeta_{1}}}+{{\upbeta_{2}}}+{{\upbeta_{3}}}>0,{{\upbeta_{2}}}+{{\upbeta_{3}}}>0,{{\upbeta_{2}}}-{{\upbeta_{3}}}>0. (1.52)

From the perspective of null Lagrangian, we already have the restriction

μ=κ=λ=0,{{\mu}}={{\kappa}}={{\lambda}}=0, (1.53)

as 𝔸{{\mathbb{A}}} needs to be 0 within the class of null Lagrangians due to H. Rund [17]. Let us now simplify the expression involving 𝔹{\mathbb{B}}. Here

B~ijkl\displaystyle\tilde{{B}}_{ijkl} =12(BijklBilkj)=12(β3β1)(δjkδilδijδkl).\displaystyle=\frac{1}{2}({{B}}_{ijkl}-{{B}}_{ilkj})=\frac{1}{2}({{\upbeta_{3}}}-{{\upbeta_{1}}})(\delta_{jk}\delta_{il}-\delta_{ij}\delta_{kl}). (1.54)

Therefore, 𝔹~[κ]=12(β3β1)[κ(trκ)𝐈]\tilde{{\mathbb{B}}}[{\upkappa}]=\frac{1}{2}({{\upbeta_{3}}}-{{\upbeta_{1}}})[{\upkappa}^{\top}-(\operatorname{tr}{\upkappa})\mathbf{I}], so that

κ𝔹~[κ]\displaystyle{\upkappa}\cdot\tilde{{\mathbb{B}}}[{\upkappa}] =12(β3β1)[κκ(trκ)2].\displaystyle=\frac{1}{2}({{\upbeta_{3}}}-{{\upbeta_{1}}})\left[{\upkappa}\cdot{\upkappa}^{\top}-(\operatorname{tr}{\upkappa})^{2}\right]. (1.55)

We denote the deviatoric part

devκ:=κ13(trκ)𝐈.\operatorname{dev}{{\upkappa}}:={{\upkappa}}-\frac{1}{3}(\operatorname{tr}{{\upkappa}})\mathbf{I}.

Then κ^κ^=(devκ^)2+13(trκ)2\hat{{{\upkappa}}}{\cdot}\hat{{{\upkappa}}}=(\operatorname{dev}\hat{{{\upkappa}}})^{2}+\frac{1}{3}(\operatorname{tr}{{\upkappa}})^{2} since trκ^=trκ\operatorname{tr}\hat{{{\upkappa}}}=\operatorname{tr}{{\upkappa}} and (devκ^)𝐈=0(\operatorname{dev}\hat{{{\upkappa}}}){\cdot}\mathbf{I}=0. Also

κ^κ~=0,κκ=(devκ^)2+13(trκ)2+κ~κ~, and κκ=(devκ^)2+13(trκ)2κ~κ~.\displaystyle\hat{{{\upkappa}}}{\cdot}\tilde{{{\upkappa}}}=0,\quad{{\upkappa}}{\cdot}{{\upkappa}}=(\operatorname{dev}\hat{{{\upkappa}}})^{2}+\frac{1}{3}(\operatorname{tr}{{\upkappa}})^{2}+\tilde{{{\upkappa}}}{\cdot}\tilde{{{\upkappa}}},\text{ and }\quad{{\upkappa}}{\cdot}{{\upkappa}}^{\top}=(\operatorname{dev}\hat{{{\upkappa}}})^{2}+\frac{1}{3}(\operatorname{tr}{{\upkappa}})^{2}-\tilde{{{\upkappa}}}{\cdot}\tilde{{{\upkappa}}}. (1.56)

Therefore, we get

κ𝔹[κ]\displaystyle{{\upkappa}}{\cdot}{{\mathbb{B}}}[{{\upkappa}}] =(2μ+κ)(devκ^)2+3λ+2μ+κ3(trκ)2+κκ~κ~,\displaystyle=(2{{\mu}}+{{\kappa}})(\operatorname{dev}\hat{{{\upkappa}}})^{2}+\frac{3{{\lambda}}+2{{\mu}}+{{\kappa}}}{3}(\operatorname{tr}{{\upkappa}})^{2}+{{\kappa}}\tilde{{{\upkappa}}}{\cdot}\tilde{{{\upkappa}}},

and

κ𝔹^[κ]\displaystyle{\upkappa}\cdot\hat{{\mathbb{B}}}[{\upkappa}] =2β2+β3+β12(devκ^)2+β2+2β3+2β13(trκ)2+2β2β3β12κ~κ~,\displaystyle=\frac{2{{\upbeta_{2}}}+{{\upbeta_{3}}}+{{\upbeta_{1}}}}{2}(\operatorname{dev}\hat{{\upkappa}})^{2}+\frac{{{\upbeta_{2}}}+2{{\upbeta_{3}}}+2{{\upbeta_{1}}}}{3}(\operatorname{tr}{\upkappa})^{2}+\frac{2{{\upbeta_{2}}}-{{\upbeta_{3}}}-{{\upbeta_{1}}}}{2}\tilde{{\upkappa}}\cdot\tilde{{\upkappa}}, (1.57)

by (1.56). For a null Lagrangian the non-trivial expression corresponds to κ𝔹[κ]{\upkappa}\cdot{{\mathbb{B}}}[{\upkappa}], while its symmetric part leads to κB^[κ]=0{\upkappa}\cdot\hat{{B}}[{\upkappa}]=0. Therefore, using (1.57), we have,

2β2+β3+β1=0,β2+2β3+2β1=0,2β2β3β1=0\displaystyle 2{{\upbeta_{2}}}+{{\upbeta_{3}}}+{{\upbeta_{1}}}=0,\quad{{\upbeta_{2}}}+2{{\upbeta_{3}}}+2{{\upbeta_{1}}}=0,\quad 2{{\upbeta_{2}}}-{{\upbeta_{3}}}-{{\upbeta_{1}}}=0

which gives

β2=0andβ2β3β1=0{{\upbeta_{2}}}=0\quad\text{and}\quad{{\upbeta_{2}}}-{{\upbeta_{3}}}-{{\upbeta_{1}}}=0 (1.58)

and thus

ϵ𝔸~[ϵ]+κ𝔹[κ]=κ𝔹~[κ]=(β2β3)[(trκ)2κκ].\boldsymbol{\epsilon}\cdot\tilde{\mathbb{A}}[\boldsymbol{\epsilon}]+{\upkappa}\cdot{{\mathbb{B}}}[{\upkappa}]={\upkappa}\cdot\tilde{{\mathbb{B}}}[{\upkappa}]=({{\upbeta_{2}}}-{{\upbeta_{3}}})\left[(\operatorname{tr}{\upkappa})^{2}-{\upkappa}\cdot{\upkappa}^{\top}\right].

Indeed, (1.58) and (1.53) are the necessary and sufficient conditions for null Lagrangian in isotropic case without chiral term 𝔻{\mathbb{D}} and within the class of null Lagrangians due to H. Rund [17]. And the Lagrangian becomes

Ψ(ϵ,κ)=C0[(trκ)2κκ]{\Psi}(\boldsymbol{\epsilon},{\upkappa})=C_{0}\left[(\operatorname{tr}{\upkappa})^{2}-{\upkappa}\cdot{\upkappa}^{\top}\right] (1.59)

for some constant C0C_{0}.

Remark 5.

Using the definition of the second invariant of a second order tensor, i.e., 2(𝐀)=12((tr𝐀)2tr𝐀2),\mathcal{I}_{2}(\mathbf{A})=\frac{1}{2}((\operatorname{tr}\mathbf{A})^{2}-\operatorname{tr}\mathbf{A}^{2}), we have an equivalent expression for (1.59), i.e., Ψ(ϵ,κ)=C02(κ).{{\Psi}}({{\upepsilon}},{{\upkappa}})=C_{0}\mathcal{I}_{2}({{\upkappa}}). It can be easily verified by direct calculation too that Ψ{{\Psi}} is a null Lagrangian.

