9.6.3.1 The deformation-temperature-electric field-magnetic field formulation

In this formulation, F, Θ, e*, and h* are the independent variables. We define the energy potential EFΘeh as the Legendre transformation of internal energy ε=ε˘(F,η,p/ρ,m/ρ)si486_e with respect to the thermal, electrical, and magnetic variables, from η to Θ, p*/ρ to e*, and m*/ρ to h*, i.e.,

EFΘeh=εΘηepρμohmρ.

si487_e  (9.71)

Refer to Table 9.4. Taking the rate of (9.71) gives

E.FΘeh=ε.Θη.ηΘ.e(pρ)¯¯¯¯¯¯¯¯pρe.μoh(mρ)¯¯¯¯¯¯¯¯μomρh.,

si488_e

and use of this result in (9.61) yields the second law statement

E.FΘeh+1ρRPF.ηΘ.pρe.μomρh.+1ρje1ρΘqgradΘ0.

si489_e  (9.72)

Via the chain rule, we have

E.FΘeh=EFΘehFF.+EFΘehΘΘ.+EFΘehee.+EFΘehhh..

si490_e  (9.73)

Substitution of (9.73) into (9.72), and use of the standard arguments, leads to

P=ρREFΘehF,η=EFΘehΘ,p=ρEFΘehe,m=ρμoEFΘehh.

si491_e  (9.74)

9.6.4 Restrictions imposed by invariance under superposed rigid body motions and conservation of angular momentum

The constitutive equations (9.74) for the deformation-temperature-electric field-magnetic field formulation (or any other formulation for that matter, e.g., (9.64) or (9.70)), obtained from restrictions imposed by the second law, must also satisfy invariance under superposed rigid body motions. In particular, we must have [30]

(EFΘeh)+=EFΘeh

si492_e  (9.75a)

when

F+=QF,Θ+=Θ,(e)+=Qe,(h)+=Qh

si493_e  (9.75b)

for all proper orthogonal tensors Q(t). Equations (9.75a) and (9.75b) are referred to as invariance requirements. It can be shown that these invariance requirements demand that

EFΘeh(F,Θ,e,h)=E˜FΘeh(C,Θ,eR,hR),

si494_e  (9.76)

where

C=FTF,eR=FTe,hR=FTh

si495_e

are the right Cauchy-Green deformation tensor, referential electric field, and referential magnetic field, respectively. We can then verify, through a change of independent variable from F to C, e* to eR, and h* to hR, that the constitutive equations (9.74) become

T=2ρFE˜FΘehCFTepμohm,η=E˜FΘehΘ,p=ρFE˜FΘeheR,m=ρμoFE˜FΘehhR.

si496_e  (9.77)

These are the invariant forms of the constitutive equations. We emphasize that E˜FΘehsi497_e denotes the free energy function with C, Θ, eR, and hR as independent variables. A useful result follows from a change of the mechanical independent variable from the right Cauchy-Green deformation C to the Lagrangian strain E, where E = 1/2 (CI), in which case (9.77) becomes

T=ρFE˘FΘehEFTepμohm,η=E˘FΘehΘ,p=ρFE˜FΘeheR,m=ρμoFE˘FΘehhR,

si498_e  (9.78)

where E˘FΘehsi499_e denotes the free energy function with E, Θ, eR, and hR as independent variables. It can be verified that (9.78) satisfies conservation of angular momentum. In the following section, we use (9.78) as a starting point for developing constitutive equations for materials with linear, reversible TEME response.

Exercises

1. Prove that the invariance requirements

(EFΘeh)+=EFΘeh

si500_e

when

F+=QF,Θ+=Θ,(e)+=Qe,(h)+=Qh

si501_e

for all proper orthogonal tensors Q(t) demand that

EFΘeh(F,Θ,e,h)=E˜FΘeh(C,Θ,eR,hR).

si502_e

2. Verify that the invariant form of the constitutive equation for the Cauchy stress, i.e.,

T=ρFE˘FΘehEFTepμohm,

si503_e

satisfies conservation of angular momentum (9.57b).

