Problem 9.4

Prove the transport theorem for surface integrals. That is, show that

ddtQa·nda=Q[a+curl(a×v)+v(diva)]·nda.

si164_e

Solution

We begin by using relationship (3.76) between n da and N dA, together with a change of independent variable from x to X, to convert the Eulerian integration to a Lagrangian integration, i.e.,

ddtQa·nda=ddtQRa·JFTNdA.

si165_e

We emphasize that the integrand on the left-hand side is a function of x and t, while the integrand on the right-hand side is a function of X and t. Then, it follows that

ddtQa·nda=ddtQRJF1a·NdA(definition(2.13))=QR·¯JF1a·NdA(fixed region of integrationQR)=QR(˙JF1a+J·¯F1a+JF1˙a)·NdA(product rule(3.22)).

si166_e

We now individually examine the three terms on the right-hand side of the above equation. For the first term, we have

QR˙JF1a·NdA=QRJ(divv)F1a·NdA(result(3.60)3)=QRa(divv)·JFTNdA(definition(2.13)).

si167_e

For the second term, we have

QRJ·¯F1a·NdA=QRJF1La·NdA(result(3.61)2)=QRLa·JF1NdA(definition(2.13))=QR(gradv)a·JFTNdA(definition(3.56)).

si168_e

For the third term, we have

QRJF1˙a·NdA=QR˙a·JFTNdA(definition(2.13)).

si169_e

Assembling the preceding results yields

ddtQa·nda=QR(˙a+a(divv)(gradv)a)·JFTNdA.

si170_e

We convert the Lagrangian integration to an Eulerian integration with a change of independent variable from X to x and use of the relationship (3.76) between N dA and n da:

ddtQa·nda=Q(˙a+a(divv)(gradv)a)·nda.

si171_e

Use of results (2.102) and (3.20)2 leads to

ddtQa·nda=Q(a+curl(a×v)+v(diva))·nda.

si172_e

Problem 9.5

Starting with the Eulerian (or spatial) statement of the Ampère-Maxwell law in integral form,

ddtQd*·nda+Qj*·nda=Qh*·ldl,

si173_e

derive the corresponding pointwise form

curlh*=(d*)+curl(d*×v)+σ*v+j*.

si174_e

Solution

We begin with the Eulerian integral form of the Ampère-Maxwell law, i.e.,

ddtQd*·nda+Qj*·nda=Qh*·ldl.

si175_e  (a)

We consider the first term on the left-hand side of (a):

ddtQd*·nda=Q[(d*)+curl(d*×v)+v(divd*)]·nda(transport theorem(9.43))=Q[(d*)+curl(d*×v)+σ*v]·nda(Gauss's law(9.46i)).

si176_e

Next, we consider the right-hand side of (a), which, after use of Stokes's theorem (9.44), becomes

Qh*·ldl=Q(curlh*)·nda.

si177_e

Substitution of the preceding results into (a) leads to

Q[(d*)+curl(d*×v)+σ*v+j*curlh*]·nda=0.

si178_e

Since the integrand is continuous and Qsi179_e is arbitrary, the localization theorem implies that

[(d*)+curl(d*×v)+σ*v+j*curlh*]·n=0.

si180_e

Furthermore, since the coefficient of n is independent of n, and n is arbitrary, it follows that the coefficient must vanish, i.e.,

curlh*=(d*)+curl(d*×v)+σ*v+j*.

si181_e

It is important to note that the electromagnetic equations (9.46f)(9.46j) are not independent. For instance, it can be shown that conservation of charge (9.46f) is implicitly contained in Maxwell's equations (9.46g)(9.46j); refer to Problem 9.6.

Problem 9.6

Show that conservation of charge (9.46f) is a consequence of Maxwell's equations (9.46g)(9.46j).

Solution

Taking the divergence of the Ampère-Maxwell law (9.46j), and subsequently using result (2.101), gives

div[(d*)]+div(σ*v)+divj*=0.

si182_e

Continuity of d* enables the order of partial differentiation to be exchanged, so

div[(d*)]=(divd*).

si183_e

This, together with result (2.99)2, implies that

(divd*)+v·gradσ*+σ*divv+divj*=0.

si184_e

We then invoke Gauss's law for electricity (9.46i) and relationship (3.20)1 to recover

˙σ*+σ*divv+divj*=0,

si185_e

the pointwise Eulerian statement of conservation of charge.

Exercises

1. Verify that t = Tn.

2. Confirm that h = q · n.

3. Prove that (9.46a)(9.46i) are the pointwise versions of the Eulerian integral forms of the fundamental laws (9.42a)(9.42i).

