• Nebyly nalezeny žádné výsledky

T HEORY OF M IXTURES

N/A
N/A
Protected

Academic year: 2022

Podíl "T HEORY OF M IXTURES "

Copied!
108
0
0

Načítání.... (zobrazit plný text nyní)

Fulltext

(1)

T HEORY OF M IXTURES

COURSE LECTURE NOTES

J

OSEF

M

ÁLEK

, O

ND ˇREJ

S

OU ˇCEK Charles University

Prague 2020

WE THANKKARELTUMAˇ , MICHALHABERA, VOJT ˇECH PATO ˇCKA AND MARKDOSTALÍK FOR THEIR HELP AND REMARKS.

(2)

Contents

1 Theory of interacting continua - introduction 2

2 Reminder of the concepts of classical equilibrium thermodynamics 3 2.1 Physical postulates and mathematical model for a macroscopic system in thermodynamic equi-

librium . . . 4

2.2 Examples . . . 6

2.3 Physical interpretation of the mathematical model . . . 7

2.4 Other thermodynamic potentials - Legendre transform . . . 9

3 Reminder of some basic concepts of classical continuum thermodynamics and mechanics of single continuum 11 3.1 Basic cornerstones of continuum thermodynamics . . . 11

3.2 Local equilibrium thermodynamics . . . 14

3.2.1 Entropic representation (for fluid mixtures) . . . 14

3.2.2 Energetic representation . . . 17

3.2.3 Other thermodynamic potentials - Enthalpy, Helmholtz and Gibbs free energy . . . 18

3.3 Structure of local equilibrium thermodynamics directly from the local form of fundamental relation . . . 20

3.4 Molar-based quantities . . . 21

3.5 Constitutive theory of continuum thermodynamics . . . 24

4 Kinematical description of mixtures 26 5 Balance equations 29 5.1 Auxiliary definitions - mass, volume other measures . . . 29

5.2 General form of a balance law in the bulk (in Eulerian description) . . . 30

5.3 Balance of mass . . . 31

5.4 Balance of linear momentum . . . 32

5.5 Balance of angular momentum . . . 34

5.6 Balance of energy (for non-polar mixture of non-polar constituents) . . . 37

5.7 Balance of entropy (Second law of thermodynamics) . . . 40

5.8 Classification of the mixture theories . . . 41

6 Class I mixtures 41 6.1 Fick-Navier-Stokes-Fourier model (ψ=ψ¯(ϑ,1ρ,c)) . . . 44

6.2 Multi-component Fick - NSF model . . . 47

6.3 Constraints: Incompressibility and Quasi-incompressibility . . . 48

6.3.1 Incompressibility . . . 48

6.3.2 Quasi-incompressibility . . . 48

6.4 Cahn-Hilliard-NSF and Allen-Cahn-NSF model . . . 53

6.5 Allen-Cahn and Cahn-Hilliard models as gradient flows . . . 60

6.6 Chemical reactions . . . 61

6.6.1 Stoichiometry . . . 62

6.6.2 Mixture of ideal gases . . . 67

6.6.3 Chemical equilibrium . . . 69

6.6.4 Chemical kinetics . . . 71

7 Class II mixtures 76 7.1 Motivation . . . 76

7.2 Balance laws . . . 77

7.3 Interaction forces - structure inferred from balance laws . . . 77

7.4 Interaction forces - macroscopic mechanical analogies . . . 79

7.4.1 Flow around a sphere . . . 79

(3)

7.5 Darcy’s law . . . 83

7.5.1 Reduction of two-component momentum balance . . . 83

7.5.2 Derivation from macroscopic analogy . . . 84

7.5.3 Derivation through homogenization . . . 85

7.6 Generalizations of Darcy’s law: models of Brinkman and Forchheimer . . . 88

7.7 A thermodynamic framework for a mixture of two liquids . . . 89

7.7.1 Compressible case . . . 91

7.7.2 Incompressible case . . . 94

8 Multiphase continuum thermodynamics 96 8.1 Introduction . . . 96

8.2 General balance laws under presence of discontinuities - single continuum . . . 97

8.3 Jump conditions . . . 101

8.4 A thermodynamic framework for boundary conditions . . . 103

8.5 General balance laws under presence of discontinuities - multicomponent continuum . . . 105

8.6 Framework of the multiphase continuum theory . . . 105

8.7 Averaging theorems . . . 105

1 Theory of interacting continua - introduction

What is a mixture? Oxford English dictionary: mixture: “A product of mixing, a complex unity or aggre- gate (material or immaterial) composed of various ingredient orconstituent partsmixed together”.

and goes on as follows:

“... mixed state or condition; co-existenceof different ingredients or different groups or classes of things mutually diffusedthrough each other.”

Goal of the course

• Basic understanding of (derivation of) models that might be capable of describing the following pro- cesses (examples of mixtures around us):

– Geophysics- Thermohaline circulation - (convection due to both temperature and concentration variations), Porous media flow - flow of water and transport of solubles through soils, rocks,..., Avalanches, Pollution spreading and transport through environment, Melting of ice and freezing of meltwater in glaciers (phase transitions), Mineral phase transitions in rocks, Generation and extraction of magma, Liquefaction (change of strength of porous water saturated solids under seismic forcing),..

– Astrophysics- Plasmas and gaseous mixtures in the stars

– Biology- Flow of tracers/fluids through biological tissues, Processes at cell membranes, Biochem- ical reactions, Remodulation of bones, Blood coagulation, Cloth formation and dissolution, Flow of blood plasma,...