9.7 Constitutive model development for thermo-electro-magneto-elastic materials: small-deformation theory

Materials exhibiting spontaneous polarization or magnetization in the presence of external electric or magnetic fields are broadly classified as ferroic materials. When coupled with mechanical stresses, ferromagnetic materials exhibit magnetostriction (i.e., deformation induced by a magnetic field), and ferroelectric materials exhibit electrostriction (i.e., deformation induced by an electric field). Materials exhibiting both orders of ferroic behavior are called multiferroics. These ferroic effects are often hysteretic, dissipative, and nonlinear, but for near-equilibrium processes in the small-deformation and small-electromagnetic-field regime, the associated thermo-electro-magneto-elastic (TEME) constitutive equations can often be approximated as linear and reversible.

9.7.1 Small-deformation kinematics, kinetics, electromagnetic fields, and fundamental laws

The small-deformation theory (or, equivalently, the small-strain or infinitesimal theory) is customarily obtained by assuming that the displacements u are small, and expanding u in a power series with respect to a small parameter.12 As a consequence of this approximation, (3.7), (3.27), (3.36), (3.54), and (3.55) become, to leading order,13,14

x=X,F=I,J=1,E=e=12[gradu+(gradu)T].

si506_e  (9.79)

Hence, in the small-deformation theory, the reference and present configurations of the body are indistinguishable, the referential (Lagrangian) and spatial (Eulerian) descriptions of a given quantity are one and the same, and the Eulerian and Lagrangian strains are identical. In the small-deformation theory, we also have

.=,Grad=grad,

si507_e  (9.80)

where ø is a scalar-valued (or vector-valued or tensor-valued) function of position and time; refer to (3.20) and (3.21). It then follows from (9.41) and (9.79) that, to leading order,

eR=e,pR=p,dR=d,hR=h,mR=m,bR=b,σR=σ,jR=j,P=T,qR=q.

si508_e  (9.81)

We assume that the electromagnetic fields are small, so their contributions to balance of linear momentum and angular momentum emerge as higher-order terms. In other words, the electromagnetic body force and body couple vanish at leading order, i.e.,

fem=0,Γem=0.

si509_e  (9.82)

An additional implication of the small-field assumption is that

e=e,p=p,d=d,h=h,m=m,b=b,σ=σ,j=j.

si510_e  (9.83)

That is, to leading order, the effective fields collapse to the corresponding standard fields, and all standard fields are equivalent (the v0 limit of (9.51)(9.54)).

As a consequence of (9.79)(9.83), we hereafter omit adjectives such as “referential,” “spatial,” “effective,” and “standard” that are no longer needed in the small-deformation/small-field theory, and instead refer to E as the infinitesimal strain, T the stress, e the electric field, p the electric polarization, d the electric displacement, and so on. Also, (9.79)(9.83) imply that the fundamental laws (9.49a)(9.49j), when expressed in Cartesian component notation, become, to leading order,

ρ=ρR,

si511_e  (9.84a)

ρui,tt=ρfmi+Tij,j,

si512_e  (9.84b)

Tij=Tji,

si513_e  (9.84c)

ρε,t=TijEij,t+eipi,t+μohimi,t+ρrt+jieiqi,i,

si514_e  (9.84d)

ρη,tρrtΘ(qiΘ),i,

si515_e  (9.84e)

σ,t+ji,i=0,

si516_e  (9.84f)

bi,i=0,

si517_e  (9.84g)

bi,t+εijkek,j=0,

si518_e  (9.84h)

di,i=σ,

si519_e  (9.84i)

di,t+ji=εijkhk,j.

si520_e  (9.84j)

The strain-displacement relationship (9.79)4, expressed in Cartesian component notation, becomes

Eij=12(ui,j+uj,i).

si521_e  (9.85)

Note that (),t denotes partial differentiation with respect to time, e.g.,

bi,tbi(x1,x2,x3,t)t.

si522_e

Also note that conservation of angular momentum (9.84c) and definition (9.85) imply that the stress T and infinitesimal strain E are symmetric.