9.2.2 Lagrangian fundamental laws

Recall from Section 9.1 the Lagrangian integral forms of the first principles:

Conservation of mass

Pρdv=PRρRdV,

si403_e  (9.48a)

Balance of linear momentum

ddtPRvρRdV=PR(fm+fem)ρRdV+PRtRdA,

si404_e  (9.48b)

Balance of angular momentum

ddtPRx×vρRdV=PRx×(fm+fem)ρRdV+PRx×tRdA+PRcemρRdV,

si405_e  (9.48c)

First law of thermodynamics

ddtPR12v·vρRdV+ddtPRερRdV=PR(fm+fem)·vρRdV+PRtR·vdA+PR(rt+rem)ρRdVPRhRdA,

si406_e  (9.48d)

Second law of thermodynamics

ddtPRηρRdVPRrtΘρRdVPRhRΘdA,

si407_e  (9.48e)

Conservation of electric charge

ddtPRσRdV=PRjR·NdA,

si408_e  (9.48f)

Gauss's law for magnetism

PRbR·NdA=0,

si409_e  (9.48g)

Faraday's law

ddtQRbR·NdA=QReR·lRdL,

si410_e  (9.48h)

Gauss's law for electricity

PRdR·NdA=PRσRdV,

si411_e  (9.48i)

Ampère-Maxwell law

ddtQRdR·NdA+QRjR·NdA=QRhR·lRdL.

si412_e  (9.48j)

An important feature of the Lagrangian integral conservation laws (9.48a)(9.48j) is that the regions PRsi413_e and QRsi414_e occupied by subsets S1si415_e and S2si416_e in the reference configuration are fixed. Hence, the regions of integration PRsi417_e and QRsi418_e do not change with time, so time derivatives of Lagrangian surface and volume integrals can be passed directly inside the integrals. Conversely, with the Eulerian integral conservation laws (9.42a)(9.42j), the regions of integration Psi419_e and Qsi420_e change with time. Thus, to take time derivatives of these time-varying integrals, the transport theorem for surface integrals (9.43) and the transport theorem for volume integrals (4.20) must be employed. The transport theorems are analogous to Leibniz's rule in multivariable calculus.

The continuous, bounded nature of the integrands in (9.48a)(9.48j) enables the divergence theorem and Stokes's theorem, and the requirement that (9.48a)(9.48j) be continuous and global, i.e., true for the entire body and all subsets, enables the localization theorem. With the traction tR and heat flux hR dependent on surface geometry only through the outward unit normal N, so tR = PN and hR = qR · N, it can be shown that application of the divergence theorem, Stokes's theorem, and the localization theorem to the Lagrangian integral equations (9.48a)(9.48j) leads to the Lagrangian pointwise equations

ρJ=ρR,

si421_e  (9.49a)

ρR˙v=ρR(fm+fem)+DivP,

si422_e  (9.49b)

ρRΓem+PFTFPT=0,

si423_e  (9.49c)

ρR˙ε=P·Gradv+ρR(rt+rem)DivqR,

si424_e  (9.49d)

ρR˙ηρRrtΘDiv(qRΘ),

si425_e  (9.49e)

˙σR+DivjR=0,

si426_e  (9.49f)

DivbR=0,

si427_e  (9.49g)

CurleR=˙bR,

si428_e  (9.49h)

DivdR=σR,

si429_e  (9.49i)

CurlhR=˙dR+jR.

si430_e  (9.49j)

Note that it can be shown that conservation of charge (9.49f) is not independent of Maxwell's equations(9.49g)(9.49j), but rather is implicitly contained in them. Recall that P is the first Piola-Kirchhoff stress, qR is the referential heat flux vector, “Div” is the Lagrangian divergence (i.e., the divergence calculated with respect to the reference configuration), “Curl” is the Lagrangian curl, and

˙a=tˆa(X,t)

si431_e

is the material derivative of an arbitrary vector a=ˆa(X,t)si432_e, i.e., the partial derivative of the referential description of a with respect to time t. Also,

Γema=cem×a

si433_e

for any vector a, i.e., Γem is a skew tensor whose corresponding axial vector is cem. Note that in deriving the pointwise version of the first law of thermodynamics (9.49d) from its integral counterpart (9.48d), we have made use of the Lagrangian form of the energy theorem for continuum electrodynamics:

PR(fm+fem)·vρRdV+PRtR·vdAddtPR12v·vρRdV=PRP·GradvdV.

si434_e  (9.50)

Compare (9.50) with (4.54), the Lagrangian form of the energy theorem for mechanics.

Exercises

1. Verify that tR = PN.

2. Confirm that hR = qR · N.

3. Prove that (9.49a)(9.49j) are the pointwise versions of the Lagrangian integral forms of the fundamental laws (9.48a)(9.48j).

4. Show that conservation of charge (9.49f) is a consequence of Maxwell's equations (9.49g)(9.49j).

9.3 Modeling of the effective electromagnetic fields

Recall that Eulerian modeling of deformable thermo-electro-magneto-mechanical materials involves two sets of spatial (or Eulerian) electromagnetic fields: the effective fields e*, d*, p*, h*, b*, m*, σ*, and j*, distinguished in our notation by superscript asterisks, and the standard fields e, d, p, h, b, m, σ, and j. (Conversely, in Lagrangian modeling, there is no notion of effective or standard fields.) The effective fields are the electromagnetic fields acting on the deforming continuum as seen in its present configuration, measured with respect to a co-moving or rest frame, i.e., a frame affixed to but not deforming with the continuum. The standard fields also act on the deforming continuum as seen in its present configuration, but are instead measured with respect to a fixed or laboratory frame.