– Chemistry- Chemical reactions, Chemical equilibrium, Chemical kinetics

– Material sciences- Multi-phase flow, Flow of aerosols, dyes, Deformation of composite materials ...

• Understanding of the assumptions required for the derivation of the models and developing framework for their possible generalizations.

• To arrive at three-dimensional thermodynamically consistent material models of mixtures which in- clude temperature effects.

• To understand how the models of Fick, Darcy, Forchheimer, Biot, Brinkman, Allen-Cahn, Cahn-Hilliard and Stephan fit into the general framework of the mixture theory.

• To provide a sufficiently colorful “palette” of material models to cover the real-world processes.

(4)

Some names & literature

• Fick (1855), On liquid diffusion.

• Darcy (1856), Les fontaines publiques de la ville de Dijon.

• Muskat (1937), The Flow of Homogeneous Fluids through Porous Media.

• Truesdell and Toupin (1960), The Classical Field Theories.

• Truesdell and Noll (1965), The non-linear field theories of mechanics.

• Atkin and Craine (1976), Continuum Theories of Mixtures: Applications

• Bowen (1976), Theory of mixtures.

• Truesdell (1984), Rational thermodynamics.

• Müller (1985), Thermodynamics

• Samohýl (1982), Racionální termodynamika chemicky reagujících smˇesí.

• de Groot and Mazur (1984), Non-equilibrium Thermodynamics.

• Rajagopal and Tao (1995), Mechanics of Mixtures.

• Drew and Passman (1998), Theory of Multicomponent Fluids.

• Pekaˇr and Samohýl (2014), The thermodynamics of linear fluids and fluid mixtures.

• Bothe and Dreyer (2015), Continuum thermodynamics of chemically reacting fluid mixtures.

Two main approaches within the theory of interacting continua

Approach 1 - a concept based on the principle of equi-presence (all components of the mixture are assumed to co-exist simultaneously in each point of the continuum; the foundations of this approach have been laid by Truesdell et al. (e.g. Truesdell and Toupin, 1960) and is sometimes referred to as Continuum mixture theory

Approach 2- a concept relying on postulating classical single-component balance laws together with interface conditions at intermediate (sufficiently fine spatial) scale, where components of the mixture are distinguishable (e.g. at the scale of the pore-space in case of porous media flow). This is followed by a suitable averaging procedure, which delivers a continuum-like mixture framework; one of the classical references to this approach is Drew and Passman (1998). This approach is often referred to as Multi-phase theory.

In this lecture, we shall mostly rely on the first approach, with the exception of Chapter 8, where founda- tions of the second approach will be presented.

2 Reminder of the concepts of classical equilibrium thermodynamics

In this section, we first address as a reminder the formal structure of classical equilibrium thermodynamics, following closely Evans’ Lecture notes entropy and PDEs. (Evans, 2017). The development of classical phys- ical notions is here performed in a mathematically formal way, a more "physical" introduction to the topic can be found, for example in Callen (1960). In the next chapters, these concepts are used under so-called assumption of local equilibrium used to transfer the thermodynamic relation to the realm of continuum physics.

(5)

2.1 Physical postulates and mathematical model for a macroscopic system in thermo- dynamic equilibrium

Let us nevertheless start with the physical postulates (c.f. Callen, 1960, p. 283-284):

A physical model for a system in thermal equilibrium

• There exist particular states (called equilibrium states) that, macroscopically are characterized com- pletely by the specification of internal energyEand set of extensive parametersX1, . . .Xm

• There exists a function (calledentropy) of the extensive parameters, defined for all equilibrium states, with the following property: the value assumed by the extensive parameters in the absence of con- straints, are those that maximize the entropy over the manifold of equilibrium states

• The entropy of a composite system is addition over the constituent subsystems (hence the entropy of each system is a homogeneous first order function of the extensive parameters. The entropy is continuous and differentiable and is monotonically increasing function of energy

• The entropy of any system vanishes in the state for whichT:=∂SE =0

Let us translate these physical postulates into a set of mathematical postulates A mathematical model for a system in thermal equilibrium

Let us suppose we are given

• an open convex subsetΣofRm+1

• aC1 function ˆS:Σ→Rsuch that (i) ˆSis concave

(ii) ∂ESˆ >0

(iii) S is positively homogeneous of degree 1:

S(ˆ λE,λX1, . . . ,Xm)=λS(E,ˆ X1, . . . ,Xm) λ>0 (2.1) We callΣthe state space andSthe entropy of the system

Sdef= S(E,ˆ X1, . . . ,Xm) Fundamental thermodynamic equation (2.2) depending on the internal energyEand other extensive quantities X1, . . . ,Xm (examples of such quantities are for example the volumeV, number (or molar number) of particles of individual atoms or moleculesNαor their massesMα, electric polarizationP, magnetizationM, etc.)

Owing to (ii), we can invert relation (2.2) and obtain aC1function ˆEas a function

E=E(S,ˆ X1, . . . ,Xm). (2.3)

We define:

• Thermodynamic temperature T

T=Tˆ=

S (2.4)

• Generalized force (or pressure)

Pk=Pˆk= −

Xk (2.5)

Lemma 2.1. (i) The functionE is positively homogeneous of degree 1:ˆ

E(ˆ λS,λX1, . . . ,λXm)=λE(S,ˆ X1, . . . ,Xm) λ>0. (2.6)

(6)

(ii) The functionsT,ˆ Pˆk are positively homogeneous of degree 0:

Tˆ(λS,λX1, . . . ,λXm)=Tˆ(S,X1, . . . ,Xm) (2.7) Pˆk(λS,λX1, . . . ,λXm)=Pˆk(S,X1, . . . ,Xm) λ>0 (2.8) These properties will leads to distinguishing so-calledextensive parameters:S,E(and alsoXk), which- so-to-speak depend on how “big” the system is andintensive parametersT, Pk, which do not depend on the “size” of the system.