1.4 Chiral isotropic model

Isotropy (which insists on a symmetry relative to all orthogonal tensors) implies centrosymmetry so the expressions discussed earlier may need to be modified with non-zero 𝔻{\mathbb{D}}. In non-centrosymmetric case of symmetry similar to ‘Isotropy’ (relative to all rotation tensors) does not include mirror-reflection, i.e., inversion, (sometimes called hemitropy) has additonal parameters. In the case of general chiral ‘isotropic’ micropolar elastic material, three additional material constants appear as compared to the non-chiral isotropic micropolar elastic material [26, 25, 39, 45, 43]. These additional material parameters allows the incorporation of a change in the signs of the corresponding terms depending on the handedness of the microstructure (owing to the presence of axial vector, rotation axis, cross product, etc). For hemitropic (chiral isotropic) material, 𝔸{{\mathbb{A}}}, 𝔹{{\mathbb{B}}}, 𝔻{{\mathbb{D}}} takes the form [21, 26, 42, 45]

{Aijkl=λδijδkl+(μ+κ)δikδjl+μδilδjk=AklijBijkl=β1δijδkl+β2δikδjl+β3δilδjk=BklijDijkl=ζδijδkl+(ν+ρ)δikδjl+νδilδjk=Dklij.\begin{cases}{{A}}_{ijkl}={{\lambda}}\delta_{ij}\delta_{kl}+({{\mu}}+{{\kappa}})\delta_{ik}\delta_{jl}+{{\mu}}\delta_{il}\delta_{jk}&={{A}}_{klij}\\ {{B}}_{ijkl}={{\upbeta_{1}}}\delta_{ij}\delta_{kl}+{{\upbeta_{2}}}\delta_{ik}\delta_{jl}+{{\upbeta_{3}}}\delta_{il}\delta_{jk}&={{B}}_{klij}\\ {{D}}_{ijkl}={{\zeta}}\delta_{ij}\delta_{kl}+({{\nu}}+{{\rho}})\delta_{ik}\delta_{jl}+{{\nu}}\delta_{il}\delta_{jk}&={{D}}_{klij}.\end{cases} (1.60)

In contrast to the classical isotropic linear elasticity which is characterized by two material constants, so called Láme, the non-centrosymmetric ‘isotropic’ micropolar material possesses nine independent material constants. The stress σ{{\upsigma}} and couple stress μ{{\upmu}} are given by (1.12). For chiral isotropic case, σij=λδijϵll+(μ+κ)ϵij+μϵji+ζδijκll+(ν+ρ)κij+νκji,μij=β1δijκll+β2κij+β3κji+ζδijϵll+(ν+ρ)ϵij+νϵji.{{\upsigma}}_{ij}={{\lambda}}\delta_{ij}{{\upepsilon}}_{ll}+({{\mu}}+{{\kappa}}){{\upepsilon}}_{ij}+{{\mu}}{{\upepsilon}}_{ji}+{{\zeta}}\delta_{ij}{{\upkappa}}_{ll}+({{\nu}}+{{\rho}}){{\upkappa}}_{ij}+{{\nu}}{{\upkappa}}_{ji},{{\upmu}}_{ij}={{\upbeta_{1}}}\delta_{ij}{{\upkappa}}_{ll}+{{\upbeta_{2}}}{{\upkappa}}_{ij}+{{\upbeta_{3}}}{{\upkappa}}_{ji}+{{\zeta}}\delta_{ij}{{\upepsilon}}_{ll}+({{\nu}}+{{\rho}}){{\upepsilon}}_{ij}+{{\nu}}{{\upepsilon}}_{ji}. For positive definite stored energy,

κ>0, 2μ+κ>0, 3λ+2μ+κ>0,\displaystyle{{\kappa}}>0,\;2{{\mu}}+{{\kappa}}>0,\;3{{\lambda}}+2{{\mu}}+{{\kappa}}>0, (1.61)
3β1+β3+β2>0,β3+β2>0,β2β3>0.\displaystyle 3{{\upbeta_{1}}}+{{\upbeta_{3}}}+{{\upbeta_{2}}}>0,\;{{\upbeta_{3}}}+{{\upbeta_{2}}}>0,\;{{\upbeta_{2}}}-{{\upbeta_{3}}}>0. (1.62)

According to (1.18)1 and (1.18)3, we require that ζδijδkl+(ν+ρ)δikδjl+νδilδjk=ζδkjδil(ν+ρ)δkiδjlνδklδji,{{\zeta}}\delta_{ij}\delta_{kl}+({{\nu}}+{{\rho}})\delta_{ik}\delta_{jl}+{{\nu}}\delta_{il}\delta_{jk}=-{{\zeta}}\delta_{kj}\delta_{il}-({{\nu}}+{{\rho}})\delta_{ki}\delta_{jl}-{{\nu}}\delta_{kl}\delta_{ji}, so that

ν+ζ=ν+ρ=0.\displaystyle{{\nu}}+{{\zeta}}={{\nu}}+{{\rho}}=0. (1.63)

Hence, (1.60)3 implies Dijkl=ζ(δijδklδilδjk),{{D}}_{ijkl}={{\zeta}}(\delta_{ij}\delta_{kl}-\delta_{il}\delta_{jk}), so that

ϵ𝔻[κ]=ζ[(trϵ)(trκ)ϵκ].\displaystyle{{\upepsilon}}{\cdot}{{\mathbb{D}}}[{{\upkappa}}]={{\zeta}}\left[(\operatorname{tr}{{\upepsilon}})(\operatorname{tr}{{\upkappa}})-{{\upepsilon}}{\cdot}{{\upkappa}}^{\top}\right]. (1.64)

Similarly, (1.60)1 and (1.60)2 implies

μ+λ=μ+κ=0, and β3+β1=β2=0,\displaystyle{{\mu}}+{{\lambda}}={{\mu}}+{{\kappa}}=0,\text{ and }{{\upbeta_{3}}}+{{\upbeta_{1}}}={{\upbeta_{2}}}=0, (1.65)

so that Aijkl=λ(δijδklδilδjk),Bijkl=β1(δijδklδilδjk),{{A}}_{ijkl}={{\lambda}}(\delta_{ij}\delta_{kl}-\delta_{il}\delta_{jk}),{{B}}_{ijkl}={{\upbeta_{1}}}(\delta_{ij}\delta_{kl}-\delta_{il}\delta_{jk}), and

ϵ𝔸[ϵ]=λ[(trϵ)2ϵϵ],κ𝔹[κ]=β1[(trκ)2κκ].\displaystyle{{\upepsilon}}{\cdot}{{\mathbb{A}}}[{{\upepsilon}}]={\lambda}\left[(\operatorname{tr}{{\upepsilon}})^{2}-{{\upepsilon}}{\cdot}{{\upepsilon}}^{\top}\right],\quad{{\upkappa}}{\cdot}{{\mathbb{B}}}[{{\upkappa}}]={\upbeta_{1}}\left[(\operatorname{tr}{{\upkappa}})^{2}-{{\upkappa}}{\cdot}{{\upkappa}}^{\top}\right]. (1.66)

Also,

σij=λϵji+λδijϵllζκji+ζδijκll,\displaystyle{{\upsigma}}_{ij}=-{{\lambda}}{{\upepsilon}}_{ji}+{{\lambda}}\delta_{ij}{{\upepsilon}}_{ll}-{{\zeta}}{{\upkappa}}_{ji}+{{\zeta}}\delta_{ij}{{\upkappa}}_{ll},
μij=β1κji+β1δijκllζϵji+ζδijϵll.\displaystyle{{\upmu}}_{ij}=-{{\upbeta_{1}}}{{\upkappa}}_{ji}+{{\upbeta_{1}}}\delta_{ij}{{\upkappa}}_{ll}-{{\zeta}}{{\upepsilon}}_{ji}+{{\zeta}}\delta_{ij}{{\upepsilon}}_{ll}.

Let

F\displaystyle F :=12(ui,j+𝙴ijsϕs)Aijkl(uk,l+𝙴klsϕs)+12ϕi,jBijklϕk,l+(ui,j+𝙴ijsϕs)Dijklϕk,l.\displaystyle:=\frac{1}{2}({u}_{i,j}+{{\mathtt{E}}}_{ijs}{\phi}_{s}){{A}}_{ijkl}({u}_{k,l}+{{\mathtt{E}}}_{kls}{\phi}_{s})+\frac{1}{2}{\phi}_{i,j}{{B}}_{ijkl}{\phi}_{k,l}+({u}_{i,j}+{{\mathtt{E}}}_{ijs}{\phi}_{s}){{D}}_{ijkl}{\phi}_{k,l}.

Let i\mathscr{E}_{i} be the components of the Euler operator for i=1,,6.i=1,\dots,6. Then for i=1,2,3i=1,2,3, i(F)=0\mathscr{E}_{i}(F)=0 gives

σij,j=λϵji,j+λϵll,iζκji,j+ζκll,i=0,{{\upsigma}}_{ij,j}=-{{\lambda}}{{\upepsilon}}_{ji,j}+{{\lambda}}{{\upepsilon}}_{ll,i}-{{\zeta}}{{\upkappa}}_{ji,j}+{{\zeta}}{{\upkappa}}_{ll,i}=0,

and for i=4,5,6i=4,5,6, i(F)=0\mathscr{E}_{i}(F)=0 gives

μij,j𝙴ijkσjk=β1κji,j+β1κll,iζϵji,j+ζϵll,i=0,{{\upmu}}_{ij,j}-{{\mathtt{E}}}_{ijk}{{\upsigma}}_{jk}=-{{\upbeta_{1}}}{{\upkappa}}_{ji,j}+{{\upbeta_{1}}}{{\upkappa}}_{ll,i}-{{\zeta}}{{\upepsilon}}_{ji,j}+{{\zeta}}{{\upepsilon}}_{ll,i}=0,

where ϵll=ul,l{{\upepsilon}}_{ll}={u}_{l,l} and κll=ϕl,l{{\upkappa}}_{ll}={\phi}_{l,l}. In fact,

λ(uj,ij+𝙴jisϕs,j)+λul,liζϕj,ij+ζϕl,li=0,β1ϕj,ij+β1ϕl,liζ(uj,ij+𝙴jisϕs,j)+ζul,li=0,-{{\lambda}}({u}_{j,ij}+{{\mathtt{E}}}_{jis}{\phi}_{s,j})+{{\lambda}}{u}_{l,li}-{{\zeta}}{\phi}_{j,ij}+{{\zeta}}{\phi}_{l,li}=0,\quad-{{\upbeta_{1}}}{\phi}_{j,ij}+{{\upbeta_{1}}}{\phi}_{l,li}-{{\zeta}}({u}_{j,ij}+{{\mathtt{E}}}_{jis}{\phi}_{s,j})+{{\zeta}}{u}_{l,li}=0,

i.e.,

λ𝙴jisϕs,j=0,ζ𝙴jisϕs,j=0.{{\lambda}}{{\mathtt{E}}}_{jis}{\phi}_{s,j}=0,\quad{{\zeta}}{{\mathtt{E}}}_{jis}{\phi}_{s,j}=0.