Various transformations have been presented in the literature that relate the effective fields to the standard fields.8 In this section, we present four popular transformation theories, each based on a different set of principles and postulates. As a result, the standard electromagnetic fields generally have different physical connotations from model to model; hence, in what follows, the fields corresponding to a particular model are distinguished with the appropriate subscript notation.

9.3.1 Minkowski model

The Minkowski model [32, 33, 38] is motivated by Einstein's special theory of relativity. In this approximation, the effective fields are related to the standard fields through semirelativistic inverse Lorentz transformations:

e*=eM+v×bM,h*=hMv×dM,d*=dM,b*=bM,p*=pM,m*=mM+v×pM,j=jMσMv,σ*=σM,

si435_e  (9.51)

where (·)M represents a standard electromagnetic field corresponding to the Minkowski model.

9.3.2 Lorentz model

The Lorentz model is motivated by Lorentz's theory of electrons [39], which postulates that a body consists of an infinitely large number of rapidly moving charged particles. The motion of these charged particles, in turn, generates rapidly fluctuating microscopic electromagnetic fields. The microscopic fields averaged over a small time interval and an infinitesimal volume are defined as the corresponding macroscopic fields. The aforementioned postulates lead to the following relationships:

e*=eL+v×bL,h*=hLov×eL,d*=dL,b*=bL,p*=pL,m*=mL,j*=jLσLv,σ*=σL,

si436_e  (9.52)

where (·)L represents a standard electromagnetic field corresponding to the Lorentz model.

9.3.3 Statistical model

The statistical model [40] is a modification of Lorentz's theory, wherein the charged particles are grouped into stable structures (e.g., atoms, molecules, or ions). The field effects of the charged particles within each stable structure are represented by microscopic electric and magnetic multipole moments (e.g., dipole, quadrupole, and octupole moments), and the macroscopic polarization and magnetization fields are defined as statistical averages of these multipole moments over a large number of stable structures. The transformations presented in this model are identical to Minkowski's, i.e.,

e*=eS+v×bS,h*=hSv×dS,d*=dS,b*=bS,p*=pS,m*=mS+v×pS,j*=jSσSv,σ*=σS,

si437_e  (9.53)

where ( )s represents a standard electromagnetic field corresponding to the statistical model.

9.3.4 CHU model

The Chu model [41] is based on the postulate that deforming bodies contribute to electromagnetic phenomena by acting, in a macroscopic sense, as electric and magnetic dipole sources for the electromagnetic fields. The transformations for the Chu model are

e*=eC+μO×hC,h*=hCov×eC,d*=dC,b*=bC,p*=pC+mC×vc2,m*=mC,j*=jCσCv,σ*=σC,

si438_e  (9.54)

where ( )c represents a standard electromagnetic field corresponding to the Chu model.

9.3.5 A comparison of the four models

Table 9.2 catalogs Maxwell's equations for each of the four models, deduced by substituting each set of transformations (i.e., (9.51)(9.54)) into (9.46g)(9.46j). Recall that since each of the four models is based on a different set of postulates, the standard electromagnetic fields generally have different physical connotations from model to model. We can deduce relationships between the standard fields by comparing the equations in Table 9.2. For instance, the Minkowski and statistical models, although developed from different perspectives, are mathematically equivalent (i.e., mM = ms, pM = ps, etc.).

Table 9.2

Maxwell's Equations for Different Transformation Models

ModelMaxwell's Equations
Minkowskicurl eM = −bMdiv dM = σM
curl hM = dM + jMdiv bM = 0
Lorentzcurl eL = −bLdiv dL = σL
curl hL = dL − curl (v × pL) + jLdiv bL = 0
Statisticalcurl eS = −bSdiv dS = σS
curl hS = dS + jSdiv bS = 0
Chucurl eC = −bC + μocurl (v × mC)div dC = σC
curl hC = dC − curl (v × pC) + jCdiv bC = 0

t0015

Problem 9.7

Verify the results presented in Table 9.2 for the Minkowski model.

Solution

Straightforward substitution of the Minkowski transformations (9.51)3, (9.51)4, and (9.51)8 into the Eulerian forms of Gauss's law for magnetism (9.46g) and Gauss's law for electricity (9.46i) gives

divbM=0,divdM=σM.

si439_e

Then, use of the transformation equations (9.51)2, (9.51)3, (9.51)7, and (9.51)8, i.e.,

h*=hMv×dM,d*=dM,j*=jMσMv,σ*=σM,

si440_e

in the Ampère-Maxwell law (9.46j) leads to

curl(hMv×dM)=dM+curl(dM×v)+σMv+(jMσMv).

si441_e

Anticommutativity of the vector product (2.61) and the distributive property of the curl (2.103a) then imply that the above equation becomes

curlhM=dM+jM.

si442_e

Similarly, Faraday's law (9.46h) is transformed by applying (9.51)1 and (9.51)4:

curl(eM+v×bM)=bMcurl(bM×v).

si443_e

Simplification of this result using (2.61) and (2.103a) leads to

curleM=bM.