Proof. 1. Clearly, it must hold for allE,X1, . . . ,Xm:

E=E( ˆˆ S(E,X1, . . .Xm),X1, . . . ,Xm). (2.9) Thus for allλ>0:

λE=E( ˆˆ S(λE,λX1, . . .λXm),λX1, . . . ,λXm)

=E(ˆ λS(E,ˆ X1, . . . ,Xm),λX1, . . . ,λXm) (2.10) due to (2.1)

2. Since ˆSisC1, so is ˆE. Differentiate (2.6) with respect toS:

λ∂

S(λS,λX1, . . . ,λXm)

| {z }

T(λS,λX1,...,λXm)

=λ∂

S(S,X1, . . . ,Xm)

| {z }

T(S,X1,...,Xm)

(2.11)

Lemma 2.2. It holds

E = 1

T,

Xk =Pk

T k=1, . . . ,m (2.12)

Proof. We definedT asT=ESˆ and the function ˆS(E,X1, . . . ,Xm) as an inverse ofE=E(S,ˆ X1, . . . ,Xm). Thus it holds

E=E( ˆˆ S(E,X1, . . . ,Xm),X1, . . . ,Xm) . (2.13) Therefore

1=dE dE=

S

E=⇒T=

S = Ã

E

!−1

. (2.14)

Similarly, by taking a derivative with respect toXk, we get 0=

S

|{z}

T

Xk+

Xk

| {z }

Pk

k=1, . . . ,m. (2.15)

The definitions (2.4) and (2.5) can be summarized by the so-calledGibbs’s formula dE=T dS−

m

X

k=1

Pkd Xk (2.16)

(7)

2.2 Examples

The extensive parameters X1, . . . ,Xm can represent various physical quantities. Let us discuss some proto- typical examples

1. Simple fluid. Possibly the simplest example of an equilibrium thermodynamic system is a homoge- neous simple fluid, for which the extensive parameters are:

• E- internal energy

• X1=V - volume

• X2=N - number of particles

consequently, the fundamental thermodynamic relation for such system readsS=S(E,Vˆ ,N), and the associated intensive parameters are

• T=ESˆ - thermodynamic temperature

• P1=P= −VEˆ - pressure

• P2= −µ= −∂NEˆ - (molar) chemical potential

The Gibbs formula for homogeneous simple fluid thus reads dE=T dS−P dV+µdN .

2. Multicomponent simple fluid. If the fluid is composed of N constituents, we are in the following setting. The extensive parameters are now

• E- internal energy

• X1=V - volume

• X2, . . . ,XN+1=N1, . . . ,NN - numbers of particles of the individual constituents

consequently, the fundamental thermodynamic relation for such system readsS=S(E,ˆ V,N1, . . . ,NN), and the associated intensive parameters are

• T=ESˆ - thermodynamic temperature

• P1=P= −VEˆ - pressure

• P2, . . .PN+1:Pα+1= −µα= −∂NEˆα - chemical potentialsα=1, . . . ,N The Gibbs formula for such a multicomponent simple fluid thus reads

dE=T dS−P dV+

N

X

α=1

µαdNα.

3. Homogeneous single-component dielectric fluid in a homogeneous electrostatic field with intensityE. From the principle of virtual work, one can deduce that the change of total energy of the volume filled with the dielectric is

dE=T dS−P dV+µdN + Z

E·dD.

whereDis the electric induction vector. Assuming a homogenous field, this relation can be written as dE=T dS−P dV+µdN +VE·dD.

For an isotropic linear dielectric body readsD=ε0E+P, wherePis the so-called polarization vector and ε0 is the permitivity of vacuum. Often one introduces a reduced energy ˜E:=E−V|E8π|2 which results

(8)

is obtained by subtracting the energy of the electrostatic field (it cannot be however interpreted as intrinsic energy of the dielectric since the value ofEis affected by the presence of the dielectric).

The Gibbs relation in terms of ˜Ereads

dE˜=T dS−P dV+µdN +VE·dP

and, by introducing the extensive quantity total polarizationP :=PV, we can rewrite it as dE˜=T dS−P dV˜ +µdN +E·dP ,

with ˜P=P+ε20|E|2 +E·P, and consider fundamental relation ˜E=E(S,b˜ V,N,P).

2.3 Physical interpretation of the mathematical model

Let us spend some time with an attempt of deeper explanation of the mathematical assumptions laid in our model and their physical interpretations and consequences.

• Theassumption ESˆ >0 impliesT>0, i.e. positivity of the thermodynamic temperature

• The1-homogeneity assumption ofSˆ is equivalent toadditivity of entropy over subsystems of an equilibrium system. Indeed, consider a bodyB in equilibrium, with internal energyE, entropy S and other extensive variablesXk,k=1, . . . ,m. Let us consider a subregionB(1), which represents a λfraction ofB, and consequently, its extensive parameters take values: E(1)=λE, X1(1)=λX1. . .Xm(1)= λXm, and alsoS(1)=λS, due to 1-homogeneity of ˆS, since

S(1)=S(Eˆ (1),X1(1), . . . ,Xm(1))=S(ˆ λE,λX1, . . . ,λXm)=λS(E,ˆ X1, . . . ,Xm)=λS.