As k(F)0\mathscr{E}_{k}(F)\equiv 0 for k=1,,6k=1,\dots,6, hence FF is a null Lagrangian if and only if λ=ζ=0{\lambda}={\zeta}=0. Thus, we do not find any other null Lagrangians besides those which are isotropic.

Remark 6.

On the other hand if the space of rotation vector field is restricted to curl free vector fields then the null Lagrangians with non-zero values of λ{\lambda} and ζ{\zeta} are also admissible.

2 Generalized elasticity for quasicrystals

We seek the null Lagrangians in the framework of the generalized elasticity theory of quasicrystals [19, 31]. In general, an (n3)(n-3)-dimensional quasicrystal can be generated by the projection of an nn-dimensional periodic structure to the 3-dimensional physical space (n=4,5,6n=4,5,6). Typically, the nn-dimensional hyperspace EnE^{n} is seen as the direct sum of two orthogonal subspaces (phonons and phasons, respectively),

En=Ep3Es(n3).\displaystyle E^{n}=E_{{p}}^{3}\oplus E_{{s}}^{(n-3)}. (2.1)

Let us restrict ourselves for illustrative purpose the case when n=6n=6. Throughout this section, we denote the phonon fields by ()p(\cdot)^{{p}} and the phason fields by ()s(\cdot)^{{s}}. It is important to note that all quantities (phonon and phason fields) depend on the referential (material) location 𝒙3{\boldsymbol{x}}\in{{\mathbb{R}}}^{3}. In the theory of quasicrystals, the equilibrium conditions are of the form (see, e.g., [33, 38])

σij,jp+fip=0,σij,js+fis=0,\displaystyle{{\sigma}}^{{p}}_{ij,j}+f_{i}^{{p}}=0,\quad{{\sigma}}^{{s}}_{ij,j}+f_{i}^{{s}}=0, (2.2)

where σijp{{\sigma}}^{{p}}_{ij} and σijs{{\sigma}}^{{s}}_{ij} are the phonon and phason stress tensors, respectively, and fipf_{i}^{{p}} is the conventional (phonon) body force density and fisf_{i}^{{s}} is a generalized (phason) body force density. The comma denotes differentiation with respect to the material coordinates. We note that the phonon stress tensor is symmetric, σijp=σjip{{\sigma}}^{{p}}_{ij}={{\sigma}}^{{p}}_{ji}, while the phason stress tensor is allowed to be asymmetric, σijsσjis{{\sigma}}^{{s}}_{ij}\neq{{\sigma}}^{{s}}_{ji} (see, e.g., [33]). The phonon and phason distortion tensors, γkl{{\upgamma}}_{kl} and κkl{{\upkappa}}_{kl}, are defined as the spatial gradients of ukp{u}_{k}^{{p}} and uks{u}_{k}^{{s}}, respectively

γkl=𝐮k,l,κkl=𝐯k,l.\displaystyle{{\upgamma}}_{kl}={\operatorname{\boldsymbol{u}}}_{k,l},\quad{{\upkappa}}_{kl}={\operatorname{\boldsymbol{v}}}_{k,l}. (2.3)

The constitutive relations between the stress tensors and the distortion tensors are

σijp=Cijklγkl+Dijklκkl,σijs=Dklijγkl+Eijklκkl.\displaystyle{{\sigma}}^{{p}}_{ij}={C}_{ijkl}{{\upgamma}}_{kl}+{D}_{ijkl}{{\upkappa}}_{kl},\quad{{\sigma}}^{{s}}_{ij}={D}_{klij}{{\upgamma}}_{kl}+{E}_{ijkl}{{\upkappa}}_{kl}. (2.4)

In the above backdrop, we investigate the null Lagrangians of the following type.
With 𝐮:Ω3\operatorname{\boldsymbol{u}}:\Omega\to\mathbb{R}^{3} and 𝐯:Ω3\operatorname{\boldsymbol{v}}:\Omega\to\mathbb{R}^{3}, the Lagrangian is given by

Ψ(𝐮,𝐯)=(γ,κ)=12γ[γ]+12κ𝔼[κ]+γ𝔻[κ],\displaystyle{\Psi}({\nabla}\operatorname{\boldsymbol{u}},{\nabla}\operatorname{\boldsymbol{v}})={{\mathcal{E}}}({{\upgamma}},{{\upkappa}})=\frac{1}{2}{{\upgamma}}{\cdot}{\mathbb{C}}[{{\upgamma}}]+\frac{1}{2}{{\upkappa}}{\cdot}{\mathbb{E}}[{{\upkappa}}]+{{\upgamma}}{\cdot}{{\mathbb{D}}}[{{\upkappa}}], (2.5)
where γ=𝐮,κ=𝐯,\text{where }{\upgamma}={\nabla}\operatorname{\boldsymbol{u}},{\upkappa}={\nabla}\operatorname{\boldsymbol{v}},

and the three constitutive tensors possess the symmetries

Cijkl=Cklij=Cijlk=Cjikl,Dijkl=Djikl,Eijkl=Eklij.\displaystyle{C}_{ijkl}={C}_{klij}={C}_{ijlk}={C}_{jikl},\quad{D}_{ijkl}={D}_{jikl},\quad{E}_{ijkl}={E}_{klij}. (2.6)

Here, Cijkl{C}_{ijkl} is the tensor of the elastic moduli of phonons, Eijkl{E}_{ijkl} is the tensor of the elastic moduli of phasons, and Dijkl{D}_{ijkl} is the tensor of the elastic moduli of the phonon-phason coupling. The symmetries of the tensors of the elastic constants can be simplified according to the specific type of the considered quasicrystal (see e.g. [38, 46]). After we substitute (2.4) and (2.3) into (2.2), we obtain

Cijkluk,ljp+Dijkluk,ljs=fip,\displaystyle{C}_{ijkl}{u}_{k,lj}^{{p}}+{D}_{ijkl}{u}_{k,lj}^{{s}}=-f_{i}^{{p}}, (2.7)
Dklijuk,ljp+Eijkluk,ljs=fis.\displaystyle{D}_{klij}{u}_{k,lj}^{{p}}+{E}_{ijkl}{u}_{k,lj}^{{s}}=-f_{i}^{{s}}. (2.8)

In the context of the Euler–Lagrange equation (1.33), here the dependent variable is 𝒚=(𝒖,𝒗)\boldsymbol{y}=({{\boldsymbol{u}}},\boldsymbol{v}), where 𝒖{{\boldsymbol{u}}} and 𝒗\boldsymbol{v} both have dimension 33, so that N=6N=6 here. Rewriting (1.35) for 𝒖{{\boldsymbol{u}}} and 𝒗\boldsymbol{v} we get,

Ψ(𝒙,𝒖,𝒗,𝒖,𝒗)=12𝒟1(α,β;i,j)ui,αuj,β+12𝒟2(α,β;i,j)vi,αvj,β+𝒟3(α,β;i,j)ui,αvj,β+𝒟1(α;i)ui,α+𝒟2(α;i)vi,α+12𝒟(0;0){\Psi}(\boldsymbol{x},{{\boldsymbol{u}}},\boldsymbol{v},\nabla{{{\boldsymbol{u}}}},\nabla{\boldsymbol{v}})=\frac{1}{2}\mathscr{D}_{1}(\alpha,\beta;i,j){u}_{i,\alpha}{u}_{j,\beta}+\frac{1}{2}\mathscr{D}_{2}(\alpha,\beta;i^{\prime},j^{\prime})v_{i,\alpha}v_{j,\beta}\\ +\mathscr{D}_{3}(\alpha,\beta;i,j^{\prime}){u}_{i,\alpha}v_{j,\beta}+\mathscr{D}_{1}(\alpha;i){u}_{i,\alpha}+\mathscr{D}_{2}(\alpha;i^{\prime})v_{i,\alpha}+\frac{1}{2}\mathscr{D}(0;0) (2.9)

where 1i,j31\leq i,j\leq 3 and i=i+3,j=j+3i^{\prime}=i+3,\,j^{\prime}=j+3.
Also rewriting given Lagrangian in terms of ui,α=𝒖{u}_{i,\alpha}=\nabla{{\boldsymbol{u}}} and vi,α=𝒗v_{i,\alpha}=\nabla\boldsymbol{v} we get,