si444_e

Exercise

1. Verify the results presented in Table 9.2 for the Lorentz, statistical, and chu models.

9.4 Modeling of the electromagnetically induced coupling terms

As discussed in Section 9.1, the thermomechanical equations ((9.46a)(9.46d) in Eulerian form) are coupled to the electromagnetic equations ((9.46f)(9.46j) in Eulerian form) through the electromagnetic body force fem, body couple cem (or, equivalently, Γem), and energy supply rate rem. The mathematical forms of these coupling terms are postulated on the basis of the interaction theories presented in Section 9.3, which are motivated by either atomic physics or empiricism. As an example, the coupling terms postulated using the Minkowski theory for apolarizable, magnetizable, deformable continuum are, in Eulerian form [32],

ρfem=σ*e*+j*×b*+(grade*)Tp*+μo(gradh*)Tm*+d*×b*+d*×b*,

si445_e  (9.55a)

ρΓem=(e*p*p*e*)+μo(h*m*m*h*),

si446_e  (9.55b)

ρrem=j*·e*+ρe*··¯(p*ρ)+ρμoh*··¯(m*ρ),

si447_e  (9.55c)

where we recall that () ⊗ () denotes the dyadic product of two vectors, “grad” is the Eulerian gradient, and

å=a+curl(a×v)+v(diva)

si448_e  (9.56)

is a convected rate of an arbitrary vector a=˜a(x,t)si449_e. coupling terms for the Lagrangian forms of the fundamental laws can be found, for instance, in [32].

The Minkowski model (9.55a)(9.55c) has as special cases interaction theories describing forces exerted by static electric fields in polarizable solids [42] and static magnetic fields in magnetizable solids [43]. Use of (9.55a)(9.55c) in the Eulerian forms of balance of linear momentum (9.46b), balance of angular momentum (9.46c), and the first law of thermodynamics (9.46d) yields

ρ˙v=ρfm+σ*e*+j*×b*+(grade*)Tp*+μo(gradh*)Tm*+d*×b*+d*×b*+divT,

si450_e  (9.57a)

TTT=(p*e*e*p*)+μo(m*h*h*m*),

si451_e  (9.57b)

˙ε=1ρRP·˙F+e*··¯(p*ρ)+μoh*··¯(m*ρ)+rt+1ρj*·e*1ρdivq.

si452_e  (9.57c)

9.4.1 An alternative approach

Electromagnetic effects in the thermomechanical balance laws can be modeled as an electromagnetically induced body force, body couple, and energy supply (as discussed previously; see also [32, 33, 44]) or, alternatively, incorporated into the constitutive response (see, for instance, [3436]). The latter is accomplished by defining a symmetrictotal stress tensor

τ=T+Tem

si453_e  (9.58)

that consists of contributions from the cauchy stress tensor T and the electromagnetic Maxwell stress tensor Tem. The Maxwell stress Tem is defined so that twice its skew part is ρΓem and its divergence is ρfem; refer to Problem 9.8. For the Minkowski formulation (9.55a)(9.55c), it can be shown that

Tem=e*d*+h*b*12(εoe*·e*+μoh*·h*)I.

si454_e  (9.59)

By formulating the first principles and the constitutive equations in terms of the total stress τ instead of the Cauchy stress T, we eliminate explicit coupling between the electromagnetic fields and the thermomechanical fields in the first principles (9.57a)(9.57c). This coupling is instead accounted for implicitly through the constitutive equation for τ.

Problem 9.8

Prove that the divergence of the Maxwell stress tensor Tem in (9.59) is equal to the electromagnetic body force vector ρfem in (9.55a).

Solution

Taking the divergence of the Maxwell stress tensor (9.59) gives

divTem=div[e*d*+h*b*12(εoe*·e*+μoh*·h*)I].

si186_e

The divergence is distributive over tensor addition (refer to (2.103b)), which implies that

divTem=div(e*d*)+div(h*b*)=12{εodiv[(e*·e*)I]+μodiv[(h*·h*)I]}.

si187_e  (a)

The first term in (a) simplifies to

div(e*d*)=(divd*)e*+(grade*)d*(result(2.99)7)=σ*e*+(grade*)d*(Gauss's law(9.46i)).

si188_e  (b)

Similarly, the second term reduces to

div(h*b*)=(divb*)h*+(gradh*)b*(result(2.99)7)=(gradh*)b*(Gauss's law(9.46g)).

si189_e  (c)

The third term in (a) simplifies to

12εodiv[(e*·e*)I]=12εograd(e*·e*)(result(2.99)6)=εo(grade*)Te*(result(2.99)3).

si190_e  (d)

Similarly, use of results (2.99)3 and (2.99)6 reduces the fourth term to

12μodiv[(h*·h*)I]=μo(gradh*)Th*.

si191_e  (e)

Combining results (b)(e) in (a), we obtain

divTem=σ*e*+(grade*)d*+(gradh*)b*=[εo(grade*)Te*+μo(gradh*)Th*].

si192_e  (f)

Relationships (9.34)1 and (9.40)1, i.e.,

d*=p*+εoe*,b*=μo(h*+m*),

si193_e

allow us to rewrite (f) as

divTem=σ*e*+[grade*(grade*)T]d*+[gradh*=(gradh*)T]b*+(grade*)Tp*+μo(gradh*)Tm*.

si194_e

Invoking definition (2.100) of the curl implies that

divTem=σ*e*+(curle*)×d*+(curlh*)×b*+(grade*)Tp*+μo(gradh*)Tm*.