9.7.2 Linear constitutive equations

With use of relationships (9.79)(9.83), the TEME constitutive equations (9.78) become, to leading order in the small-deformation/small-field theory,

Tij=ρψEij,η=ψΘ,pi=ρψei,mi=ρμoψhi,

si523_e  (9.86)

where we have introduced ψ=E˘FΘehsi524_e for notational brevity. In what follows, starting from (9.86), we derive linear TEME constitutive equations for ferroic materials within the near-equilibrium regime. To obtain a functional form of the free energy ψ that characterizes this linear, reversible TEME response, we perform a Taylor series expansion of ψ in terms of its independent variables E, Θ, e, and h in the neighborhood of a thermodynamic equilibrium state xo = (Eo, Θ°, eo, ho):

E˘FΘehψ(E,Θ,e,h)=ψ|xo+122ψEijEklxoEijEkl+122ψeiejxoeiej+122ψhihjxohihj+122ψΘ2xoΘ2+2ψEijekxoEijek+2ψEijhkxoEijhk+2ψeihjxoeihj+2ψΘEijxoΘEij+2ψΘeixoΘei+2ψΘhixoΘhi+higher oreder terms.

si525_e  (9.87)

Note that the TEME quantities Eij, Θ, ei, and hi in (9.87) are perturbed about the equilibrium statexo. Disregarding the higher-order terms (i.e., truncating at second order), we substitute the free energy expansion (9.87) into (9.86) to obtain the linear form of the constitutive equations:

Tij=2ψ¯¯EijEklxoEkl+2ψ¯¯Eijekxoek+2ψ¯¯Eijhkxohk+2ψ¯¯EijΘxoΘ,

si526_e  (9.88a)

pi=2ψ¯¯EjkeixoEjk2ψ¯¯eiejxoej2ψ¯¯eihjxohj2ψ¯¯eiΘxoΘ,

si527_e  (9.88b)

μomi=2ψ¯¯EjkhixoEjk2ψ¯¯ejhixoej2ψ¯¯hihjxohj2ψ¯¯hiΘxoΘ,

si528_e  (9.88c)

η¯=2ψ¯¯EijΘxoEij2ψ¯¯eiΘxoei2ψ¯¯hiΘxohi2ψ¯¯Θ2xoΘ,

si529_e  (9.88d)

where ψ¯¯=ρψsi530_e and η¯=ρηsi531_e denote the free energy per unit volume and the free entropy per unit volume, respectively. The constant coefficients arising in the linearized constitutive equations (9.88a)(9.88d) are characterized using experimental data specific to the material. For instance, 2ψ¯¯/(EijEkl)xosi532_e represents the component of the elasticity tensor at equilibrium state xo, which is specific to the material being characterized; refer to Table 9.7. The other coefficients can be described in a similar manner. With use of the nomenclature shown in Table 9.7, the free energy per unit volume (9.87) becomes

ψ¯¯=12CijklEijEkl12χeijeiej12χmijhihj12cΘ2deijkEijekdmijkEijhkβijEijΘχemijeihjLeiΘhi,

si533_e  (9.89)

and the linearized set of constitutive equations (9.88a)(9.88d) simplify to

Tij=CijklEkldeijkekdmijkhkβijΘ,

si534_e  (9.90a)

pi=dejkiEjk+χeijej+χemijhj+LeiΘ,

si535_e  (9.90b)

μomi=dmjkiEjk+χemjiej+χmijhj+LmiΘ,

si536_e  (9.90c)

η¯=βijEij+Leiei+Lmihi+cΘ.

si537_e  (9.90d)

Table 9.7

Material Constants and Their Representations for Linear Reversible Processes

ConstantRepresentationConstantRepresentation
Elasticity constantCijkl=2ψ¯¯EijEklxosi118_ePiezoelectric constantdeijk=2ψ¯¯Eijekxosi119_e
Piezomagnetic constantdmijk=2ψ¯¯Eijhkxosi120_eCoefficient of thermal stressβij=2ψ¯¯ΘEijxosi121_e
Electric susceptibilityχeij=2ψ¯¯eiejxosi122_eMagnetoelectric constantχemij=2ψ¯¯eihjxosi123_e
Pyroelectric constantLei=2ψ¯¯eiΘxosi124_eMagnetic susceptibilityχmij=2ψ¯¯hihjxosi125_e
Pyromagnetic constantLmi=2ψ¯¯hiΘxosi126_eSpecific heatc=2ψ¯¯Θ2xosi127_e

t0040

9.7.3 Material symmetry

The symmetric infinitesimal strains and stresses, as well as the existence of a free energy function ψ¯¯si538_e, lead to the following restrictions on the material constants in the constitutive equations (9.90a)(9.90d):