Ψ(𝒙,𝒖,𝒗,𝒖,𝒗)=12Ciαjβui,αuj,β+12Eiαjβvi,αvj,β+Diαjβui,αvj,β.{\Psi}(\boldsymbol{x},{{\boldsymbol{u}}},\boldsymbol{v},\nabla{{\boldsymbol{u}}},\nabla\boldsymbol{v})=\frac{1}{2}{C}_{i\alpha j\beta}{u}_{i,\alpha}{u}_{j,\beta}+\frac{1}{2}{E}_{i\alpha j\beta}v_{i,\alpha}v_{j,\beta}+{D}_{i\alpha j\beta}{u}_{i,\alpha}v_{j,\beta}. (2.10)

By comparing (2.10) and (2.9) we get,

|SiαSiβSjαSjβ|\displaystyle\begin{vmatrix}S^{\alpha}_{i}&S^{\beta}_{i}\\ S^{\alpha}_{j}&S^{\beta}_{j}\end{vmatrix} =𝒟1(α,β;i,j)=Ciαjβ,\displaystyle=\mathscr{D}_{1}(\alpha,\beta;i,j)={C}_{i\alpha j\beta},\quad |SiαSiβSjαSjβ|\displaystyle\begin{vmatrix}S^{\alpha}_{i^{\prime}}&S^{\beta}_{i^{\prime}}\\ S^{\alpha}_{j^{\prime}}&S^{\beta}_{j^{\prime}}\end{vmatrix} =𝒟2(α,β;i,j)=Eiαjβ,\displaystyle=\mathscr{D}_{2}(\alpha,\beta;i^{\prime},j^{\prime})={E}_{i\alpha j\beta}, (2.11)
|SiαSiβSjαSjβ|\displaystyle\begin{vmatrix}S^{\alpha}_{i}&S^{\beta}_{i}\\ S^{\alpha}_{j^{\prime}}&S^{\beta}_{j^{\prime}}\end{vmatrix} =𝒟3(α,β;i,j)=Diαjβ,\displaystyle=\mathscr{D}_{3}(\alpha,\beta;i,j^{\prime})={D}_{i\alpha j\beta},\quad |SiαSiβS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{i}&S^{\beta}_{i}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟1(α;i)=0,\displaystyle=\mathscr{D}_{1}(\alpha;i)=0, (2.12)
|SiαSiβS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{i^{\prime}}&S^{\beta}_{i^{\prime}}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟2(α;i)=0,\displaystyle=\mathscr{D}_{2}(\alpha;i^{\prime})=0,\quad |S|ααS|αβS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{|\alpha}&S^{\beta}_{|\alpha}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟(0;0)=0.\displaystyle=\mathscr{D}(0;0)=0. (2.13)

These are necessery and sufficient conditions on ,𝔼\mathbb{C},{\mathbb{E}} and 𝔻{\mathbb{D}} so that (2.10) becomes null Lagrangian within the family stated by H. Rund [17]. In the following we further explore possible explicit conditions on ,𝔼{\mathbb{C}},{\mathbb{E}} and 𝔻{\mathbb{D}}.

From (2.11) and (2.12)1 we get the following restrictions on ,𝔼{\mathbb{C}},{\mathbb{E}} and 𝔻{\mathbb{D}},

Cijkl=Cklij,Eijkl=Eklij,Dijkl=Dklij\displaystyle{C}_{ijkl}={C}_{klij},\;{E}_{ijkl}={E}_{klij},\;{D}_{ijkl}={D}_{klij} (2.14)
Cijkl=Ckjil,Eijkl=Ekjil,Dijkl=Dkjil\displaystyle{C}_{ijkl}=-{C}_{kjil},\;{E}_{ijkl}=-{E}_{kjil},\;{D}_{ijkl}=-{D}_{kjil} (2.15)
Cijkl=Cilkj,Eijkl=Eilkj,Dijkl=Dilkj\displaystyle{C}_{ijkl}=-{C}_{ilkj},\;{E}_{ijkl}=-{E}_{ilkj},\;{D}_{ijkl}=-{D}_{ilkj} (2.16)
andCijkl=0,\displaystyle and\quad{C}_{ijkl}=0,\; Eijkl=0\displaystyle{E}_{ijkl}=0 if i=ki=k or j=lj=l (2.17)
Dijkl=0\displaystyle{D}_{ijkl}=0 if j=lj=l (2.18)

Also we recall that Cijkl=Cjikl{C}_{ijkl}={C}_{jikl} and Dijkl=Djikl{D}_{ijkl}={D}_{jikl} are given symmetry restrictions. As a result, we get

Cijkl=Ckjil=Cjkil=Cikjl,\displaystyle{C}_{ijkl}=-{C}_{kjil}=-{C}_{jkil}={C}_{ikjl},
Cijkl=Ckjil=Cjkil=Cjikl=Cijkl\displaystyle\therefore\quad{C}_{ijkl}=-{C}_{kjil}=-{C}_{jkil}=-{C}_{jikl}=-{C}_{ijkl}
Cijkl=0=0.\displaystyle\implies{C}_{ijkl}=0\implies{\mathbb{C}}=0.
Similarly,𝔻=0.\displaystyle\text{Similarly,}\quad{\mathbb{D}}=0.

Therefore the necessary sufficient conditions for null Lagrangian of above type (2.5) (with (2.6)) and within the family stated by H. Rund [17] are

=0,𝔻=0,\displaystyle{\mathbb{C}}=0,\quad{\mathbb{D}}=0, (2.19)
Eijkl=Eklij,Eijkl=Ekjil,Eijkl=Eilkj,\displaystyle{E}_{ijkl}={E}_{klij},\quad{E}_{ijkl}=-{E}_{kjil},\quad{E}_{ijkl}=-{E}_{ilkj}, (2.20)
andEijkl=0if i=k or j=l,\displaystyle\text{and}\quad{E}_{ijkl}=0\quad\text{if $i=k$ or $j=l$}, (2.21)

and the null Lagrangian is given by

Ψ(𝒙,𝒖,𝒗,𝒖,𝒗)=12κ𝔼[κ].\displaystyle{\Psi}(\boldsymbol{x},{{\boldsymbol{u}}},\boldsymbol{v},\nabla{{\boldsymbol{u}}},\nabla\boldsymbol{v})=\frac{1}{2}{\upkappa}\cdot{\mathbb{E}}[{\upkappa}]. (2.22)
Remark 7.

It can be easily verified by direct calculation that Ψ{\Psi} in (2.22) is indeed a null Lagrangian. Again, in contrast to the case of linear elasticity above null Lagrangian can be interpreted as stored energy functionals in the linear generalized elasticity theory of quasicrystals.

3 Linear electro-magneto-elasticity

With 𝒖:Ω3{\boldsymbol{u}}:\Omega\to\mathbb{R}^{3}, φ:Ω{\varphi}:\Omega\to\mathbb{R} and ψ:Ω{\psi}:\Omega\to\mathbb{R}, assuming that Cijlk,Pikl,Qikl,Eil,Bil,Ail{{C}}_{ijlk},{{P}}_{ikl},{{Q}}_{ikl},{E}_{il},{{B}}_{il},{{A}}_{il} are the elastic, piezoelectric, piezomagnetic, dielectric, magnetic permeability and electromagnetic coupling constants, respectively, the Lagrangian (enthalpy functional) for a linear electro-magneto-elastic material is given by [12]

2H(𝒖,𝒆,𝒉)=Cijklui,juk,lEijφ,iφ,jBijψ,iψ,j\displaystyle 2H({\boldsymbol{u}},{\boldsymbol{e}},{\boldsymbol{h}})={{C}}_{ijkl}{u}_{i,j}{u}_{k,l}-{E}_{ij}{\varphi}_{,i}{\varphi}_{,j}-{{B}}_{ij}{\psi}_{,i}{\psi}_{,j}
+2Pkijφ,kui,j+2Qkijφ,kui,j2Aijφ,iψ,j.\displaystyle+2{{P}}_{kij}{\varphi}_{,k}{u}_{i,j}+2{{Q}}_{kij}{\varphi}_{,k}{u}_{i,j}-2{{A}}_{ij}{\varphi}_{,i}{\psi}_{,j}. (3.1)

where the material constants satisfy the following conditions of symmetry

Cijkl=Cjikl=Cklij,Pkij=Pkji,Qkij=Qkji,Eij=Eji,Aij=Aij,Bij=Bji.{{C}}_{ijkl}={{C}}_{jikl}={{C}}_{klij},\quad{{P}}_{kij}={{P}}_{kji},\quad{{Q}}_{kij}={{Q}}_{kji},\quad{E}_{ij}={E}_{ji},\quad{{A}}_{ij}={{A}}_{ij},\quad{{B}}_{ij}={{B}}_{ji}.