si195_e  (g)

Faraday's law (9.46h) and the Ampère-Maxwell law (9.46j), i.e.,

curle*=(b*)curl(b*×v),curlh*=(d*)+curl(d*×v)+σ*v+j*,

si196_e

are then used in (g), which, noting anticommutativity of the vector product, leads to

divTem=σ*e*+j*×b*+(grade*)Tp*+μo(gradh*)Tm*+[(d*)+curl(d*×v)+σ*v]×b*+d*×[(b*)+curl(b*×v)]

si197_e

It then follows that

divTem=σ*e*+j*×b*+(grade*)Tp*+μo(gradh*)Tm*+d*×b*+d*×b*=ρfem,

si198_e

where b*si199_e and d*si200_e represent convected rates as defined in (9.56).

Exercises

1. Prove that T·L=1JP·˙Fsi455_e.

2. Verify that use of the electromagnetic energy (9.55c) in the Eulerian form of the first law of thermodynamics (9.46d) leads to (9.57c).

3. Verify that

(a) twice the skew part of the Maxwell stress tensor Tem in (9.59) is ρΓem in (9.55b), and

(b) the total stress tensor τ in (9.58) is symmetric.

9.5 Thermo-electro-magneto-mechanical process

The pointwise Eulerian field equations (9.46a), (9.46f)(9.46j), and (9.57a)(9.57c) constitute the first principles of our model, true for all thermo-electro-magneto-mechanical (TEMM) materials. For the purposes of developing the constitutive equations that supplement these first principles and characterize particular TEMM materials, we conceptually divide the fields appearing in (9.46a), (9.46f)(9.46j), and (9.57a)(9.57c) into three groups:

{x,η,p*,m*},{P,Θ,e*,h*,ε,q,j*},{ρ,fm,rt,σ*},

si456_e  (9.60)

denoted the independent variables, dependent variables, and balancing terms, respectively.9 (Note that the set of independent variables contains slots occupied by one mechanical, one thermal, one electrical, and one magnetic quantity, respectively, from left to right.) Hence, the first Piola-Kirchhoff stress P, temperature Θ, electric field e*, magnetic field h*, internal energy ε, heat flux q, and conductive current j* are determined from constitutive equations that, in general, depend on the history of the motion x, entropy η, electric polarization p*, and magnetization m*, and possibly their rates or gradients. A group of quantities x, η, p*, m*, P, Θ, e*, h*, ε, q, j*, ρ, fm, rt, and σ* that satisfy the governing equations (9.46a), (9.46f)(9.46j), and (9.57a)(9.57c) for all space and time in the domain of interest describes a TEMM process.

9.6 Constitutive model development for thermo-electro-magneto-elastic materials: large-deformation theory

In this section, we specialize to (1) thermo-electro-magneto-elastic (TEME) materials that are capable of undergoing large elastic strains before yielding and (2) material response that is reversible (i.e., path independent and rate insensitive). We then illustrate the development of a general finite-deformation TEME constitutive framework using the principles of continuum thermodynamics, and show how this framework can be simplified using restrictions imposed by the second law of thermodynamics, invariance, angular momentum, and material symmetry. Similar approaches have been used to develop constitutive models for particular classes of smart materials in the finite-deformation regime, e.g., electroactive elastomers and magnetosensitive elastomers; see, for instance, [3537,4560,]. The large-deformation constitutive framework presented in this section, on the other hand, is intended to model a more general range of smart material behavior (e.g., magneto-electric coupling).

9.6.1 The reduced clausius-duhem inequality, work conjugates

We algebraically combine the Eulerian forms of the Clausius-Duhem inequality (9.46e) and the first law of thermodynamics (9.57c) to produce the reduced Clausius-Duhem inequality

˙ε+1ρRP·˙F+Θ˙η+e*··¯(p*ρ)+μoh*··¯(m*ρ)+1ρj*·e*1ρΘq·gradΘ0.

si457_e  (9.61)

Analogously to classical thermodynamics [61], this fundamental statement of the second law consists of contributions from conjugate pairs of thermal, electrical, magnetic, and mechanical quantities. Each of these conjugate pairs (or work conjugates) is the product of the rate of an extensive quantity in rate form (˙Fsi458_e, ˙ηsi459_e, (˙p*/ρ)si460_e, and (˙m*/ρ)si461_e) and an intensive quantity (P, Θ, e*, and h*).10 The utility of the second law, and our identification of the extensive and intensive quantities, will become apparent in the following section.

Exercise

1. Verify that the Clausius-Duhem inequality (9.46e) can be algebraically combined with the first law of thermodynamics (9.57c) to produce the reduced Clausius-Duhem inequality (9.61).

9.6.2 The all-extensive formulation

Recall that we specialized to reversible TEME material response, which implies that the deformation, although it may be large, is elastic and fully recoverable, and the material only undergoes nondissipative TEMM processes. It follows, then, that these TEMM processes can be completely described through the principles of classical equilibrium thermodynamics11 [61].