The complementary subregionB(2)(such thatB=B(1)∪B(2)), has extensive parametersE(2)=(1−λ)E, X(2)1 =(1−λ)X1. . .Xm(2)=(1−λ)Xmand thusS(2)=(1−λ)Sby 1-homogeneity of ˆS. Consequently, we get S=S(1)+S(2). The argument can be reversed and by assuming additivity of entropy over subsystems of an equilibrium system, we would deduce 1-homogeneity of ˆS: Take division ofB intoN subsystems each of the same size, i.e. N1 fraction ofB. By the additivity assumption, it must hold

S=

N

X

i=1

S(i)=NSˆ µ1

NE, 1

NX1, . . . , 1 NXm

¶ ,

which must hold for allN∈N. Using the same argument for subsystems of the size MN,M∈{1, . . . ,N}, we obtain thatqS(E,ˆ X1, . . . ,Xm)=S(qE,ˆ qX1, . . . ,qXm) for allq∈Q,q>0. By the assumption of continuity of ˆSand by the density argument ofQinR, we obtain the desired result.

• On the other hand, we have proven the intensive parametersT,P1, . . . ,Pm, to be 0-homogeneous. This implies that these parameters take the same value irrespective of the size of the subregion, i.e.

T=T(1)=T(2)=. . . ., Pk=P(1)k =P(2)k =. . . . k=1, . . . ,m.

Concavity of S. Let us now consider two isolated bodies made out of the same substanceˆ B(1) and B(2)which do not form subregions of one equilibrium system, but each of them is in equilibrium. Let

S(1)=S(Eˆ (1),X11, . . . ,X1m) S(2)=S(Eˆ (2),X12, . . . ,Xm2)

The total entropy of the two systems in this configuration isS(1)+(2)=S(1)+S(2). If we now combine the two bodies into one (e.g. by allowing transfer of all extensive parameters such as energy, mass,...) but without doing any work, neither allowing heat transfer from outside, we end up with a systemB(3) with extensive parametersE(3)=E(1)+E(2), X(3)k =X(1)k +Xk(2),k=1, . . . ,m. Since irreversible processes may have taken place inside the systemB(3)before it equilibrated, we cannot assume more than

S(3)≥S(1)+S(2)

(9)

i.e. we assume that the final equilibrium entropy of the joint system is greater or equal than the sum of the two entropies of the originally isolated subsystems. So this means that

S(3)=S(Eˆ (1)+E(2),X1(1)+X1(2), . . . ,Xm(1)+Xm(2))≥S(1)+S(2)=S(Eˆ (1),X1(1), . . . ,X(1)m)+S(Eˆ (2),X(2)1 , . . . ,Xm(2)) . This however implies that ˆSis concave, since taking any 0<λ<1:

Sˆ³

λE(1)+(1−λ)E(2),λX1(1)+(1−λ)X1(2), . . . ,λXm(1)+(1−λ)X(2)m´

≥Sˆ³

λE(1),λX1(1), . . . ,λXm(1)´ +Sˆ³

(1−λ)E(2), (1−λ)X(2)1 , . . . , (1−λ)Xm(2)´

=λSˆ³

E(1),X1(1), . . . ,Xm(1)´

+(1−λ) ˆS³

E(2),X(2)1 , . . . ,X(2)m´ where in the last equality, we employed the 1-homogeneity of ˆS.

Convexity ofEˆas a consequence of concavity of ˆS. Let us take anyS(1),X1(1), . . .X(1)m andS(2),X(2)1 , . . .X(2)m and any 0<λ<1. Define

E(1)=E(Sˆ (1),X1(1), . . .Xm(1)) E(2)=E(Sˆ (2),X1(2), . . .Xm(2)) and thus

S(1)=S(Eˆ (1),X1(1), . . .Xm(1)) S(2)=S(Eˆ (2),X1(2), . . .X(2)m) . Since ˆSis concave, it must hold

Sˆ³

λE(1)+(1−λ)E(2),λX1(1)+(1−λ)X1(2), . . . ,λX(1)m +(1−λ)X(2)m´

λS(Eˆ (1),X1(1), . . . ,X(1)m)+(1−λ) ˆS(E(2),X1(2), . . . ,Xm(2)) . Since it holds

E=E( ˆˆ S(E,X1, . . . ,Xm),X1, . . . ,Xm) and since ESˆ =T≥0, we obtain

λE(1)+(1−λ)E(2)=Eˆ³

S(ˆ λE(1)+(1−λ)E(2), . . . ,λX(1)k +(1−λ)X(2)k , . . . ), . . . ,λXk(1)+(1−λ)Xk(2), . . .´

≥Eˆ³

λS(Eˆ (1), . . . ,X(1)k , . . . )+(1−λ) ˆS(E(2), . . . ,Xk(2), . . . ), . . . ,λX(1)k +(1−λ)X(2)k , . . .´

=E(ˆ λS(1)+(1−λ)S(2), . . . ,λXk(1)+(1−λ)Xk(2), . . . ).

So we obtained

λE(Sˆ (1),X1(1), . . . ,Xm(1))+(1−λ) ˆE(S(2),X1(2), . . . ,X(2)m)≥E(ˆ λS(1)+(1−λ)S(2), . . . ,λX(1)k +(1−λ)X(2)k , . . . ) which implies that ˆEis convex.