Here, δjk\delta_{jk} is the Kronecker symbol and summation over repeated indices is implied. In other words, with 𝒆=φ{\boldsymbol{e}}=-{\nabla}{\varphi} and 𝒉=ψ{\boldsymbol{h}}=-{\nabla}{\psi},

H(𝒖,𝒆,𝒉)=12ϵ:ϵ12𝐄𝒆𝒆12𝐁𝒉𝒉𝙿ϵ𝒆𝚀ϵ𝒉𝐀𝒆𝒉.\displaystyle H({\boldsymbol{u}},{\boldsymbol{e}},{\boldsymbol{h}})=\frac{1}{2}{{\mathbb{C}}}{\upepsilon}:{\upepsilon}-\frac{1}{2}{{\mathbf{E}}}{\boldsymbol{e}}{\cdot}{\boldsymbol{e}}-\frac{1}{2}{\mathbf{B}}{\boldsymbol{h}}{\cdot}{\boldsymbol{h}}-{{\mathtt{P}}}{\upepsilon}{\cdot}{\boldsymbol{e}}-{{\mathtt{Q}}}{\upepsilon}{\cdot}{\boldsymbol{h}}-{\mathbf{A}}{\boldsymbol{e}}{\cdot}{\boldsymbol{h}}. (3.2)

In the theory of electro-magneto-elasticity, the stress [12] is

𝝈\displaystyle{\boldsymbol{\sigma}} =𝒖+𝙿φ+𝚀ψ=ϵ𝙿𝒆𝚀𝒉,\displaystyle={{\mathbb{C}}}\nabla{\boldsymbol{u}}+{{\mathtt{P}}}^{\top}\nabla{\varphi}+{{\mathtt{Q}}}^{\top}\nabla{\psi}={{\mathbb{C}}}{\upepsilon}-{{\mathtt{P}}}^{\top}{\boldsymbol{e}}-{{\mathtt{Q}}}^{\top}{\boldsymbol{h}}, (3.3)
𝒅\displaystyle{\boldsymbol{d}} =𝙿𝒖𝐄φ𝐀ψ=𝙿ϵ+𝐄𝒆+𝐀𝒉,\displaystyle={{\mathtt{P}}}\nabla{\boldsymbol{u}}-{{\mathbf{E}}}\nabla{\varphi}-{\mathbf{A}}\nabla{\psi}={{\mathtt{P}}}{\upepsilon}+{{\mathbf{E}}}{\boldsymbol{e}}+{\mathbf{A}}{\boldsymbol{h}}, (3.4)
𝒃\displaystyle\boldsymbol{b} =𝚀𝒖𝐀φ𝐁ψ=𝚀ϵ+𝐀𝒆+𝐁𝒉.\displaystyle={{\mathtt{Q}}}\nabla{\boldsymbol{u}}-{\mathbf{A}}\nabla{\varphi}-{\mathbf{B}}\nabla{\psi}={{\mathtt{Q}}}{\upepsilon}+{\mathbf{A}}{\boldsymbol{e}}+{\mathbf{B}}{\boldsymbol{h}}. (3.5)

The classical variation of electromagnetic enthalpy functional

(𝒖,φ,ψ)=ΩH(𝒖,𝒆,𝒉)𝑑v,\mathcal{H}({\boldsymbol{u}},{\varphi},{\psi})=\int_{\Omega}H({\boldsymbol{u}},{\boldsymbol{e}},{\boldsymbol{h}})dv,

provides the Euler–Lagrange differential equations. In the context of the Euler–Lagrange equation (1.33), the dependent variable is 𝒚=(𝒖,φ,ψ)\boldsymbol{y}=({{\boldsymbol{u}}},{\varphi},{\psi}), while 𝒖{{\boldsymbol{u}}} has dimension 33, φ{\varphi} and ψ{\psi} have dimension 11, so that N=5N=5. Rewriting (1.35) for 𝒖,φ{{\boldsymbol{u}}},{\varphi} and ψ{\psi} we get,

Ψ(𝒖,φ,ψ)=12𝒟1(α,β;i,j)ui,αuj,β+12𝒟2(α,β;4,4)φ,αφ,β+12𝒟3(α,β;5,5)ψ,αψ,β+𝒟4(α,β;4,i)ui,βφ,α+𝒟5(α,β;5,j)ui,βψ,α+𝒟6(α,β;4,5)φ,αψ,β+𝒟1(α;i)ui,α+𝒟2(α;4)φ,α+𝒟3(α;5)ψ,α+12𝒟(0;0){\Psi}(\nabla{{{\boldsymbol{u}}}},\nabla{{\varphi}},\nabla{{\psi}})=\frac{1}{2}\mathscr{D}_{1}(\alpha,\beta;i,j){u}_{i,\alpha}{u}_{j,\beta}+\frac{1}{2}\mathscr{D}_{2}(\alpha,\beta;4,4){\varphi}_{,\alpha}{\varphi}_{,\beta}+\frac{1}{2}\mathscr{D}_{3}(\alpha,\beta;5,5){\psi}_{,\alpha}{\psi}_{,\beta}\\ +\mathscr{D}_{4}(\alpha,\beta;4,i){u}_{i,\beta}{\varphi}_{,\alpha}+\mathscr{D}_{5}(\alpha,\beta;5,j){u}_{i,\beta}{\psi}_{,\alpha}+\mathscr{D}_{6}(\alpha,\beta;4,5){\varphi}_{,\alpha}{\psi}_{,\beta}\\ +\mathscr{D}_{1}(\alpha;i){u}_{i,\alpha}+\mathscr{D}_{2}(\alpha;4){\varphi}_{,\alpha}+\mathscr{D}_{3}(\alpha;5){\psi}_{,\alpha}+\frac{1}{2}\mathscr{D}(0;0) (3.6)

where 1i,j31\leq i,j\leq 3.
Comparing (3.1) and (3.6) we get,

|SiαSiβSjαSjβ|\displaystyle\begin{vmatrix}S^{\alpha}_{i}&S^{\beta}_{i}\\ S^{\alpha}_{j}&S^{\beta}_{j}\end{vmatrix} =𝒟1(α,β;i,j)=Ciαjβ,\displaystyle=\mathscr{D}_{1}(\alpha,\beta;i,j)={{C}}_{i\alpha j\beta},\quad |S4αS4βS4αS4β|\displaystyle\begin{vmatrix}S^{\alpha}_{4}&S^{\beta}_{4}\\ S^{\alpha}_{4}&S^{\beta}_{4}\end{vmatrix} =𝒟2(α,β;4,4)=Eαβ\displaystyle=\mathscr{D}_{2}(\alpha,\beta;4,4)=-{E}_{\alpha\beta} (3.7)
|S5αS5βS5αS5β|\displaystyle\begin{vmatrix}S^{\alpha}_{5}&S^{\beta}_{5}\\ S^{\alpha}_{5}&S^{\beta}_{5}\end{vmatrix} =𝒟3(α,β;5,5)=Bαβ,\displaystyle=\mathscr{D}_{3}(\alpha,\beta;5,5)=-{B}_{\alpha\beta},\quad |S4αS4βSiαSiβ|\displaystyle\begin{vmatrix}S^{\alpha}_{4}&S^{\beta}_{4}\\ S^{\alpha}_{i}&S^{\beta}_{i}\end{vmatrix} =𝒟4(α,β;4,i)=Pαiβ\displaystyle=\mathscr{D}_{4}(\alpha,\beta;4,i)={{P}}_{\alpha i\beta} (3.8)
|S5αS5βSiαSiβ|\displaystyle\begin{vmatrix}S^{\alpha}_{5}&S^{\beta}_{5}\\ S^{\alpha}_{i}&S^{\beta}_{i}\end{vmatrix} =𝒟5(α,β;5,i)=Qαiβ,\displaystyle=\mathscr{D}_{5}(\alpha,\beta;5,i)={{Q}}_{\alpha i\beta},\quad |S4αS4βS5αS5β|\displaystyle\begin{vmatrix}S^{\alpha}_{4}&S^{\beta}_{4}\\ S^{\alpha}_{5}&S^{\beta}_{5}\end{vmatrix} =𝒟6(α,β;4,5)=Aαβ\displaystyle=\mathscr{D}_{6}(\alpha,\beta;4,5)=-{A}_{\alpha\beta} (3.9)
|SiαSiβS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{i}&S^{\beta}_{i}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟1(α;i)=0,\displaystyle=\mathscr{D}_{1}(\alpha;i)=0,\quad |S4αS4βS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{4}&S^{\beta}_{4}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟2(α;4)=0,\displaystyle=\mathscr{D}_{2}(\alpha;4)=0, (3.10)
|S5αS5βS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{5}&S^{\beta}_{5}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟3(α;5)=0,\displaystyle=\mathscr{D}_{3}(\alpha;5)=0,\quad |S|ααS|αβS|βαS|ββ|\displaystyle\begin{vmatrix}S^{\alpha}_{|\alpha}&S^{\beta}_{|\alpha}\\ S^{\alpha}_{|\beta}&S^{\beta}_{|\beta}\end{vmatrix} =𝒟(0;0)=0\displaystyle=\mathscr{D}(0;0)=0 (3.11)