Cijkl=Cjikl=Cijlk=Cklij,deijk=dejik,dmijk=dmjik,χeij=χeji,χmij=χmji,βij=βji.

si539_e

These restrictions reduce the total number of unknown material constants to 169. The number of unknown material constants can be further reduced using crystal symmetry arguments. Materials that undergo one or more linear thermo-electro-mechanical processes (refer to Figure 9.2) can be classified into 32 crystallographic symmetry groups. These groups are based on rotation, reflection, and inversion symmetry of the crystal structure. A detailed description of each of these symmetry groups is provided in [63]. For magnetic materials, the concept of time inversion symmetry becomes an additional consideration, which increases the total number of possible symmetry groups from 32 to 122 (90 magnetic and 32 crystallographic symmetry groups).

f09-02-9780123946003
Figure 9.2 Multiphysics interaction diagram illustrating various linear TEMM effects.

Symmetry of the stress, strain, and material constant tensors allows further simplification of the constitutive equations (9.90a)(9.90d) using Voigt notation. Voigt notation is a standard mapping, typically used to reduce the order (or rank) of symmetric tensors. The indices are mapped as follows:

111,222,333,234,135,126.

si540_e

For example, Voigt notation simplifies the customary three-by-three matrix representation of the symmetric second-order stress tensor (refer, for instance, to (2.34)) to a single column matrix with six independent components. Similar reductions are accomplished for matrix representations of the fourth-order elasticity tensor, third-order piezoelectric and piezomagnetic coupling tensors, and second-order strain tensor. With use of this shorthand notation, the constitutive equations (9.90a)(9.90d), for the special case of a fully coupled TEME material with hexagonal crystal symmetry (i.e., the C6v crystallographic symmetry group), can be presented in matrix form as

T1T2T3T4T5T6p1p2p3μom1μom2μom3η¯=C11C12C1300000de1300dm13β11C12C11C1300000de1300dm13β11C13C13C3300000de3300dm33β33000C4400de41de5100dm51000000C440de51de410dm5100000000C660000000000de41de510χe1100χem11000000de51de4100χe1100χem1100de13de13de3300000χe3300χem33Le30000dm510χem1100χm11000000dm51000χem1100χm1100dm13dm13dm3300000χem3300χm33Lm3β11β11β3300000Le300Lm3cE1E2E3E4E5E6e1e2e3h1h2h3Θ

si541_e

where C66 = 1/2(C11C12). Clearly, crystal symmetry considerations greatly reduce the number of unknown material constants, which in turn reduces the number of experiments needed to completely characterize a material.

9.8 Linear, reversible, thermo-electro-magneto-mechanical processes

In this section, we discuss the wealth of physical phenomena and material behavior that can be described by the constitutive equations (9.90a)(9.90d) and characterized as linear, reversible, thermo-electro-magneto-mechanical (TEMM) processes. The multiphysics interaction diagram (see Figure 9.2) describes all combinations of linear TEMM processes. Each of the thermal, electrical, magnetic, and mechanical physical effects are defined by their corresponding extensive and intensive variables, marked at the inner and outer quadrilateral corners of the multiphysics interaction diagram, respectively.

The diagonal edges joining the inner and outer quadrilaterals signify the uncoupled processes, i.e., elasticity, polarization, magnetization, and heat capacity. Coupled processes are described through six subset diagrams, each relating two of the four physical effects. Each of the six subset diagrams, the corresponding coupled processes, and the materials that exhibit these properties are highlighted in Table 9.8; also see Figure 9.3.

Table 9.8

Subset Diagrams of the Fully Coupled TEMM Multiphysics Interaction Diagram

u09-01-9780123946003

t0045

f09-03-9780123946003
Figure 9.3 The thermomechanical panel, a subset of the multiphysics interaction diagram in Figure 9.2.

The coupled processes therein can be categorized as either (1) a primary process—a coupled process that relates the intensive parameter of one physical effect to the extensive parameter of the second physical effect—or (2) a secondary process—a coupled process that is a superposition of two or more primary processes. In other words, primary processes are direct or one-step processes that describe the coupling between any two physical effects, whereas secondary processes are multistep processes that are a superposition of two or more primary effects.