Analogously to classical equilibrium thermodynamics, our fundamental energy potential is the specific internal energy ε, which employs extensive quantities as its independent variables [61]. Hence, for the fundamental formulation, the natural independent variables are the extensive quantities F, η, p*/ρ, and m*/ρ appearing as rates in (9.61), and the natural dependent variables are the conjugate intensive quantities P, Θ, e*, and h*. Thus, for thermodynamic consistency, and to respect the reversible elastic nature of the process, we adjust the division (9.60) so that the dependence of the response on the present position x, polarization p*, and magnetization m* is through the deformation gradient F = Grad x, p*/ρ, and m*/ρ, respectively. Additionally, we demand that the response depends on these independent variables only through their values at the present time t, not their histories, rates, or gradients, i.e.,

P=˘P(F,η,p*ρ,m*ρ),Θ=˘Θ(F,η,p*ρ,m*ρ),e*=˘e*(F,η,p*ρ,m*ρ),h*=˘h*(F,η,p*ρ,m*ρ),ε=˘ε(F,η,p*ρ,m*ρ).

si462_e  (9.62)

Notation (9.62) indicates that P, Θ, e*, h*, and ε are explicit functions of F, η, p*/p, and m*/ρ evaluated at present position x and time t, and implicit functions of x and t. Note that a superscript breve is used to distinguish a function from its value (or output). As with all materials in the TEMM theory, the response functions (9.62) for a TEME material must satisfy the second law of thermodynamics, invariance requirements, conservation of angular momentum, and material symmetry conditions.

In what follows, we demonstrate how restrictions imposed by the second law of thermodynamics (9.61) yield a set of constitutive equations that provide the dependent variables P, Θ, e*, and h* (the intensive quantities) as partial derivatives of the specific internal energy ε (the fundamental thermodynamic energy potential) with respect to the independent variables F, η, p*/ρ, and m*/ρ (the extensive quantities), respectively. Use of the chain rule on ε=˘ε(F,η,p*/ρ,m*/ρ)si463_e gives

˙ε=˘εF·˙F+˘εη˙η+˘ε(p*ρ)··¯(p*ρ)+˘ε(m*ρ)··¯(m*ρ),

si464_e

and substitution of this result into the second law (9.61) leads to

(1ρRP˘εF)·˙F+(Θ˘εη)˙η+(e*˘ε(p*ρ))··¯(p*ρ)+(μoh*˘ε(m*ρ))··¯(m*ρ)+1ρj*·e*1ρΘq·gradΘ0.

si465_e  (9.63)

Inequality (9.63) must hold for all processes. Since the coefficients of the rates (˙Fsi466_e, ˙ηsi467_e, etc.) in (9.63) are independent of the rates themselves, and the rates may be varied independently and are arbitrary, it follows that the coefficients vanish, i.e.,

P=ρR˘εF,Θ=˘εη,e*=˘ε(p*ρ),h*=1μo˘ε(m*ρ).

si468_e  (9.64)

What remains of inequality (9.63), i.e.,

j*·e*1Θq·gradΘ0,

si469_e  (9.65)

is called the residual dissipation inequality. The residual dissipation inequality quantifies irreversibilities in a thermodynamic process. In this case, the first term in (9.65) represents Joule heating due to current flow, and the second term represents heat conduction; both are transport processes. Accordingly, unlike the other dependent variables (see (9.64)), the conductive current j* and heat flux q are not derivable from a thermodynamic energy potential. Instead, their constitutive equations are specified empirically (e.g., Ohm's law for j* and Fourier's law for q), with the caveat that they must satisfy inequality (9.65).

We collectively coin the set of extensive independent variables {F, η, p*/ρ, m*/ρ}, the thermodynamic energy potential {F,η,p*/ρ,m*/ρ}si470_e, and the constitutive equations (9.64) the all-extensive formulation. The all-extensive formulation, along with the other formulations that we present in this section, correlate with a particular experiment, the independent variables being controlled and the dependent variables being the measured responses.

9.6.2.1 Polarization and magnetization as independent variables

From an experimental point of view, it is more practical to control the electric polarization p* and magnetization m* than the polarization per unit mass p*/ρ and magnetization per unit mass m*/ρ. Hence, we modify the all-extensive formulation presented in Section 9.6.2 to accommodate the use of p* and m* as independent variables. We proceed by using the chain rule

·¯(p*ρ)=1ρ˙p*+1ρ(FT·˙F)p*,·¯(m*ρ)=1ρ˙m*+1ρ(FT·˙F)m*

si471_e  (9.66)

to rewrite the fundamental form (9.61) of the Clausius-Duhem inequality:

˙ε+[1ρRP+1ρ(e*·p*+μoh*·m*)FT]·˙F+Θ˙η+1ρe*·˙p*+μoρh*·˙m*+1ρj*·e*1ρΘq·gradΘ0,

si472_e  (9.67)

where we have used

·¯(1ρ)=1ρdivv,divv=trL=tr(˙FF1)=FT·˙F.

si473_e  (9.68)

In the modified form (9.67) of the second law, polarization p* and magnetization m* appear as rates (i.e., natural independent variables). Accordingly, the thermodynamic energy potential ε in this formulation is a function of F, η, p*, and m*, i.e., ε=ˉε(F,η,p*,m*)si474_e; a superscript bar is used instead of a superscript breve to signify a different internal energy function with the same value. Use of the chain rule gives

˙ε=ˉεF·˙F+ˉεη˙η+ˉεp*·˙p*+ˉεm*·˙m*,

si475_e

and substitution of this result into (9.67) leads to

(1ρRP+1ρ(e*·p*+μoh*·m*)FTˉεF)·˙F+(Θˉεη)˙η+(1ρe*ˉεp*)·˙p*+(μoρh*ˉεm*)·˙m*+1ρj*·e*1ρΘq·gradΘ0.

si476_e  (9.69)

Since the coefficients of ˙Fsi477_e, ˙ηsi478_e, ˙p*si479_e, and ˙m*si480_e in inequality (9.69) are independent of the rates, and the rates may be varied independently and are arbitrary, it follows that the coefficients vanish, i.e.,

P=ρRˉεFJ(e*·p*+μoh*·m*)FT,Θ=ˉεη,e*=ρˉεp*,h*=ρμoˉεm*.

si481_e  (9.70)

We collectively coin the set of independent variables {F, η, p*, m*}, the thermodynamic energy potential ε=ˉε(F,η,p*,m*)si482_e, and the constitutive equations (9.70) the deformation-entropy-polarization-magnetization formulation.

Exercise

1. Verify the relationships in (9.68).

9.6.3 Other formulations

Knowledge of the internal energy function ε=˘ε(F,η,p*/ρ,m*/ρ)si483_e or ε=ˉε(F,η,p*,m*)si484_e is sufficient to characterize a reversible TEME material. Said differently, specifying the functional form of the energy potential determines P, Θ, e*, and h* in (9.64) and (9.70). The functional form of the energy potential is ascertained experimentally; this experimental characterization most straightforwardly accomplished when the independent and dependent variables synchronize with those one wishes to control and measure, respectively, in an experiment. From an experimental perspective, it is often more practical to control intensive quantities than extensive quantities; e.g., temperature is easier to control than entropy or internal energy. To change an independent variable from extensive to intensive, a new free energy is defined through a Legendre transformation of the internal energy, i.e.,

new free energy=internal energy(intensive)(extensive).

si485_e

Following this blueprint, we present in Tables 9.39.6 a catalogue of the 15 possible Legendre transformations of the internal energy. These Legendre transformations are divided into four families, each employing a common set of thermomechanical independent variables: family 1, deformation and entropy (both extensive); family 2, deformation (extensive) and temperature (intensive); family 3, stress (intensive) and entropy (extensive); and family 4, stress and temperature (both intensive). For the sake of brevity, we employ the compact notation Eabcd for the Legendre-transformed energy potentials, where the superscript letters a, b, c, and d are placeholders for an appropriate mechanical, thermal, electrical, and magnetic independent variable, respectively. This compact notation denotes that the Legendre-transformed energy potential EFΘpm, for instance, is a function of deformation F, temperature Θ, electric polarization per unit mass p*/ρ, and magnetization per unit mass m*/ρ. Note that the Legendre transformations in Tables 9.39.6 provide explicit connections between the new free energies Eabcd and the primitive internal energy ε of the system, whose evolution is governed by the first law of thermodynamics (9.57c).

Table 9.3

Energy Family 1

Independent VariablesEnergyLegendre Transformation
F, η, p*/ρ, m*/ρε
F, η, e*, m*/ρEFηemEFηem=εe*·p*ρsi104_e
F, η, p*/ρ, h*EFηphEFηph=εμoh*·m*ρsi105_e
F, η, e*, h*EFηehEFηeh=εe*·p*ρμoh*·m*ρsi106_e

Table 9.4

Energy Family 2

Independent VariablesEnergyLegendre Transformation
F, Θ, p*/ρ, m*/ρEFΘpmEFΘpm = ε − Θη
F, Θ, e*, m*/ρEFΘemEFΘem=εΘηe*·p*ρsi107_e
F, Θ, p*/ρ, h*EFΘphEFΘph=εΘημoh*·m*ρsi108_e
F, Θ, e*, h*EFΘehEFΘeh=εΘηe*·p*ρμoh*·m*ρsi109_e

Table 9.5

Energy Family 3

Independent VariablesEnergyLegendre Transformation
P, η, p*/ρ, m*/ρEPηpmEPηpm=ε1ρRP·Fsi110_e
P, η, e*, m*/ρEPηemEPηem=ε1ρRP·Fe*·p*ρsi111_e
P, η, p*/ρ, h*EPηphEPηph=ε1ρRP·Fμoh*·m*ρsi112_e
P, η, e*, h*EPηehEPηeh=ε1ρRP·Fe*·p*ρμoh*·m*ρsi113_e

Table 9.6

Energy Family 4

Independent VariablesEnergyLegendre Transformation
P, Θ, p*/ρ, m*/ρEPΘpmEPΘpm=ε1ρRP·FΘηsi114_e
P, Θ, e*, m*/ρEPΘemEPΘem=ε1ρRP·FΘηe*·p*ρsi115_e
P, Θ, p*/ρ, h*EPΘphEPΘph=ε1ρRP·FΘημoh*·m*ρsi116_e
P, Θ, e*, h*EPΘehEPΘeh=ε1ρRP·FΘηe*·p*ρμoh*·m*ρsi117_e

The constitutive equations associated with the 15 formulations shown in Tables 9.39.6 are obtained via restrictions imposed by the second law of thermodynamics, invariance requirements, conservation of angular momentum, and material symmetry; refer to [62] for a complete catalogue. In the following sections, as a representative example, we investigate the deformation-temperature-electric field-magnetic field formulation. However, before we proceed, note the following:

(1) A procedure for introducing free energies that use the electric displacement d* and magnetic induction b* as independent variables, and the derivation of their associated constitutive equations, is presented in Appendix F.

(2) For formulations that employ either the electric polarization per unit mass p*/ρ or the magnetization per unit mass m*/ρ as an independent variable, as was the case with the all-extensive formulation in Section 9.6.2, p* or m* can be introduced as the independent variable ex post facto using the procedure described in Section 9.6.2.1. Refer to Problem 9.9.

Problem 9.9

Using the second law of thermodynamics and the appropriate Legendre transformation, derive the constitutive equations for the deformation-temperature-polarization-magnetic field formulation.

Solution

In this formulation, F, Θ, p*/ρ, and h* are the independent variables. We define the energy potential EFΘph as the Legendre transformation of internal energy ε=˘ε(F,η,p*/ρ,m*/ρ)si201_e with respect to the thermal and magnetic variables, from η to Θ and m*/ρ to h*, i.e.,

EFΘph=εΘημoh*·m*ρ.

si202_e

Refer to Table 9.4. The rate of the Legendre transformation is

˙EFΘph=˙εΘ˙ηη˙Θμoh*··¯(m*ρ)μom*ρ·˙h*,

si203_e

and subsequent use of this result in the second law of thermodynamics (9.61) leads to

˙EFΘph+1ρRP·˙Fη˙Θ+e*··¯(p*ρ)μom*ρ·˙h*+1ρj*·e*1ρΘq·gradΘ0.

si204_e  (a)

Use of the chain rule on ˙EFΘphsi205_e gives

˙EFΘph=EFΘphF·˙F+EFΘphΘ˙Θ+EFΘph(p*ρ)··¯(p*ρ)+EFΘphh*·˙h*,

si206_e

and substitution of this result into inequality (a) leads to the constitutive equations

P=ρREFΘphF,η=EFΘphΘ,e*=EFΘph(p*ρ),m*=ρμoEFΘphh*.

si207_e

We now illustrate how the procedure described in Section 9.6.2.1 can be used to introduce the polarization p* (instead of the polarization per unit mass p*/ρ) as the electrical independent variable. Equation (9.66)1, i.e.,

·¯(p*ρ)=1ρ˙p*+1ρ(FT·˙F)p*,

si208_e

is used to rewrite the second law inequality (a) as

˙EFΘph+[1ρRP+1ρ(e*·p*)FT]·˙Fη˙Θ+1ρe*·˙p*μom*ρ·˙h*+1ρj*·e*1ρΘq·gradΘ0.

si209_e  (b)

Note that the energy potential EFΘph in (b) uses the polarization p* instead of the polarization per unit mass p*/ρ as the electrical independent variable. It then follows that

˙EFΘph=EFΘphF·˙F+EFΘphΘ˙Θ+EFΘphp*·˙p*+EFΘphh*·˙h*.

si210_e

Subsequent use of this result in (b) leads to the constitutive equations

P=ρREFΘphFJ(e*·p*)FT,η=EFΘphΘ,e*=ρEFΘphp*,m*=ρμoEFΘphh*.

si211_e

Problem 9.10

Using the second law of thermodynamics and the appropriate Legendre transformation, derive the constitutive equations for the temperature-stress-electric field-magnetic field formulation.

Solution

The temperature-stress-electric field-magnetic field formulation correlates with a thermodynamic process where {P, Θ, e*, h*} are the independent variables and EPΘeh is the characterizing thermodynamic energy potential. EPΘeh is defined as the Legendre transformation of internal energy ε=˘ε(F,η,p*/ρ,m*/ρ)si212_e with respect to the mechanical, thermal, electrical, and magnetic variables, from F to P, η to Θ, p*/ρ to e*, and m*/ρ to h*,

EPΘeh=ε1ρRP·FΘηe*·p*ρμoh*·m*ρ.

si213_e

Refer to Table 9.6. Taking the rate of the Legendre transformation gives

˙EPΘeh=˙ε1ρRP·˙F1ρRF·˙PΘ˙ηη˙Θe*··¯(p*ρ)p*ρ·˙e*μoh*··¯(m*ρ)μom*ρ·˙h*,

si214_e

and substitution of this result into the second law (9.61) yields

˙EPΘeh1ρRF·˙Pη˙Θp*ρ·˙e*μom*ρ·˙h*+1ρj*·e*1ρΘq·gradΘ0.

si215_e

It can be shown that use of the chain rule on ˙EPΘehsi216_e leads to the constitutive equations

F=ρREPΘehP,η=EPΘehΘ,p*=ρEPΘehe*,m*=ρμoEPΘehh*,

si217_e

an exercise that we leave to the reader.

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

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