Thus, within the family of null Lagrangians stated by H. Rund [17], we find that these are necessary and sufficient conditions on ,𝐄,𝐁,𝙿,𝚀,\mathbb{C},{\mathbf{E}},{\mathbf{B}},{{\mathtt{P}}},{{\mathtt{Q}}}, and 𝐀{\mathbf{A}} so that (3.1) becomes a null Lagrangian. As in the case of previous section, we obtain the explicit form of conditions on the material constants ,𝐄,𝐁,𝙿,𝚀\mathbb{C},\mathbf{E},{\mathbf{B}},{{\mathtt{P}}},{{\mathtt{Q}}} and 𝐀{\mathbf{A}}. From (3.7)1 we get =0\mathbb{C}=0. From (3.7)2 and (3.8)1 we get 𝐄=0{\mathbf{E}}=0 and 𝐁=0{\mathbf{B}}=0 respectively. From (3.8)2 and (3.9)1 we get the following restrictions on 𝙿{{\mathtt{P}}} and 𝚀{{\mathtt{Q}}},

Pkij=Pjik,\displaystyle{{P}}_{kij}=-{{P}}_{jik},\; Qkij=Qjik\displaystyle{{Q}}_{kij}=-{{Q}}_{jik} (3.12)
Pkij=0,\displaystyle{{P}}_{kij}=0,\; Qkij=0\displaystyle{{Q}}_{kij}=0 if j=kj=k (3.13)

Also given Pkij=Pkji,Qkij=Qkji{{P}}_{kij}={{P}}_{kji},{{Q}}_{kij}={{Q}}_{kji}.

Pkij=Pjik=Pjki=Pikj=Pijk=Pkji=Pkij\displaystyle\therefore\quad{{P}}_{kij}=-{{P}}_{jik}=-{{P}}_{jki}={{P}}_{ikj}={{P}}_{ijk}=-{{P}}_{kji}=-{{P}}_{kij}
𝙿=0.\displaystyle\implies{{\mathtt{P}}}=0.

Same holds for 𝚀{{\mathtt{Q}}} so 𝚀=0{{\mathtt{Q}}}=0 as well. From (3.9)2 we get the following restrictons on 𝐀{\mathbf{A}},

Aij=Aji,\displaystyle{A}_{ij}=-{A}_{ji},
so that Aij=0 if i=j.\displaystyle\text{ so that }{A}_{ij}=0\text{ if $i=j$}.

But it is given that Aij=Aji{A}_{ij}={A}_{ji}, 𝐀=0\implies\mathbf{A}=0. Therefore, within the family of null Lagrangians stated by H. Rund [17], we do not obtain any non-zero null Lagrangians in this case. This conclusion resonates with that in linear elasticity where null Lagrangians cannot be interpreted as stored energy functionals.

4 Conclusion

In this paper, we explored the concept of null Lagrangians within the linear framework of generalized elasticity. In particular, we have constructed a family of null Lagrangians for linear Cosserat or micropolar elastic media as well as linear quasicrystal models. We found out that there does not exist a non-trivial null Lagrangian for (linear) piezoelectric and piezomagnetic media, or in general linear electro-magneto-elasticity. For the cases where the non-trivial null Lagrangian exists, we provide a split of the stored energy where the null Lagrangian part of it vanishes if and only if the relevant elasticity tensor obeys certain symmetry conditions as a reminder of Cauchy relations in classical elasticity. We have presented analysis of isotropic models and few other special cases as well. At the end of the analyses presented, a question arises about the extension to nonlinear models of the discussed formulations; this is currently under investigation [48].

Acknowledgments

BLS acknowledges the partial support of SERB MATRICS grant MTR/2017/000013.

Appendix A Characterization of null Lagrangians

If the right-hand side of (1.33) is to vanish for all values of 2yjxβxγ\frac{{\operatorname{\partial}}^{2}y^{j}}{{\operatorname{\partial}}x^{{{\beta}}}{\operatorname{\partial}}x^{{{\gamma}}}}, the coefficient of these quantities which is symmetric in β{{{\beta}}} and γ{{{\gamma}}}, must be skew-symmetric in these indices, i.e.,

2Ψyβjyγk=2Ψyγjyβk.\frac{{\operatorname{\partial}}^{2}{\Psi}}{{\operatorname{\partial}}y^{j}_{{{\beta}}}{\operatorname{\partial}}y^{k}_{{{\gamma}}}}=-\frac{{\operatorname{\partial}}^{2}{\Psi}}{{\operatorname{\partial}}y^{j}_{{{\gamma}}}{\operatorname{\partial}}y^{k}_{{{\beta}}}}.

Following the analysis of [17], in general, Ψ{\Psi} is a polynomial in its dependence on the derivative of yy. In the class of Lagrangians which are interesting in this document, it is sufficient to consider M=2.M=2. Thus, Ψ{\Psi} can be written as

Ψ=Ψ(2)+Ψ(1)+Φ,{\Psi}={\Psi}^{(2)}+{\Psi}^{(1)}+\Phi, (A.1)
where Ψ(1):=Ai1α1yα1i1,Ψ(2):=12Ai1i2α1α2yα1i1yα2i2.\text{where }{\Psi}^{(1)}:=A^{\alpha_{1}}_{i_{1}}y^{i_{1}}_{\alpha_{1}},\quad{\Psi}^{(2)}:=\frac{1}{2}A^{\alpha_{1}\alpha_{2}}_{i_{1}i_{2}}y^{i_{1}}_{\alpha_{1}}y^{i_{2}}_{\alpha_{2}}. (A.2)

Here, Φ,𝐀,𝔸\Phi,\mathbf{A},{{\mathbb{A}}} are functions of 𝒙\boldsymbol{x} and 𝒚\boldsymbol{y} of class 𝒞2\mathcal{C}^{2}. Clearly Ai1i2α1α2=Ai2i1α2α1.A^{\alpha_{1}\alpha_{2}}_{i_{1}i_{2}}=A^{\alpha_{2}\alpha_{1}}_{i_{2}i_{1}}. Then (1.34) becomes

n=12k(Ψ(n))+k(Φ)0.\sum_{n=1}^{2}{\mathscr{E}}_{k}({\Psi}^{(n)})+{\mathscr{E}}_{k}(\Phi)\equiv 0. (A.3)

On substituting from (1.33) we see that this represents a set of PDEs for 𝐀\mathbf{A} and 𝔸{{\mathbb{A}}}, whose precise form we have to determine.

From (A.2) we have Ψ(2)yγk=Aki2γα2yα2i2Ψ(1)yγk=Akγ.\frac{{\operatorname{\partial}}{\Psi}^{(2)}}{{\operatorname{\partial}}y^{k}_{{{\gamma}}}}=A^{{{{\gamma}}}\alpha_{2}}_{ki_{2}}y^{i_{2}}_{\alpha_{2}}\quad\frac{{\operatorname{\partial}}{\Psi}^{(1)}}{{\operatorname{\partial}}y^{k}_{{{\gamma}}}}=A^{{{\gamma}}}_{k}. Therefore

k(Ψ(2))\displaystyle{\mathscr{E}}_{k}({\Psi}^{(2)}) =xγ(Aki2γα2)yα2i2+[yi1(Aki2α1α2)12yk(Ai1i2α1α2)]yα1i1yα2i2+Akjγβ2yjxβxγ,\displaystyle=\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{{\gamma}}}\alpha_{2}}_{ki_{2}}\right)y^{i_{2}}_{\alpha_{2}}+\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{1}}}\left(A^{\alpha_{1}\alpha_{2}}_{ki_{2}}\right)-\frac{1}{2}\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{k}}\left(A^{\alpha_{1}\alpha_{2}}_{i_{1}i_{2}}\right)\right]y^{i_{1}}_{\alpha_{1}}y^{i_{2}}_{\alpha_{2}}+A^{{{{\gamma}}}{{{\beta}}}}_{kj}\frac{{\operatorname{\partial}}^{2}y^{j}}{{\operatorname{\partial}}x^{{{\beta}}}{\operatorname{\partial}}x^{{{\gamma}}}},
k(Ψ(1))\displaystyle{\mathscr{E}}_{k}({\Psi}^{(1)}) =xγ(Akγ)+[yi1(Akα1)yk(Ai1α1)]yα1i1.\displaystyle=\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{\gamma}}}_{k}\right)+\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{1}}}\left(A^{\alpha_{1}}_{k}\right)-\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{k}}\left(A^{\alpha_{1}}_{i_{1}}\right)\right]y^{i_{1}}_{\alpha_{1}}.