Owing to the linear nature of the constitutive model under consideration, any coupled TEMM effect can be studied as a superposition of the uncoupled and coupled primary processes highlighted in Table 9.8. For example, a linear thermo-electro-mechanical process can be described as the superposition of a linear thermoelectric (pyroelectric) process and a linear electromechanical (piezoelectric) process.

Depending on the smart material being modeled, appropriate terms can be chosen from equations (9.90a)(9.90d) to describe its behavior. This will be demonstrated in the next section for the special case of piezoelectric materials.

9.9 Specialization of the small-deformation thermo-electro-magneto-elastic framework to piezoelectric materials

As discussed in Section 9.7, ferroelectric materials inherently exhibit nonlinear hysteretic behavior. However, for small deformations and small electromagnetic fields, approximately linear responses are observed in materials such as barium titanate (BaTiO3), poly(vinylidene fluoride), and lead zirconate titanate. Ferroelectric materials like these that operate in a predominantly linear regime are known as piezoelectric materials.

Piezoelectric materials exhibit spontaneous polarization (at temperatures below the Curie point) in the presence of external electric fields. Below the Curie temperature, these materials exhibit a domain structure that lacks a center of symmetry, i.e., the centers of positive and negative charge are not identical. As a result, each unit cell acts as an electric dipole with a positive end and a negative end. Piezoelectrics change dimension in an electric field because the dipole length can be changed by the field: If a voltage is placed across the material, the dipoles respond to the field and change their dipole length, thereby changing the dimension of the crystal. Alternatively, if the crystal is mechanically stretched or compressed, the length of the dipole is changed, creating a voltage difference if there is no conductive path between the two ends of the dipole.

Since a necessary condition for the occurrence of piezoelectricity is the absence of a center of symmetry, piezoelectric materials are intrinsically anisotropic. Since piezoelectricity couples elasticity and polarization, piezoelectric material properties cannot be discussed without reference to the elasticity constant and the electric susceptibility (or electric permittivity); refer to Table 9.7. In what follows, we specialize the linear thermo-electro-magneto-elastic (TEME) framework developed in Section 9.7 to model piezoelectric material behavior. We make the following assumptions for a piezoelectric material operating well below the Curie temperature:

(1) Piezoelectric materials exhibit strains on the order of 10-100 microstrain. Infinitesimal strain theory can thus be used to describe the kinematics.

(2) Piezoelectric materials are used in transducer applications that operate in the low-frequency regime, typically ranging from 10 to 500 Hz. For this range of frequencies, the dynamic behavior of the electromagnetic fields may be ignored. In other words, the electromagnetic fields may be regarded as quasi-static.

(3) Thermal and magnetic effects may be neglected.

(4) The medium is nonconductive. Thus, when an external voltage is applied to the medium, no charge distribution is formed, i.e., σi = 0, and there is no free current, i.e., ji = 0.

These assumptions allow us to simplify the fundamental laws (9.84a)(9.84j) to

Tij,j+ρfmi=ρui,tt,

si542_e  (9.91a)

Tij=Tji,

si543_e  (9.91b)

di,i=0,

si544_e  (9.91c)

ei=,i,

si545_e  (9.91d)

where ø is the electric potential. Note that (9.91d) is a consequence of Faraday's law: in the absence of time-varying magnetic fields, Faraday's law demands that the electric field is curl free, which, in turn, implies that the electric field is the gradient of a scalar potential.

Similarly, the linear constitutive equations (9.90a)(9.90d) reduce to

Tij=CijklEkldeijkek

si546_e  (9.92a)

pi=dejkiEjk+χeijej,

si547_e  (9.92b)

where

Eij=12(ui,j+ui,j),di=εoei+pi.

si548_e

Note that the constitutive equations (9.92a) and (9.92b) relate stress and electric polarization to strain and electric field. Utilizing Voigt notation and imposing hexagonal crystal symmetry (C6v group), we can express (9.92a) and (9.92b) in matrix form as

T1T2T3T4T5T6p1p2p3=C11C12C1300000de13C12C11C1300000de13C13C13C3300000de33000C4400de41de5100000C440de51de41000000C66000000de41de510χe1100000de51de4100χe110de13de13de3300000χe33

si549_e

Exercise

1. Derive a linear reversible framework for the constitutive modeling of magnetostrictive materials, i.e., materials that couple magnetic and mechanical fields. Use the magnetic field h and stress T as the independent variables, and the magnetization m and strain E as the dependent variables. Disregard thermal and electrical effects.