Entropy maximization and energy minimizationfor thermally isolated systems.

– Entropy maximization principle - The equilibrium value of any unconstrained internal pa- rameter is such as to maximize the entropy of the system for the given value of total internal energy.

– Energy minimization principle- The equilibrium value of any unconstrained internal param- eter is such as to minimize the energy for the given value of the total entropy.

(10)

Figure 1: Entropy function ˆS(E, . . . ,Xk, . . . ) and visualization of the entropy maximization (left) end energy minimization (right) principle. The unconstrained variableXkattains equilibrium valueXk, such that either entropy is maximal provided that the energy is fixed at valueE (left), or such that the energy is minimal provided that the entropy is fixed at valueS.

2.4 Other thermodynamic potentials - Legendre transform

Assume that H:Rn→(−∞,+∞] is a convex, lower semicontinuous function, which is proper (i.e. not identi- cally equal to infinity).

Definition 1. The Legendre transform ofH(p) denotedH(q) is defined as H(q)=sup

p∈Rn(p·q−H(p)) (2.17)

One can show, that alsoH(q) is convex, lower semicontinuous and proper. Furthermore,

(H)=H, (2.18)

i.e.His a Legendre transform ofH, that’s why we call the pairH,H as dual convex functions.

Provided that on top of the assumptions before,His moreover also aC2and strictly convex, then for each q∈Rnthere exists a unique pointpwhere (p·q−H(p)) is maximal, namely the unique pointp, where

q=DpH(p) , providing relationq=q(p)ˆ (2.19) which can be inverted to give

p=p(q).ˆ (2.20)

The Legendre transform then reads

H(q)=p(q)ˆ ·q−H( ˆp(q)) (2.21) As a consequence, one obtains also

DqH(q)=p

q−DpH(p)¢

Dqpˆ=p, (2.22)

so to summarize

q=DpH(p) p=DqH(q) (2.23)

So far, we have seenS=S(E,ˆ X1, . . . ,Xm) andE=E(S,ˆ X1, . . . ,Xm). Now Legendre transform is used to define otherthermodynamic potentials

The Helmholtz free energyF is

F(T,V, . . . )def= inf

S ( ˆE(S,V, . . . )−T S) (2.24)

(11)

The enthalpyHis

H(S,P, . . . )def=inf

V ( ˆE(S,V, . . . )+PV) (2.25)

The Gibbs free energyGis

G(T,P, . . . )def= inf

S,V( ˆE(S,V, . . . )−T S+PV) (2.26) Assuming ˆEis strictly convex andC2and that the infima are attained at a unique point in the domain of ˆE, we can rewrite the potentials as

The Helmholtz free energy

F(T,V, . . . )=E−T S, where S=S(T,ˆ V, . . . ) solves T=

E(S,ˆ V, . . . )

S (2.27)

The enthalpy

H(S,P, . . . )=E+PV, where V=V(S,ˆ P, . . . ) solves P= −

E(S,Vˆ , . . . )

V (2.28)

The Gibbs free energy

G(T,P, . . . )=E−T S+PV, where S=S(Tˆ ,P, . . . ) solves T=

E(S,Vˆ , . . . )

S , P= −

E(S,Vˆ , . . . )

V (2.29) Based on the definitions and the assumption ˆEis strictly convex andC2, we can prove the following

Lemma 2.3. 1. E(S,ˆ V, . . . )is locally strictly convex in(S,V)

2. Fˆ(T,V, . . . )is locally strictly concave in T, locally strictly convex in V . 3. H(S,ˆ P, . . . )is locally strictly concave in P, locally strictly convex in S.

4. G(T,ˆ P, . . . )is locally strictly concave in(T,P).

Proof. First, the statement (1) is already assumed hence it is true. To prove (2), let us recall the definition of F

Fˆ(T,V, . . . )=E( ˆˆ S(T,V, . . . ),V, . . . )−TS(Tˆ ,V, . . . ) , with

T=

E( ˆˆ S(T,V, . . . ),V, . . . )

S (2.30)

This implies

T =

S

T −Sˆ−T

T = −S(T,ˆ V, . . . )

V =

S

V +

V −T

V =

V( ˆS(T,V, . . . ),V, . . . )= −P(T,ˆ V, . . . ) which implies

2

T2= −

S(Tˆ ,V, . . . )

T (2.31)

2

V2= −

P(T,ˆ V, . . . )

V =

2

SV

S(Tˆ ,V, . . . )

V +

2

V2 (2.32)

(12)

Differentiating (2.30) with respect to (T,V), we obtain 1=

2

S2

S(Tˆ ,V, . . . )

T 0=

2

S2

S(Tˆ ,V, . . . )

V +

2

VS Thus (2.31) and (2.32) imply

2F

T2 = − µ2

S2

1

2F

V2 =

2

V2− µ 2

SV

2µ2

S2

1

and since we assumed that ˆEis strictly convex in (S,V, . . . ), it holds

2

S2 >0,

2

V2 >0

µ2

S2

¶ µ2

V2

− µ 2

SV

2

>0 , and thus

2

T2<0 ,

2

V2>0 . (2.33)

The remaining two statements are proved analogously.

Exercise1. Prove the statements (3) and (4).

3 Reminder of some basic concepts of classical continuum thermody- namics and mechanics of single continuum

3.1 Basic cornerstones of continuum thermodynamics

A priori homogenization- continuum mechanics introduces the notion ofa material pointas a point- wise representative of some sufficiently small/sufficiently large control volume of real material; mate- rial properties assigned to such material point are averages (volume/time/stochastic) of the properties of the real material contained in the control volume.

X referential conf.

x=χ(X,t)

current conf.

B

κ0(B)

κt(B)

Figure 2: Essential kinematical concept of continuum mechanics - notion of motionχof an abstract bodyB, viewed as a mapping from some reference configurationκ0(B)⊂R3 to the current configurationκt(B)⊂R3 at given timet.

(13)

Kinematics

Motion - mapping

χ(·,t) :κ0(B)→κt(B) (3.1) R3×R→R3:x=χ(X,t) xk=χk(XK,t) , (3.2) which is assumed to be sufficiently smooth and invertible.

– Spatial gradient

GradΦ(X,t) = Φ(X,t)

X (3.3)

gradφ(x,t) = ∂φ

x (3.4)

– Velocity

V(X,t) = ∂χ

t

¯

¯

¯

¯X

in Lagrangean description (3.5)

v(x,t) = V(χ1(x,t),t) in Eulerian description (3.6) – Material time derivative

D

D tΦ(X,t) = Φ(X,t)

t

¯

¯

¯

¯X

(3.7) D

D tφ(x,t) = ∂φ(χ(X,t),t)

t

¯

¯

¯

¯X= ∂φ(x,t)

t

¯

¯

¯

¯x+v(x,t)·gradxφ(x,t) (3.8) – Deformation gradient

F(X,t)=∂χ(X,t)

X (F)kK(X,t)=∂χ

k(X,t)

XK (3.9)

– Green deformation tensor

C(X,t) = FTF (3.10)

(C)I J = (F)iI(F)jJδi j (3.11) (3.12) – Cauchy deformation tensor

c(x,t) = F−TF−1 (3.13)

(c)i j = (F−1)Ii(F−1)JjδI J (3.14) – Finger deformation tensor

B(X,t) = FFT (3.15)

(B)i j = (F)iI(F)jJδI J (3.16) – Piola deformation tensor

b(x,t) = F1FT (3.17)

(b)I J = (F1)Ii(F)Jjδi j (3.18) – Velocity gradient

L(x,t) = gradv (3.19)

(L)ij = vi

xj (3.20)

(14)

Exercise2. Show that ˙F=LF. Exercise3. Show that ˙B=LB+BLT.

• Volume and mass measures, constraints

Consider a mixture bodyΩand letP ⊂Ωopen subset, then we define

∀Pt“nice” M(P) . . . mass contained inP V(P) . . . volume ofP

Constraint: ∀Pt: 0= dtdV(Pt) (using the substitution theorem and identity ˙

det(F)=divvdetF), we obtain the incompressibility constraint 0=R

Ptdivv=0, localization gives divv=0

• Balance equations (in Eulerian description) Balance of mass

∂ρ

t +div(ρv)=0 , (3.21)

whereρis the density,vis the velocity.

Balance of linear momentum

(ρv)

t +div(ρvv)=divT+ρb, (3.22) whereTis the Cauchy stress,bis the intensity of the body forces.

Balance of angular momentum (for non-polar continuum)

T=TT. (3.23)

Balance of total energy

t µ

ρ µ

e+1 2|v|2

¶¶

+div µ

ρ µ

e+1 2|v|2

v

=div (Tvq)+ρb·v+ρr, (3.24) whereeis the specific internal energy, andris the energy supply (e.g. radiation).

In the framework of continuum thermodynamics, one has to consider also Balance of entropy

(ρη)

t +div¡ ρηv¢

+divqη+ρrη=ξ, where ξ≥0 . (3.25) Here ηis the specific entropy, qη is the entropy flux, rη is the specific entropy supply, and ξis the entropy production, which, by the second law of thermodynamics must be non-negative,

Exercise4. Show that (3.21)-(3.24) allow to rewrite (3.24) in a more compact form (balance of internal energy)

ρe˙=T:D−divq+ρr. (3.26)

Assuming the body forcesband energy supplyrare given, the balance equations (3.21), (3.22), (3.24) represent system of of 5 evolutionary equations for unknownsρ,e,v=(v1,v2,v3), T=(T11,T12, T13, T22,T23,T33),q=(q1,q2,q3) (14 unknowns), and it is known that it is in general impossible to predict the evolution of theTand q from this system knowing the initial state and boundary conditions. It is necessary to provide closure relations - relations describing howTand qdepend on the mechanical and thermal state of the system.

(15)

3.2 Local equilibrium thermodynamics

In this course we shall rarely leave the realm of so called local equilibrium thermodynamics - a thermody- namic theory in which each material point (representing sufficiently small subsystem of the whole system), is assumed to be in local thermodynamic equilibrium. By saying local, we mean, that this equilibrium only concerns the subsystem, but naturally, this material point behaves as an open subsystem, exchanging mass, energy, entropy etc., with the neighbouring points, with which equilibrium may not have been reached.

Based on this idea, we can exploit the notions and relations that have been derived for macroscopic systems in thermodynamic equilibrium and simply apply them on some small representative volume of our system. We will then deduce some local relations, and postulate them to hold point-wise in our continuum model. In the following application we will be exclusively interested in the description of fluid mixtures in the absence of electric and magnetic fields. We shall often tacitly assume this without further emphasizing this point explicitly.

3.2.1 Entropic representation (for fluid mixtures)

So consider this volume element dΩ, assumed to be in thermodynamic equilibrium, for which we apply the formal equilibrium structure, briefly reminded in Section 2. So again, we postulate existence of a function S called “entropy”, which is a function of energy of the system, and it extensive variables - in most cases we will restrict ourselves to volume, and masses of individual components of the system:

S=S(E,ˆ V,Mα) , (3.27)

or more precisely

S(dΩ)=S(E(dˆ Ω),V(dΩ),Mα(dΩ)), (3.28) where E(dΩ), V(dΩ) andMα(dΩ) are the energy, volume and masses of component in the subdomaindΩ. The function ˆSwas postulated (see (2.1)) to be

• positive 1−homogeneous with respect to the extensive variables, meaning

S(ˆ λE,λV,λMα)=λS(E,ˆ V,Mα) ∀λ>0 . (3.29)

• increasing function of energy

E>0 (3.30)

• ˆSis concave

From (3.27), we get by differentiating dS=

EdE+

VdV+

N

X

α=1

Mα

dMα. (3.31)

In chapter 2, we defined the thermodynamic temperatureϑ, thermodynamic pressurepand chemical poten- tialsµαby relations

1 ϑ=

E , p ϑ=

V , −µα ϑ =

Mα , (3.32)

so we can rewrite (3.31) in the classical form

ϑdS=dE+pdV−

N

X

α=1

µαdMα . (3.33)

Differentiating (3.29) w.r.t. λatλ=1, we obtained the correspondingEuler relation S(E,ˆ V,Mα)=

EE+

VV+

N

X

α=1

MαMα, (3.34)

(16)

or, equivalently, using the introduced definitions

ϑS=E+pV−

N

X

α=1

µαMα . (3.35)

Differentiating (3.35) and using (3.31), we obtained theGibbs-Duhem relation:

Sdϑ=V d p−

N

X

α=1

Mαdµα . (3.36)

Let us now explore the consequences of 1-homogeneity of entropy:

• Let us first take λ:= 1

M , withM:=PN α=1Mα: Sˆ

µ E M, V

M,Mα

M

= 1

MS(E,ˆ V,Mα) . (3.37)

Recognizing in the arguments of the function on the l.h.s., the specific energy e, inverse of density of the mixture 1ρ and concentrations (mass fractions) cα, defined as

edef= E

M , 1 ρ

def= V

M , cαdef= Mα

M α=1, . . . ,N, (3.38)

and observing that the function on the l.h.s. is a specific entropyη, we can write this relation as 1

MS(E,ˆ V,Mα)=ηdef=ηˆ(e,1

ρ,c1, . . . ,cN) , (3.39) differentiating the last equality yields

dη=ηˆ

ed e+ηˆ

1ρd µ1

ρ

¶ +

N

X

α=1

ηˆ

cαd cα (3.40)

Dividing the Gibbs-Duhem relation (3.36) byM, we get ηdϑ=1

ρd p−

N

X

α=1

cαdµα . (3.41)

Dividing the Euler relation (3.35) byM, we obtain its local form ϑη=e+p

ρ

N

X

α=1

µαcα. (3.42)

Finally, differentiating (3.42) and subtracting (3.41), we recover the local version of (3.33):

ϑdη=d e+pd µ1

ρ

N

X

α=1

µαd cα . (3.43)

Note that (3.43) implies the following local definitions of temperature, pressure and chemical potential:

1 ϑ= ηˆ

e

¯

¯

¯

¯1 ρ,c1,...cN

, p

ϑ = ηˆ

1ρ

¯

¯

¯

¯

¯e,c1,...,cN

, −µϑα= ηˆ

cα

¯

¯

¯

¯e,1 ρ,cβ6=α

(3.44)

(17)

Remark1. In the definition of ˆη(3.39) we did not employ the constraintPN

α=1cα=1, which follows from the definition of mass fractionscαin (3.38). Sometimes it is more convenient to employ this constraint when defining the specific entropy ˆηand eliminate one of the concentrations (withou loss of generality cN) and define reduced specific entropy ¯η

η¯(e,1

ρ,c1, . . . ,cN1)def= ηˆ Ã

e,1

ρ,c1, . . . ,cN1, 1−

N1

X

β=1

cβ

!

. (3.45)

Immediatelly, we then obtain the following relations 1

ϑ= η¯

e

¯

¯

¯

¯1

ρ,c1,...cN−1

, p

ϑ = η¯

1ρ

¯

¯

¯

¯

¯e,c1,...,cN−1

, −µαµN

ϑ = η¯

cα

¯

¯

¯

¯e,1

ρ,cβ β=1,...,N−1&β6=α

. (3.46)

Note that the only difference when employing this reduction arises in the last set of relations, i.e. defin- ing the partial derivatives of ¯ηwith respect to reduced (independent) set of concentrationsc1, . . . ,cN−1, which define thedifferenceof chemical potentials with respect to the eliminated component.

Clearly one can also employ the constraint in the Gibbs relation (3.43) and get ϑdη=d e+pd

µ1 ρ

N1

X

α=1

(µαµN)d cα. (3.47)

If one defines the relative chemical potentials µ¯αdef

= µαµN , (3.48)

one can simply use the reduced description (i.e. considerη=η¯³

e,1ρ,c1, . . . ,cN−1´

the Gibbs relation as

ϑdη=d e+pd µ1

ρ

N1

X

α=1

µ¯αd cα . (3.49)

• Now, let’s consider λ:= 1

V . Then 1-homogeneity of ˆSyields Sˆ

µE V, 1,Mα

V

= 1 V

S(E,Vˆ ,Mα) . (3.50)

Recognizing in the arguments of the function on the l.h.s., the volume density of energyρe, and partial densityρα, and observing that the function on the l.h.s. is a volume density of entropyρη, we can write this relation as

1 V

S(E,ˆ V,Mα)=ρη=ρηc(ρe,ρ1, . . . ,ρN) , (3.51) differentiating the last equality yields

d(ρη)= ρηc

(ρe)d(ρe)+

N

X

α=1

∂ρηc

∂ραdρα. (3.52)

Dividing the Gibbs-Duhem relation (3.36) byV, we get ρηdϑ=d p−

N

X

α=1ραdµα . (3.53)

Dividing the Euler relation (3.35) byV, we get ϑρη=ρe+p−

N

X

α=1

µαρα . (3.54)

(18)

Finally, differentiating (3.54) and subtracting (3.53), we recover the local version of (3.33):

ϑd(ρη)=d(ρe)−

N

X

α=1

µαdρα . (3.55)

Note that (3.55) implies the following alternative local definitions of temperature, and chemical poten- tial:

1

ϑ= ρηc

(ρe)

¯

¯

¯

¯ρ

α

, −µα

ϑ = ρηc

∂ρα

¯

¯

¯

¯ρe,ρ

β6=α

(3.56) Note: One might be puzzled now, where has the pressure vanished from our description in (3.55) and (3.56). In this setting we can introduce the pressure through the Euler relation (3.54). A natural question is whether thermodynamic pressure defined in such a manner coincides with (3.44).

Exercise5. Confirm by calculation that two discussed definitions of thermodynamic pressure are com- patible. HINT: Assume that you start from fundamental thermodynamic relationρη=ρηc(ρe,ρ1, . . . ,ρN) and realize that the introduced definitions imply

ηˆ µ

e,1

ρ,c1, . . . ,cN

= 1

MS(E,ˆ V,M1, . . . ,MN)= V M

1 V

S(E,ˆ V,M1, . . . ,MN)= V MSˆ

µE V, 1,M1

V , . . . ,MN

V

=1

ρρηc(ρe,ρ1, . . . ,ρN) .

Use this relation, define pressure through the Euler relation (3.54) and try to compute what is ϑ ηˆ1

ρ

¯

¯

¯

¯e,c

1,...,cN

=ϑ1 ρ

¯

¯

¯

¯e,c

1,...,cN

³

ρηc ρ

´

(ρe,ρ1, . . . ,ρN) .

3.2.2 Energetic representation

Alternatively, using the fact, that SEˆ = 1ϑ >0, assuming sufficient smoothness of ˆS we could invert (3.27) and write instead

E=E(S,ˆ V,Mα) . (3.57)

This function was proved in 2.1 to be also 1-homogeneous, i.e.

Eˆ µ S

M, V M,Mα

M

= 1

ME(S,ˆ V,Mα) . (3.58)

Differentiating this relation and comparing with (3.33), we obtained alternative definitions of absolute temperature, pressure and chemical potential:

ϑ=

S , p= −

V , µα=

Mα . (3.59)

Both the Euler relation (3.35) and the Gibbs-Duhem relation (3.36) are recovered in the same form, and one can proceed to the local forms as before, by applying the 1-homogeneity w.r.t total mass and total volume. The same form of local Euler and Gibbs-Duhem relations are recovered, plus we obtain the following two sets of alternative definitions:

ϑ=

∂η

¯

¯

¯

¯1

ρ,c1,...cN

, −p=

1ρ

¯

¯

¯

¯

¯η,c1,...,cN

, µα=

cα

¯

¯

¯

¯η,1 ρ,cβ6=α

(3.60)

for

e=e(ˆη,1

ρ,c1, . . . ,cN) , (3.61)

and

ϑ= ρce

(ρη)

¯

¯

¯

¯ρα

, µα= ρce

∂ρα

¯

¯

¯

¯ρe,ρβ6=α

(3.62) for

ρe=ρce(ρη,ρ1, . . . ,ρN) . (3.63)

Odkazy

Související dokumenty

• In Section 3, we introduce the Riemann–Hilbert problem for the orthogonal polynomials and transform this problem into one which can be controlled as n → ∞... • In Section 4,

After the determination of the time necessary for the adsorption and desorption, the experiments were carried out with a model mixture (5 % v/v ethanol-water mixture), which resulted

If we take nuclear prominence in the form of a separate nuclear tone as a possible cue for main clause status (as discussed in Section 2), we have to conclude that the

Finally, as an application of the abstract existence result, we demonstrate that by choosing the maps t 7→ γ 1 (t), γ 2 (t) in particular ways, we can recover the existence of at

Thus, in Section 5, we show in Theorem 5.1 that, in case of even dimension d &gt; 2 of a quadric the bundle of endomorphisms of each indecomposable component of the Swan bundle

We note that in the case m = 1, the class K 1,n (D) properly contains the classical Kato class K n (D) introduced in [1] as the natural class of singular functions which replaces the

The paper is organized as follows: in the next section (Section 2) we will discuss several technical results; in Section 3 we obtain a reformulation of the definition of

In Section 4, we prove that the eigenvectors of the transition matrix of neutral two-dimensional Markov processes can be decomposed as the product of “universal” polynomials (in