From (A.3) we get,

[xγ(Aki2γα2)+yi2(Akα2)yk(Ai2α2)]yα2i2+[yi1(Aki2α1α2)12yk(Ai1i2α1α2)]yα1i1yα2i2\displaystyle\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{{\gamma}}}\alpha_{2}}_{ki_{2}}\right)+\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{2}}}\left(A^{\alpha_{2}}_{k}\right)-\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{k}}\left(A^{\alpha_{2}}_{i_{2}}\right)\right]y^{i_{2}}_{\alpha_{2}}+\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{1}}}\left(A^{\alpha_{1}\alpha_{2}}_{ki_{2}}\right)-\frac{1}{2}\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{k}}\left(A^{\alpha_{1}\alpha_{2}}_{i_{1}i_{2}}\right)\right]y^{i_{1}}_{\alpha_{1}}y^{i_{2}}_{\alpha_{2}}
+Akjγβ2yjxβxγ+[xγ(Akγ)Φyk]=0.\displaystyle\quad\quad\quad\quad\quad+A^{{{{\gamma}}}{{{\beta}}}}_{kj}\frac{{\operatorname{\partial}}^{2}y^{j}}{{\operatorname{\partial}}x^{{{\beta}}}{\operatorname{\partial}}x^{{{\gamma}}}}+\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{\gamma}}}_{k}\right)-\frac{{\operatorname{\partial}}\Phi}{{\operatorname{\partial}}y^{k}}\right]=0. (A.4)

This holds for all values of yαiy^{i}_{\alpha} and 2yjxβxγ\frac{{\operatorname{\partial}}^{2}y^{j}}{{\operatorname{\partial}}x^{{{\beta}}}{\operatorname{\partial}}x^{{{\gamma}}}}. So every term is zero separately. In particular,

[xγ(Aki2γα2)+yi2(Akα2)yk(Ai2α2)]=0,[xγ(Akγ)Φyk]=0.\displaystyle\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{{\gamma}}}\alpha_{2}}_{ki_{2}}\right)+\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{2}}}\left(A^{\alpha_{2}}_{k}\right)-\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{k}}\left(A^{\alpha_{2}}_{i_{2}}\right)\right]=0,\quad\left[\frac{\operatorname{\partial}}{{\operatorname{\partial}}x^{{{\gamma}}}}\left(A^{{{\gamma}}}_{k}\right)-\frac{{\operatorname{\partial}}\Phi}{{\operatorname{\partial}}y^{k}}\right]=0. (A.5)

for k=1,,N.k=1,\dots,N.

Let us consider 33 functions (of 𝒞2\mathcal{C}^{2} smoothness) {Sα(𝒙,𝒚)}α=1,2,3\{S^{\alpha}(\boldsymbol{x},\boldsymbol{y})\}_{\alpha=1,2,3} and define

Siα:=Sαyi;S|βα:=Sαxβ.S^{\alpha}_{i}:=\frac{{\operatorname{\partial}}S^{\alpha}}{{\operatorname{\partial}}y^{i}};\quad S^{\alpha}_{|\beta}:=\frac{{\operatorname{\partial}}S^{\alpha}}{{\operatorname{\partial}}x^{\beta}}.

Recall 𝔇\mathfrak{D}s as defined by (1.36). Therefore,

𝔇(α1;i1)xα1\displaystyle\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1};i_{1})}{{\operatorname{\partial}}x^{\alpha_{1}}} =|Si1|α1α1Si1|α1α2S|α2α1S|α2α2|+|Si1α1Si1α2S|α1α2α1S|α1α2α2|\displaystyle=\begin{vmatrix}S^{\alpha_{1}}_{i_{1}|\alpha_{1}}&S^{\alpha_{2}}_{i_{1}|\alpha_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix}+\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{|\alpha_{1}\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{1}\alpha_{2}}\end{vmatrix}
(α1,α2\alpha_{1},\alpha_{2} both being dummy indices, the latter determinant is 0) (A.6)
=yi1|S|α1α1S|α1α2S|α2α1S|α2α2||S|α1α1S|α1α2Si1|α2α1Si1|α2α2|=𝔇(0;0)yi1𝔇(α1;i1)xα1\displaystyle=\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{1}}}\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{1}}&S^{\alpha_{2}}_{|\alpha_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix}-\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{1}}&S^{\alpha_{2}}_{|\alpha_{1}}\\ S^{\alpha_{1}}_{i_{1}|\alpha_{2}}&S^{\alpha_{2}}_{i_{1}|\alpha_{2}}\end{vmatrix}=\frac{{\operatorname{\partial}}\mathfrak{D}(0;0)}{{\operatorname{\partial}}y^{i_{1}}}-\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1};i_{1})}{{\operatorname{\partial}}x^{\alpha_{1}}}
2𝔇(α1;i1)xα1\displaystyle\implies 2\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1};i_{1})}{{\operatorname{\partial}}x^{\alpha_{1}}} =𝔇(0;0)yi1.\displaystyle=\frac{{\operatorname{\partial}}\mathfrak{D}(0;0)}{{\operatorname{\partial}}y^{i_{1}}}. (A.7)

Also,

𝔇(α1,α2;i1,i2)xα2\displaystyle\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1},\alpha_{2};i_{1},i_{2})}{{\operatorname{\partial}}x^{\alpha_{2}}} =|Si1|α2α1Si1|α2α2Si2α1Si2α2|+|Si1α1Si1α2Si2|α2α1Si2|α2α2|\displaystyle=\begin{vmatrix}S^{\alpha_{1}}_{i_{1}|\alpha_{2}}&S^{\alpha_{2}}_{i_{1}|\alpha_{2}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix}+\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{i_{2}|\alpha_{2}}&S^{\alpha_{2}}_{i_{2}|\alpha_{2}}\end{vmatrix}
=yi1|S|α2α1S|α2α2Si2α1Si2α2|+yi2|Si1α1Si1α2S|α2α1S|α2α2||S|α2α1S|α2α2Si1i2α1Si1i2α2||Si1i2α1Si1i2α2S|α2α1S|α2α2|\displaystyle=\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{1}}}\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\\ S^{\alpha_{1}}_{i_{2}}&S^{\alpha_{2}}_{i_{2}}\end{vmatrix}+\frac{\operatorname{\partial}}{{\operatorname{\partial}}y^{i_{2}}}\begin{vmatrix}S^{\alpha_{1}}_{i_{1}}&S^{\alpha_{2}}_{i_{1}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix}-\begin{vmatrix}S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\\ S^{\alpha_{1}}_{i_{1}i_{2}}&S^{\alpha_{2}}_{i_{1}i_{2}}\end{vmatrix}-\begin{vmatrix}S^{\alpha_{1}}_{i_{1}i_{2}}&S^{\alpha_{2}}_{i_{1}i_{2}}\\ S^{\alpha_{1}}_{|\alpha_{2}}&S^{\alpha_{2}}_{|\alpha_{2}}\end{vmatrix}
=𝔇(α1;i2)yi1+𝔇(α1;i1)yi2\displaystyle=-\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1};i_{2})}{{\operatorname{\partial}}y^{i_{1}}}+\frac{{\operatorname{\partial}}\mathfrak{D}(\alpha_{1};i_{1})}{{\operatorname{\partial}}y^{i_{2}}}
(the remaining two determinants cancel each other).\displaystyle\text{(the remaining two determinants cancel each other)}.

Comparing this with (A.5)1 we get

Ai1i2α1α2=𝔇(α1,α2;i1,i2),Ai1α1=𝔇(α1;i1).A^{\alpha_{1}\alpha_{2}}_{i_{1}i_{2}}=\mathfrak{D}(\alpha_{1},\alpha_{2};i_{1},i_{2}),\quad A^{\alpha_{1}}_{i_{1}}=\mathfrak{D}(\alpha_{1};i_{1}). (A.8)

From (A.7), 12𝔇(0;0)yi1=Ai1α1xα1=Φyi1\frac{1}{2}\frac{{\operatorname{\partial}}\mathfrak{D}(0;0)}{{\operatorname{\partial}}y^{i_{1}}}=\frac{{\operatorname{\partial}}A^{\alpha_{1}}_{i_{1}}}{{\operatorname{\partial}}x^{\alpha_{1}}}=\frac{{\operatorname{\partial}}\Phi}{{\operatorname{\partial}}y^{i_{1}}} by (A.5)2, which upon integrating yields

Φ=12𝔇(0;0).\Phi=\frac{1}{2}\mathfrak{D}(0;0).

Therefore, reverting back to (A.1) we get (1.35).

Using (A.1), (A.2), (A.4) and (A.8) we can summarize the result as:

Theorem 1 ([17], pg 257-258).

Any function of the form (1.35) satisfies Euler-Lagrange equation (1.34) identically, where 𝒞2\mathcal{C}^{2} functions Sα(𝐱,𝐲)S^{\alpha}(\boldsymbol{x},\boldsymbol{y}) are entirely arbitrary.