References

[30] Green AE, Naghdi PM. A unified procedure for construction of theories of deformable media. I. Classical continuum physics. Proc. R. Soc. Lond. A. 1995;448(1934):335–356.

[31] Green AE, Naghdi PM. A re-examination of the basic postulates of thermomechanics. Proc. R. Soc. Lond. A. 1991;432(1885):171–194.

[32] Hutter K, van de Ven AAF, Ursescu A. Electromagnetic Field Matter Interactions in Thermoelastic Solids and Viscous Fluids. Berlin: Springer-Verlag; 2006.

[33] Pao Y-H. Electromagnetic forces in deformable continua. In: Nemat-Nasser S, ed. New York: Pergamon Press; 209–305. Mechanics Today. 1978;vol. 4.

[34] Kovetz A. Electromagnetic Theory. Oxford: Oxford University Press; 2000.

[35] Kankanala SV, Triantafyllidis N. On finitely strained magnetorheological elastomers. J. Mech. Phys. Solids. 2004;52(12):2869–2908.

[36] Dorfmann A, Ogden RW. Some problems in nonlinear magnetoelasticity. Zeit. Ang. Math. Phys. 2005;56(4):718–745.

[37] Dorfmann A, Ogden RW. Nonlinear electroelastic deformations. J. Elast. 2006;82(2):99–127.

[38] Eringen AC, Maugin GA. Electrodynamics of Continua I: Foundations and Solid Media. New York: Springer-Verlag; 1990.

[39] Lorentz HA. The Theory of Electrons. second ed. New York: Dover; 2004.

[40] de Groot SR, Suttorp LG. Foundations of Electrodynamics. Amsterdam: North-Holland Publishing Company; 1972.

[41] Fano RM, Chu LJ, Adler RB. Electromagnetic Fields, Energy, and Forces. Cambridge: The MIT Press; 1968.

[42] Maxwell JC. A Treatise on Electricity and Magnetism. Oxford: Clarendon Press; 1873.

[43] Brown WF. Magnetoelastic Interactions. New York: Springer-Verlag; 1966.

[44] Green AE, Naghdi PM. Aspects of the second law of thermodynamics in the presence of electromagnetic effects. Quart. J. Mech. Appl. Math. 1984;37(2):179–193.

[45] Dorfmann A, Ogden RW. Magnetoelastic modelling of elastomers. Euro. J. Mech. A: Solids. 2003;22(4):497–507.

[46] Dorfmann A, Ogden RW. Nonlinear magnetoelastic deformations. Quart. J. Mech. Appl. Math. 2004;57(4):599–622.

[47] Dorfmann A, Ogden RW. Nonlinear magnetoelastic deformations of elastomers. Acta Mech. 2004;167(1-2):13–28.

[48] Dorfmann A, Ogden RW. Nonlinear electroelasticity. Acta Mech. 2005;174(3-4):167–183.

[49] Steigmann DJ. Equilibrium theory for magnetic elastomers and magnetoelastic membranes. Int. J. Non-Linear Mech. 2004;39(7):1193–1216.

[50] McMeeking RM, Landis CM. Electrostatic forces and stored energy for deformable dielectric materials. ASME J. Appl. Mech. 2005;72(4):581–590.

[51] McMeeking RM, Landis CM, Jimenez SMA. A principle of virtual work for combined electrostatic and mechanical loading of materials. Int. J. Non-Linear Mech. 2007;42(6):831–838.

[52] Vu DK, Steinmann P, Possart G. Numerical modelling of non-linear electroelasticity. Int. J. Numer. Meth. Eng. 2007;70(6):685–704.

[53] Voltairas PA, Fotiadis DI, Massalas CV. A theoretical study of the hyperelasticity of electro-gels. Proc. R. Soc. Lond. A. 2003;459(2037):2121–2130.

[54] Suo Z, Zhao X, Greene WH. A nonlinear field theory of deformable dielectrics. J. Mech. Phys. Solids. 2008;56(2):467–486.

[55] Zhao X, Suo Z. Electrostriction in elastic dielectrics undergoing large deformation. J. Appl. Phys. 2008;104(12):123530.