References

  • [1] E. Cosserat and F. Cosserat. Théorie des corps déformables. Librairie Scientifique A. Hermann et Fils (english translation by D. Delphenich, 2007), reprint 2009, Paris (1909)
  • [2] A. E. H. Love, A Treatise on the Mathematical Theory of Elasticity, 4th edn. Cambridge University Presss Cambridge, UK (1927).
  • [3] C. Carathéodory, Uber die Variationsrechnung bei mehrfachen Integralen, Acta Szeged Sect. Scient. Mathem. 4, 193 (1929).
  • [4] A. W. Landers, Jr., Invariant Multiple Integrals in the Calculus of Variations. Univ. Chicago Press, Chicago (1942). Also: in Contributions to the Calculus of Variations, 1938-1941, 175 (Univ. Chicago Press, 1942).
  • [5] J. Ericksen and C. Truesdell. Exact theory of stress and strain in rods and shells. Arch. Rational Mech. Anal., 1: 295–323 (1958).
  • [6] D. G. B. Edelen, The Null Set of the Euler-Lagrange Operator, Arch. Rational Mech. Anal., 11, p. 117-121. (1962)
  • [7] J. Ericksen, Nilpotent energies in liquid crystal theory, Arch. Rational Mech. Anal., 10, 189–196. (1962)
  • [8] M. Picone, Criteri sufficienti peril minima assoluto di un integrale pluridimensionale del primo ordine nel vettore minimante, Mem. Accad. Naz. Lincei VI (1962) 281-338.
  • [9] M. Picone, Criteri sufficienti peril minima assoluto di un integrale pluridimensionale del primo ordine nel vettore minimante in parte o completamente libero alla frontiera del dominio di integrazione, Mem Accad. Naz. Lincei VI (1964) 341-374.
  • [10] M. De Franchis, La piu generale funzione d’invarianza per criteri sufficienti di minima con condizioni di Dirichlet per integrali pluridimensionali del primo ordine dipendenti da un vettore a piu componenti, Rend. Accad. Naz. Lincei XXXVII (1964) 130-140.
  • [11] E. L. Aero & E. V. Kuvshinski, Continuum theory of asymmetric elasticity. Microrotation effect, Solid State Physics, 5(9), 2591–2598 (in Russian) (1963) (English translation: Soviet Physics–Solid State, 5 (1964), 1892–1899)
  • [12] Berlincourt, D. A., Curran, D. R., & Jaffe, H. Piezoelectric and piezomagnetic materials and their function in transducers. Physical Acoustics: Principles and Methods, 1(Part A), 169-270.. (1964)
  • [13] R. Toupin. Theories of elasticity with couple-stresses. Arch. Rational Mech. Anal., 17: 85–112 (1964).
  • [14] R. Mindlin. Microstructure in linear elasticity. Arch. Rational Mech. Anal., 16: 51–78 (1964).
  • [15] E. L. Aero and E. V. Kuvshinski, Continuum theory of asymmetric elasticity. Equilibrium of an isotropic body, Soviet Physics–Solid State. 6, 2141–2148. (1965)
  • [16] A.C. Eringen, Linear theory of micropolar elasticity. J. Math. Mech., 15, 909–923 (1966)
  • [17] H. Rund, The Hamiltonian-Jacobi Theory in the Calculus of Variation, Its Role in Mathematics and Physics, D. Van Nostrand Company LTD (1966)
  • [18] H. Rund, Integral formulae associated with the Euler-Lagrange operators of multiple integral problems in the calculus of variations, Aequationes Mathematicae, volume 11, page 212-229 (1974)
  • [19] D.M. Barnett and J. Lothe, Dislocations and Line Charges in Anisotropic Piezoelectric Insulators, phys. stat. sol. (b) 67 105–111. (1975)
  • [20] A.C.Eringen and C.B.Kadafar, Polar field theories, in Continuum Physics, A. C. Eringen, Ed., vol. 4, pp. 1–73, Academic Press, New York, NY, USA, (1976).
  • [21] J. P. Nowacki, Green’s function for a hemitropic micropolar continuum, Green’s function for a hemitropic micropolar continuum, (8) 235. (1977)
  • [22] D. G. B. Edelen, Isovector Methods for Equations of Balance with Programs for Computer Assistance in Operator Calculations and an Exposition of Practical Topics of the Exterior Calculus. Sijthoff and Noordhoff, The Netherlands (1980).
  • [23] Gurtin, M. E. An introduction to continuum mechanics. New York: Academic Press. (1981)
  • [24] J. M. Ball, J. C. Currie and P. J. Olver, Null Lagrangians, weak continuity, and variational problems of arbitrary order, J. Funct. Anal. 41 (1981) 135-174.
  • [25] R. S. Lakes and R. L. Benedict. Noncentrosymmetry in micropolar elasticity. International Journal of Engineering Science, 20(10):1161–1167 (1982)
  • [26] W. Nowacki, Theory of Asymmetric Elasticity, PWN-Polish Scientific Publishers, ISBN: 0-08-027584-2. (1986)
  • [27] D. G. B. Edelen & D. C. Lagoudas, Null Lagrangians, admissible tractions, and finite element methods, International Journal of Solids and Structures, Volume 22, Issue 6, , Pages 659-672 (1986)
  • [28] G. A. Maugin. Continuum mechanics of electromagnetic solids. Oxford: Elsevier. (1988)
  • [29] P. J. Olver and J. Sivaloganathan, Structure of null Lagrangians, Nonlinearity 1 (1988) 389-398.
  • [30] A. C. Eringen. & G. A. Maugin. Electrodynamics of continua. New York: Springer. (1990).
  • [31] V.I. Alshits, A.N. Darinskii and J. Lothe, On the existence of interface waves in half-infinite anisotropic elastic media with piezoelectronic and piezomagnetic properties, Wave Motion 16 265–283. (1992)
  • [32] R. Lakes.Materials with structural hierarchy.Nature, 361:511–515 (1993)
  • [33] D.H. Ding, W. Wang, C. Hu and R. Yang, Generalized elasticity theory of quasicrystals, Phys. Rev. B 48 7003–7010. (1993).
  • [34] A. Campanella, M. L. Tonon, A Note on the Cauchy Relations, Meccanica, p. 105-108. (1994)
  • [35] E. Virga, Variational Theories for Liquid Crystals, Appl. Math. and Math. Comput., 8, Chapman & Hall (1994).
  • [36] M. R. Lancia, G. V. Caffarelli, P. Podio-Guidugli, Null Lagrangians in Linear Elasticity, Mathematical Models and Methods in Applied Sciences, p. 415-427. (1995)
  • [37] A. C. Eringen. Microcontinuum Field Theories - Volume 1: Foundations and Solids. Springer Verlag, New York - Berlin - Heidelberg (1999)
  • [38] C. Hu, R. Wang and D.H. Ding, Symmetry groups, physical property tensors, elasticity and dislocations in quasicrystals, Rep. Prog. Phys. 63 1–39. (2000)
  • [39] R. Lakes. Elastic and viscoelastic behavior of chiral materials. In K. Golden, G. Grimmert, R. James, G. Milton, and P. Sen, editors, Int. J. Mech. Sci., volume 43, pages 1579–1589. (2001)
  • [40] F. W. Hehl, Y. Itin, The Cauchy Relations in Linear Elasticity Theory, Journal of Elasticity, p. 185-192. (2002)
  • [41] Carillo, S. Null Lagrangians and surface interaction potentials in nonlinear elasticity, Series on advances in mathematics for applied sciences, 62: 9-18. (2002).
  • [42] P. Sharma, Size-dependent elastic fields of embedded inclusions in isotropic chiral solids, International Journal of Solids and Structures, Volume 41, Issues 22–23, Pages 6317-6333, (2004)
  • [43] D. Natroshvili, I. .G. Stratis. Mathematical problems of the theory of elasticity of chiral materials for Lipschitz domains. Math. Methods Appl. Sci., 29, 445–478. (2006)
  • [44] W. Pietraszkiewicz & V. A. Eremeyev. On natural strain measures of the non-linear micropolar continuum. International Journal of Solids and Structures, 46(3–4), 774–787. (2009)
  • [45] H. Joumaa and M. Ostoja-Starzewski. Stress and couple-stress invariance in non-centrosymmetric micropolar planar elasticity. Proc. R. Soc. London A, 467(2134):2896–2911 (2011)
  • [46] T.Y. Fan, Mathematical Theory of Elasticity of Quasicrystals and Its Applications, Science Press, Beijing and Springer-Verlag, Berlin, Heidelberg, 2011.
  • [47] Y. Itin, F. W. Hehl, The constitutive tensor of linear elastic Its decompositions, Cauchy relations, Lagrangians, and wave propagation, J. Math. Phys., 54, 042903. (2013)
  • [48] Sharma, B. L. & Basak, N. Null lagrangians in Cosserat elasticity. in preparation. (2020)