[56] Zhu J, Stoyanov H, Kofod G, Suo Z. Large deformation and electromechanical instability of a dielectric elastomer tube actuator. J. Appl. Phys. 2010;108(7):074113.

[57] Qin Z, Librescu L, Hasanyan D, Ambur DR. Magnetoelastic modeling of circular cylindrical shells immersed in a magnetic field. Part I: Magnetoelastic loads considering finite dimensional effects. Int. J. Eng. Sci. 2003;41(17):2005–2022.

[58] Qin Z, Hasanyan D, Librescu L, Ambur DR. Magnetoelastic modeling of circular cylindrical shells immersed in a magnetic field. Part II: Implications of finite dimensional effects on the free vibrations. Int. J. Eng. Sci. 2003;41(17):2023–2046.

[59] Richards AW, Odegard GM. Constitutive modeling of electrostrictive polymers using a hyperelasticity-based approach. ASME J. Appl. Mech. 2010;77(1):014502.

[60] Oates WS, Wang H, Sierakowski RL. Unusual field-coupled nonlinear continuum mechanics of smart materials. J. Intel. Mater. Syst. Struct. 2012;23(5):487–504.

[61] Callen HB. Thermodynamics and an Introduction to Thermostatistics. second ed. New York: Wiley; 1985.

[62] Santapuri S, Lowe RL, Bechtel SE, Dapino MJ. Thermodynamic modeling of fully coupled finite-deformation thermo-electro-magneto-mechanical behavior for multifunctional applications. Int. J. Eng. Sci. 2013;72:117–139.

[63] Newnham RE. Properties of Materials: Anisotropy, Symmetry, Structure. New York: Oxford University Press; 2005.


star “To view the full reference list for the book, click here

1 A familiar example of this notion is a truss in rigid body statics: not only is the entire structure in equilibrium, but also each joint and each member.

2 In order to avoid notational conflicts later in the chapter, we change the notation for the body from Bsi221_e (as was used in previous chapters) to Bsi222_e.

3 Refer to Section 3.5 for a discussion of material curves, material surfaces, and material volumes.

4 Recall from Section 4.9 that we are free to label the volume occupied by subset S1si253_e by its present volume Psi254_e or its reference volume PRsi255_e.

5 In order to avoid notational conflicts later in the chapter, we change the notation for the referential traction from p (as was used in previous chapters) to tR, and the notation for the mechanical body force from b (as was used in previous chapters) to fm.

6 In this chapter, we change the notation for the heat supply rate from r (as was used in previous chapters) to rt.

7 When validating the dimensional homogeneity of an equation in the finite-deformation theory, one finds it useful to differentiate the fundamental units of length from one another, with LP corresponding to the present configuration and LR corresponding to the reference configuration.

8 A constraint on the mathematical forms of these transformations is that the effective fields reduce to the standard fields in the absence of motion. That is, when v = 0, then e* should collapse to e, d* should collapse to d, and so on.

9 In this division, the magnetic flux b*, electric displacement d*, and Cauchy stress T are relegated to secondary dependent variables, i.e., variables that can be calculated algebraically from the independent and primary dependent variables in (9.60). Also note that v, F, and L can be calculated from x using tensor calculus.

10 In classical thermodynamics, extensive quantities are properties of a thermodynamic system that depend on its size or quantity, whereas intensive quantities are independent of its size or quantity. This intensive-extensive terminology is akin to the notion of generalized thermodynamic force-displacement pairs, with each such pair contributing to the internal energy of the system. In this chapter, we extend these classical thermodynamic concepts to the analogous setting of continuum thermodynamics [62].

11 Although we derive our constitutive framework assuming that the system is in thermodynamic equilibrium (i.e., it only undergoes reversible processes), our framework can be used to model irreversible systems that operate in a regime close to equilibrium [61].

12 The assumption of small displacements implies that both strains and rotations are small.

13 The leading-order terms within a mathematical equation, expression, or model are the terms with the largest magnitude. For instance, in the small-deformation theory, F=I+Q(ε)si504_e, where ε is a small parameter and Q(ε)si505_e denotes the higher-order (or lower-magnitude) terms.

14 Note the common notation e for the Eulerian strain tensor and the electric field. This should not lead to any ambiguity, however, as this is the first and last time the Eulerian strain tensor appears in this chapter.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset