Thermodynamics - American Chemical Society

of a system reflects the extent to which the system can affect any other because the ... of the entity at a given time, and the influences imposed upo...
0 downloads 0 Views 2MB Size
13

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

A Thermodynamic Theory for Nonequilibrium Processes RICHARD A. GAGGIOLI Department of Mechanical Engineering, Marquette University, 1515 W. Wisconsin Ave., Milwaukee, WI 53233 WILLIAM B. SCHOLTEN Intertechnology, Inc., Warrenton, VA 22186 Traditional developments of thermodynamics defined entropy and hence thermostatic property relations - only for equilibrium states. This is because entropy has been defined only for states that can be connected reversibly to some common reference state. In the present approach, this restriction is overcome by eliminating reversibility in favor of a more general concept - called ideality - which includes reversibility as a special case. The key to the generalization is to base the Second Law on the property availability (1,2,3); the availability at any state of a system reflects the extent to which the system can affect any other because the state is not in equilibrium with some reference datum. A process is said to be ideal if there is no destruction of this capacity to affect other systems; that is, if there is no destruction of a v a i l a b i l i t y . Not every ideal process is reversible; for example, if a given state is not an equilibrium state (internally) it cannot be a state of a reversible process. However, it can be a state of an ideal process. In particular, it is the initial state of any process which would furnish its a v a i l a b i l i t y ; (i.e., any process which goes from the given state to one in equilibrium with the reference, and does indeed have the maximum effect on another system). Thus, the concept of ideality can be used to replace the concept of reversibility as a more general criterion for the perfection of processes. Another feature of the present theory is that it provides a formalism for deducing a complete mathematical representation of a phenomenon. Such a representation consists, typically, of (1) Balance equations for extensive properties (such as the "equations of change for mass, energy and entropy); (2) Thermokinematic functions of state (such as pv = RT, for simple perfect gases); (3) Thermokinetic functions of state (such as the Fourier heat conduction equation • 240, which would not change the number of oxygen atoms. Governing Equations for a Simple Compressible Flow A continuum flow (or deformation) is said, here, to be simple if there is only one scalar constraint and it is conserved. To f a c i l i t a t e the discussion of balance equations, each con­ straint will be modeled by a three dimensional Euclidian manifold. (Here "manifold" can be taken as a synonym for "space" as in (4, Pg 9).) Using this scheme, processes are modeled by employing a parametric transformation between each manifold and the "manifold of observation," or reference manifold. The parameter, to be de­ noted by σ , is called the process parameter; since a process oc­ curs as σ varies continuously from one value to another. (For a model of a "real" process, σ would normally be taken as time, t.) Here, the reference manifold will be called the volume mani­ fold, and the manifold for the single scalar constraint will be called the mass manifold. The word place will be used to mean point in the volume manifold. Then the balance equation (or " l o ­ cal equation of change") for an extensive property of a continuum is an expression for the rate of change of the amount of that prop­ erty per unit of volume at a particular place. (In these continu­ um models, balance equations can be integrated over a region R in the volume manifold to yield a balance for the region, which equates the rate of change of the amount of the extensive property in R to the sum of a net rate of transport function and a net rate of pro­ duction function. In applying the theory to Lumped Parameter mod­ els , these region balances are written directly without recourse to.

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

13.

GAGGiOLi AND SCHOLTEN

Nonequilibrium

Processes

209

the local equations of change. See (2) for further elaboration.) Thus, the equation of change for mass is f=-V.[ V]

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

P

(1)

where ρ is the (mass) density ( i . e . , the amount of the mass mani­ fold "constraint" per unit amount of the volume manifold and pV is the mass flux. V is called the (mass) velocity and re­ flects the transport of the mass manifold through the reference manifold. Similarly, the local equation of change for energy is written

4 -

= -V[peW]

(2)

where e is the energy per unit mass, and W is the energy velocity. The local equation of change for availability follows the same pattern as those for energy and mass, with the addition of a production rate per unit volume term ap. Thus

= - V-[paZ] + ap

(3)

where a is the availability per unit mass, and Ζ is the availabil­ ity velocity. Because the equations of change will be complemented by con­ stitutive relations, it is advantageous to re-write the balance equations to reflect changes at a "particle of matter," rather than at a "place in space." This is accomplished by using the "hydrodynamic" or "substantial" derivative. (Nevertheless, the tensor nature of the quantities in these equations is such that they are "volume tensors" and not "mass tensors." See (5) for elaboration) Thus, ρ 5X = VV Do

(4)

where ν Ρ

Ξ

1/p

(5)

- - V.{pe[W-V]}

(6)

—Da = - V.{pa[Z-V]} + 4

(7)

Do

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

THERMODYNAMICS:

210

SECOND L A W ANALYSIS

Note that [W-V] and [Z-V] reflect the rates at which energy and a v a i l a b i l i t y respectively are moving relative to the mass mani­ fold. By Postulate I there is an energy transport associated with mass transport. By definition, V reflects the transport of the mass manifold through the volume manifold. Hence, the rate of en­ ergy transport with mass transport through the volume manifold is given by £'V, where £ is a second order tensor. Then

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

peW = σ-V + £

(8)

defines £ — which can be called the "heat flux" — as any energy flux over and above that associated with motion of the mass mani­ fold. Furthermore, it is customary to a r b i t r a r i l y break down the portion £*V into two parts: (1) a part representing energy car­ ried in the mass — i . e . , the energy per unit mass e times the mass flux pV — plus (2) a part for the energy flux (associated with the flow) relative to the mass flux, denoted as TT-V — which can be called the "work flux." This is accomplished without intro­ ducing additional variables (i.e., the variable £ is replaced by π) , by defining ττ as 1

Ξ

ο - pe£

(9)

where £ is the fundamental tensor; JT is called the stress tensor. Hence £*V = π-V + peV

(10)

The expression for the energy flux for all simple flows is then peW = peV + π-V + £

(11)

and the equation of change for energy for all simple flows is Ρ ]§• - " ν-[π-ν + £ ]

(12)

Similar to Eq. 8, the a v a i l a b i l i t y flux can be expressed as a portion associated with the mass transport, ω-V, plus an "excess" jr i . e . , paZ = ω-V + r

(13)

As before, can be broken up into two parts, one of which repre­ sents the availability carried in the mass — paV Η Ξ ω - paj>

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

(14)

13.

GAGGiOLi A N D SCHOLTEN

NonequiUbrium

Processes

211

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

Furthermore, inasmuch as the a v a i l a b i l i t y contained by a vol­ ume manifold is the portion of the energy content of the manifold which could be transmitted out of the manifold were the "contained" manifolds brought to complete stable equilibrium, then the work flux of a v a i l a b i l i t y (through a particle) will be related to the work flux of energy by

where is the stress tensor (at the particle) in the hypotheti­ cal complete equilibrium state. (Equation 15 may be explained ( 1 , 2 4 ) with the physical argument that TTQ = PQ£ where PQ is the normal compressive stress in the reference atmosphere — atmos­ pheric pressure. Then PQV is the portion, of the "work flux" of energy, which would be delivered to the atmosphere; hr - PQ£]·V - the energy flux over and above that delivered to the atmosphere- can be supplied to any other system ( i . e . , it is totally avail­ able).) Thus the equation of change of a v a i l a b i l i t y for simple flows is written as Da ρ — = - ν·[π·ν - π - V + r] + a DO —U — ρ

(16)

Next, the thermokinematic equations of change will be obtained These are found by considering ideal relaxation processes, i . e . , hypothetical ideal processes from any and every state visited dur­ ing the real process to an arbitrary dead reference state. The foregoing equations of change are combined, for such a process, to eliminate all terms without parameter derivatives of properties; thence, the existence of functions relating the properties (i.e., the "state principle") can be deduced. In physical terms, a simple compressible f l u i d is one which cannot exert shear stresses without viscous dissipation. In the­ oretical terms, it is defined as a f l u i d wherein, during ideal processes, all energy and a v a i l a b i l i t y transports with mass trans­ ports are collinear with V; i . e . σ_ = σ£ and ω = so that σ/V = oV and ω-V = u)V (and ττ-V = pV and Ξ/V = [p - PQ]V) . Then, for ideal relaxation processes Eqs. 12 and 16 become, respectively p §

= -?-[pv + i ]

(17)

P

= - V-[{p - p }V +r]

(18)

and ~

Q

where the symbols p, c[ and _r denote ideal process variables. ing Eq. 4, Eqs. 17 and 18 can be rewritten as

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

Us­

212

THERMODYNAMICS:

p

S

- - pp

p

S

- -

ë -4

-

v

SECOND L A W ANALYSIS

- i

(

1

9

)

(

2

0

)

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

and [

P -

- 4

S

P

7 [

-

p

o]

-

v

4

respectively. Thus, two independent equations remain, but with three terms which have no parameter derivatives of properties. Note that the terms V-Vp can be eliminated by restricting consideration to processes with ν Ξ 0, and that the terms and .V_r can be eliminat­ ed by consideration of processes with c[ Ξ 0. During any ideal pro­ cess with £ Ξ 0, the quantity e

s is constant.

Ξ

e+

PQV

(21)

- a

For, combining Eqs. 4, 12, and 16 yields

P j g - -

V . f c - r] - a

(22)

p

and the right hand side is zero for ideal processes that are adiabatic — i.e., which have c[ Ξ 0. (If _r φ 0, then c[ φ 0, by Post­ ulate 1(a) .) It remains, now, to combine Eqs. 19, 20 and 22 for ideal pro­ cesses in such a way to eliminate c[ and jr. For ideal processes, define α by _ V-X α =

(23)

Then from Eq. 22, for ideal processes, p j g = - [1 - β]ν·Ι

(24)

Thus, Eqs. 19 and 20 can be rewritten as p — = - pp — - V-Vp + 4— — Da Da — 1-a Da J,

(25)

F

D(a-p v) 0V/ n

p

4

Dv P

p



π Γ

D 4 - 4

v

ideal , , ρα DsI

t P - P o

1

+

i 4 4 J

, ,v 0

l

d

e

a

l

(

2

6

Consider f i r s t all those particle states that can be connec­ ted to the dead reference state while V Ξ 0. Then, for such

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

)

13.

G A G G i O L i AND SCHOLTEN

NonequiUbrium

Processes

213

ideal processes, the terms V*Vp and V ' V P Q are zero in Eqs. 25 and 2 6 . The derivatives in these equations are derivatives at a fixed particle. Inasmuch as e,v, and s are properties, the change in value of each, from any one of the states at hand to the dead re­ ference state is independent of the intermediate states visited. Hence it follows from Eqs. 25 and 2 6 , with V = 0, that there exist functions

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

= u(s,v)]

(27)

I E 0

- ρ ν = b(s,v)]

v Ξ

(28)

0

It also follows that, for these states, Ρ

:

=

P=-- -(s,v)

s

v

I4 8l< * >

8

( 3 0 )

Note that ρ and α are thus functions of properties, so they too are properties. The properties s, ν and all functions thereof are called internal properties (thus u and b are called internal en­ ergy and internal a v a i l a b i l i t y respectively). The functions in Eqs. 27 through 30 are called thermostatic constitutive relations or internal "equations" of state — or better, internal functions of state. Suppose, now, that starting from the dead state (where s Ξ S Q and ν Ξ VQ) the velocity V is ideally changed, while s is held constant — i . e . , while £ Ξ 0 — and while ν is held constant. Then, the dependence of e and a upon V must be reflected by the invariants of V inasmuch as a and e are scalar quantities. The vector V has only one independent invariant; select Then De Pr-

=

Ρ

Da Ρ

Da

7— 0 σ

2

3a

Ρ

τ4Γ~

5 3 ( v

=

D(V )

2j

Da

3a

D(V )

() 31

2

7>

2

,~ . (32) 0

ϋ σ

3(ν )

A process of a simple compressible material is here said to be non-relat i v i s t i c if 9a/3(V ) and 3e/3(V ) are independent of s,v, V. Then symbolize these quantities by 1/2 k and l/2k respec­ tively: 3a_, _ (33) 2k 9(V ) s,v 2

2

1

2

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

THERMODYNAMICS:

214

3

2

(v )

s,v

SECOND L A W ANALYSIS

( 3 4 )

-ae

for non-relativistic simple compressible processes.

Then

2

e = u(s,v) + V /2k

(35)

e 2

a - p v = b(s,v) + V /2k Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

Q

(36)

a

Equations 35 and 36 are called thermokinematic functions of state. (Note that the variable s was introduced along with Eq. 23 in order to f a c i l i t a t e elimination of c[ and jr from Eqs. 19 and 20 re­ spectively. A more natural way to eliminate these variables would be to simply multiply Eq. 19 by a, then subtract from Eq. 20. In the latter procedure entropy s would never be defined, rather a function for "internal a v a i l a b i l i t y " b(e,v) would arise. The choice of introducing s was made in order that the traditional re­ sults would be obtained.) The existence of the relations given by Eqs. 27, 28, 29 and 30 can be called the internal state principle. The existence of the relations given by Eqs. 35, 36, 29 and 30 can be called the non-relativistic thermokinematic state principle. (Note the role of α in the foregoing developments of the thermokinematic functions of state: By introducing a, covariant derivatives could be elim­ inated in favor of parameter derivatives of properties. Thus, for example, Eq. 25 is obtained in l i e u of Eq. 19. Then by considera­ tion of line integrals of the resultant equation from any given state to a common reference state, because the integrals are inde­ pendent of the path, the functions of state can be deduced. For example, Eq. 27 is deduced via Eq. 25. Hence, α can be called an integrating factor.) Next, the thermokinetic equations of change will be deduced. This is accomplished by deriving an expression for the availabil­ ity destruction for real processes, and then invoking Postulate II, which decrees that a process will occur whenever it would de­ stroy a v a i l a b i l i t y . Consider that, differentiating Eq. 36 and using Eqs. 4 and 30 in the result yields +

"δ-ν·ϊ ά"

£ - * * · Ϊ

+

Ι-*·ΪΙ

(37)

a

Let 3. be the second order tensor such that r = 3-£

(38)

Then, substituting Eqs. 22 and 38 into the difference between Eqs. 37 and 16, and solving for a gives Ρ

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

13.

GAGGiOLi A N D S C H O L T E N

a P

NonequiUbrium

Processes

= V-[j3-£] - aV-£ - p[l-a]V-V DV + p[l-a]V«— + [1-α]ν·[π·ν] — Da

215

(39)

(By proper selection of scales (8) for a v a i l a b i l i t y , velocity and mass, k can be set equal to unity; for convenience, this has been done in Eq. (39) and will be done throughout the remainder of this paper.) Equation (39) can be rearranged into a

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

a

p

= £·[ν·β] + [3 -ag]:VcL

DV (40) + [1-a] [u -p£]:VV+ [l-a]V- ρ — + V-π Da — Suppose that the material is opaque (_3) ; i . e . , that the local availability destruction at a particle is independent from that of the neighboring particles and depends only on the local internal state and the variables in Eq. 40. Also assume for the present that there is no coupling between the terms of Eq. 40. This as­ sumption is made only for simplicity; it can be relaxed, but that complicates the following logic and conclusions. Consider f i r s t , the special case such that, V Ξ VC[ Ξ 0. Then, as will be discussed presently, it follows from Postulate II that +

K

DV ρ

—= -

(41) V-π

3 = α&

(42)

4 = L(s,v,Va)

(43)

τ_ Ξ π - ρ£ = A(s,v,VV)

(44)

The foregoing are deduced from Eq. 40: If Eq. 41 were not satisfied, then it follows from Eq. 40 that a v a i l a b i l i t y would be destroyed were V non-zero; by Postulate II, V would then be non­ zero. But this refutes the supposition that V = 0, and hence Eq. 41 must be satisfied. Similar logic leads to Eq. 42. To deduce Eq. 43 note that since Va = V-3. t 0, on the basis of Eq. 40, Posulate II says that there will be c[ φ 0, when Va φ 0, to destroy a v a i l a b i l i t y . Thus L yields zero when Va = 0. And, when Va is non­ zero, then c[ is non-zero and has the opposite direction to Va. Equation 44 is deduced analogously. The special case at hand is the case when V is small — in accord with the earlier r e s t r i c ­ tion, v i z . non-relativistic systems — and when the invariants of Vc[ are small. Equations 41, 42, 43, and 44 are called the thermokinetic functions of state for this special class of problems at hand. Equation 41 is recognized as the traditional equation of change for momentum. 2

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

THERMODYNAMICS:

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

216

SECOND L A W

ANALYSIS

Now a complete set of governing equations is given, for ex­ ample, by (a) the four thermokinetic functions of state above, along with (b) the thermokinematic functions of state — Eqs. 27 and 28 —, (c) the local equations of change for mass, energy, and a v a i l a b i l i t y — Eqs. 4, 12 and 16 —, and (d) the "auxiliary" re­ lationships given by the symmetry of ττ, the definition of (see Eq. 44), and Eqs. 5, 35, 36 and 38. (The equation of change for entropy — equation 22 — can be used in l i e u of that for either availability or energy.) It should also be mentioned that, for L and Λ isotropic fonc­ tions, truncated Maclaurin expansions of Eqs. 43 and 44 and use of Eq. 29 to change the independent variables from (s,v) to (α,ρ) re­ sults in the traditional forms for Fourier's law of heat condition and Newton's law of viscosity, v i z . £ = - κ(ρ,α)να

(45)

and f

Σ = λ(ρ,α) [V-V]g - μ(ρ,α) [VV + VV ]

(46)

(It is readily shown that α defines a Temperature scale.) The theory has thus derived these "laws" as truncated expansions of a Maclaurin series, and shows that transport properties are depen­ dent upon (α,ρ). The theory also predicts the existence of (α,ρ) dependence for u,b, and v. However, the functions themselves must be determined v i a experiment. The methods for establishing these functions from experimentally measurable functions are not a part of this paper, inasmuch as they follow the lines presented in the typical presentations of thermodynamics. Governing Equations for a Simple Compressible Reactive stituent Flow (with Internal Constraints)

Multicon-

The case to be considered here is that of a continuum mixture of chemical species in which chemical reactions occur. (A lumpedparameter model is given in (2).) The model for each different chemical compound is called a constituent. Were the constituents inert, then the amount of each would be an external constraint. However, since compounds can be produced or destroyed by chemical reactions in the mixtures being considered, the amount of each constituent is a non-conserved property. Then, if a mixture con­ tains Ν constituents, and if R independent reactions are allowed to take place, Ν - R of the constituents can be selected as com­ ponents ; and the amount of each component is a conserved property, i.e. its value can change only by being transported — it cannot be produced. Thus, every component is a constituent, but not every constituent is a component. However, the component amount, and the constituent amount of the same species are different. The amount of a constituent in a mixture at any instant represents the

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

13.

GAGGiOLi A N D S C H O L T E N

NonequMbrium

217

Processes

actual amount present. On the other hand, given a set of compo­ nents, the amount of each component does not represent the amount present (except in inert mixtures), rather it represents the amount of that component that would be required to produce the mixture from that component set. A more elaborate discussion on the distinction between constituents and components is given in (3, pp. 193-194). The number of components C in a mixture equals the number of constituents Ν minus the number of independent reactions R. It is always possible to select R independent reactions which are socalled formation reactions: A formation reaction is one which yields only one constituent, from all the components. For the for­ mation-reaction of constituent 3, the stoichiometric coefficient of 3 is unity; the stoichiometric coefficient of component α is denoted by ί-νρ }, α = 1,..., C. The amount per unit volume of component α at place ρ at "time" σ will be designated Ύ (ρ;σ) and the amount per unit volume of constituent 3 at place ρ at "time" σ will be designated pg. The α

α

C ρ(ρ;σ)

Ν γ (ρ;σ) = 4

Ρ (ρ;σ)

α

α=1

(47)

ρ

3-1 1

then represents the total amount of "materiaï per unit volume at place ρ at "time" σ. (For simplicity of the presentation to f o l ­ low, these amounts are expressed in " i n e r t i a l " units (e.g. grains, pounds mass) rather than "mole units (e.g. gram moles, pound moles).) Hence the set (ρ,γ-ρ ... Yc) Is not an independent set of properties; any one of these can be found in terms of the other C. In addition, y and x4 will designate the fraction amounts of component α and constituent 3 respectively; that is 11

a

Ύ Υ (ρ;σ) 'α4'

χ

3

(ρ;σ)

Ξ

τ -

α = 1, ... C

(48)

3 = 1, .. .Ν

(49)

ρ(ρ;σ) Ρο(ρ;σ)

( ρ ; σ )

It follows that

y =xV Y.->:».-* > > - ι a

β

v

o r

β

s e t

Ύ

Thus the set (p,y4> ··· c - l ) 4 ÎYl» ···» (Ρ · is equivalent to any independent set from (ρ,γ4,... γ4).

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

(50)

THERMODYNAMICS:

218

SECOND L A W

ANALYSIS

The fraction amount of each constituent 3 which is not a component, is called the formation reaction extent for that spe­ cies ; it represents the extent to which the formation reaction for that species has proceeded. The amount of component α which has been "used up" by reaction 3 to form constituent 3 is given by V 3 Xfi" It then follows that the fraction amount of component α is related to the fraction amount of constituent α by Ν

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

y = χ 'α α

> ν χ Ζ—/ 3α 3 3=C+1 Λ

Λ

α = 1, ... C

(51)

where species "C+l through Ν" are taken to be those constituents which are not components, and species "1 through C" are taken to be the components. This notational convention will be used throughout the rest of this section. An equilibrium state is one which would not change were it isolated. Thus, considering the intensive states of the model at hand, if the specific energy, velocity, specific volume, and the amounts of each component in a set are uniquely specified, then no interaction with the environment would be possible. Hence, if con­ sideration is restricted to simple compressible equilibrium states, they are fixed by specification of (e,V,v,y4, y4-l)· Hence, considering only equilibrium states, the amount of each constitu­ ent is fixed by specification of (e,V,v,y4, ..., Yn]) · If it is desired to consider also states which are not at re­ active equilibrium — which could reach the equilibrium states via chemical reaction at fixed (e,V,v,y4, y4 4) — then s p e c i f i ­ cation of these variables clearly does not suffice to f i x the con­ stitution. Although, by definition, the states being considered are defined by (e,V,v,x , XJJ_4) » convenient to describe them with (e,V,v,y4,.. ., vc_].) P-4 fonnationreaction extents, i.e. with (e,V,v,y-L, _-L> xcdb-l» * * ' N 4 " In this section the "total amount of components f u l f i l l s the same role as "mass" in the preceding section. Hence Eq. 4 is now regarded as the balance equation for the total amount of compo­ nents. For compatibility with this, the velocity V is now de­ fined as it:

w i l 1

b e

m

o

r

e

1

us t b e

V C

, X

Ν Ρ V α—α

α=1 and

(52)

is the velocity of the ath constituent manifold. An equation of change for the formation reaction extent χ Ρ OF

+ P g r

v

- - *J

e

β = C + 1

Ν

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

is (53)

13.

GAGGiOLi A N D S C H O L T E N

Nonequiltbrium

219

Processes

where

3[V3

h

1, ...,

ΞP

(54) Ν

Note that Eqs. 52 and 54 yield Ν

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

Σ

j

a

(55)

0

α=1

Thus J3 represents the flux of constituent 3 relative to the bulk velocity of the mixture ( i . e . the diffusion flux — and this has been defined for all constituents), and r4 represents the net production rate of constituent 3 per unit amount of mixture. The presence of this production term reflects the fact that the con­ stituent amounts are not conserved. If constituent α is also a component, then there is a possi­ b i l i t y that the species can be destroyed by more than one reaction. Hence the equations of change for x , a = 1, ...C., will appear as a

Ν

Dx α = - V-J + e

0-1,

Ρ

Da

0=C+1

(56)

The term vg rg represents the net production rate of constituent α by formation reaction β per unit amount of mixture. An equation of change for the fraction component amount y is a

a

~ Da

= - V-j

v

+ >

_a

a

V-j

/ 3 =C+13a J

J

1,

(57)

Note the absence of production terms, which is to be expected since component amounts are conserved. Next, balance equations for energy and availability are writ­ ten as Ν De P

D?

π-V +

era

α=1 and

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

(58)

THERMODYNAMICS:

220

SECOND L A W ANALYSIS

Ν

V. [π- ].ν+2][6 -« ]£α + r + aP Ρο&

α



(59)

a=l

In writing these the volume plus the amount of each consti­ tuent except the C , i.e. {ν, x4, ...Χς_ΐ, Χ 4 + Ί > ··· XflK were taken as an independent set of constraints for the system. (Alter­ natively, one could take the amount of each of the Ν constituents as the independent constraints. In one case, the Ν - 1 consti­ tuent manifolds and the overall mass manifold are selected as in­ dependent; in the other case, the Ν constituent manifolds are se­ lected.) A term reflecting the flux of energy and availability with the flux of each of these constraints was incorporated in each of the respective balance equations. That such terms exist follows from Postulate I and II. A further assumption underlies these balance equations, namely that only the volume transport, represented by pV, has inertia, and that the effects of this iner­ t i a are reflected by the same momentum equation as that in the preceding section; v i z . Eq. 41. Were there inertia associated with, for example, the flux j , then the scalar coefficient 6 would have to be replaced by""â second order tensor. Consistent with this assumption is the further assumption that during ideal processes

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

ttl

a

1

or α

α

J

a

(60) α J

ideal

The next step in order to derive the pertinent thermokinematic functions of state, is to write Eqs. 58 and 59 for ideal processes, and then combine the results to eliminate covariant derivatives of fluxes in favor of coefficients times parameter derivatives of specific intensive properties. In the previous section this could be accomplished merely by multiplying one equation by α and then subtracting the other. Before that can be done here, however, the terms in the summation signs must be expressed in terms of parameter derivatives. This is done by using Eqs. 56 and 57. The results are

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

13.

GAGGioLi A N D S C H O L T E N

ΝonequiUbrium

Processes

221

Ν

+ p

Σ

1 Da

(61)

3

3=C+1 and C-l Db PP



Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

6

]

S Σ ~~~0α 5 Γ

ideal

+



a=l

Ρ

V-r

Da

e=c+i

(62)

where C-l λ

3

Ξ

δ

3

a 3a

+

3 = C=l,

(63)

a=l and u and b are defined as in the previous section (Eqs. 35 and 36. (It is readily shown that is the chemical a f f i n i t y of re­ action 3.) It still remains to express the R reaction rates in terms of property derivatives. This is done by again employing a scheme analogous to Eq. 23, that is, define

ε

Λ

3 = C + 1,

= r

(64)

ideal Thus εg is seen to represent the rate at which constituent 3 is being "formed" by reaction to the total rate of change of that constituent. Furthermore, if the mixture is inert, then εg = 0 for all 3 = C+l, ... N, and each formation reaction extent (each X3 3 = C+l,... N) becomes a conserved property; therefore the ideal process is reversible ( a l l the constraints are external — i.e. conserved). Substituting Eq. 64 in turn into Eqs. 61 and 62 yields S

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

222

THERMODYNAMICS:

Du Da J

Ρ Ρ

Dv , 5α" +

a.

p

ideal

SECOND L A W ANALYSIS

Da

a=l

Ν

Dx„ (65)

3=c+i Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

and, (using that 6 4 is a constant for a particle) C-l b+

6

y

Σ 0cc a a=l Da

y

Dv , P P

J

° a

P

+

ία"

2 A 4 a=l

ideal

+ Ρ

V-r

(66)

3=C+1 Now Eq. 23 can be employed to eliminate the remaining nonparametric derivatives, v i z . V»c[ and V-_r. Multiplying Eq. 65 by α and subtracting the result from Eq. 66 gives, after some re­ arranging : C-l

Ν λ

χ

0β β β= C+1

α=1 Do

C-l Du f

1

Da

'

/

J

a Da

a=l Ν +

p

Σ 3=C+1 L

α 1



[ ι

-

ε

β

]+

Vo6

Da

(67)

Now, as before, it follows that there exists a function f, such that

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

13.

GAGGiOLi A N D S C H O L T E N

NonequiUbrium

223

Processes

C-l + b

+

Σ?ΟΛ Σ

A

α=1

x

oe 3

S=C+1

(68)

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

and (69)

α = f (u,v, ... ) 1

(70)

-[1 - a]p = f (u,v, ... ) 2

[1 - α]δ = f α

[1-α]λ [1-ε ] + ρ

β

e

B

A

œ

-

a + 2

f 3

+

1

( u , v , ... )

(

u

>

v

>

··· )

α = 1,

»

=

c

+

1

>

C-l

N

···

(71) (72)

By procedures similar to those used in obtaining Eq. 40, the rate of availability destruction per unit volume is found to be [1 - a] J [π - pgj: VV

+ 4 V ™

a

a=l a# +

Ρ

Γ

β

ε

β Da

3

1-a

3=C+1

[6-6 ]V-j + -f=- •Va a a a 1-a J

(73)

a=l a# From this the following thermokinetic functions of state are deduced: £ =_q(u,v,y,x,Va,V6)

(74)

14 = 4(u.v.y.x.VO.Va)

a = 1,...C-1, C+l,...Ν f

π-pjg = x(u,v,y,x,[VV + VV ], A,V-j_) 6-6 a a = φa(u,v,y,x,V-1,X,V-V) > 4 » » —

(75) (76)

a = 1,...C-1, C+l,...Ν (77)

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

224

THERMODYNAMICS:

SECOND L A W

ANALYSIS

_Dx

(

e

V β

Da

R (u,v,y,x,X,V.j,V-V), β β = C+l ,.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

0

C+l

9 · · ·

Ν

(78)

· ·Ν

(79)

(In the l i s t s of independent variables, the subscripts have been deleted from x,y,6,j4 etc. for convenience.) The deduction of these equations from the above expression for ap, is analogous to that employed previously. Note that Eq. 79 follows from the fact that the intensive state of the complete stable equilibrium configuration of a particle is constant and that 644 represents the amount of energy transported. In summary then, a complete set of governing equations is given by: Equations of change: Eqs. 4, 56, 57, 58, 59 and 41. Thermokinematic functions of state: Eqs. 68 through 72. Thermokinetic functions of state: Eqs. 74 through 78. Auxiliary equations: Eqs. 5, 35, 36, 38, 50 and 55. These constitute (22 + 7R + 6C) equations in the (22 + 7R + 6C) unknowns p, v, e, u4 a, b, a, p, a , V(3), j (3R+3C), r«(R),6 a (C-l), ( 3 ) , _β(3), A (R), e (R), ô4(C-l) , y4C) , x (R). Also, if the six equations for Da/Da, f(u,v,y,x), _r = a = b + p4v + y2/2k are deleted then the remaining system could be solved for all the variables except a,b,ap, _r. It should also be mentioned that other alternative sets of independent governing equations could be selected, via changes of variables. For example, simple rearrangement of Eq. 67 would y i e l d a function for s = b - u rather than for b. With such alternatives, different but theoretically equivalent sets of governing equations could be deduced; which would prove to be the more easily and more economically employed sets would be largely a matter of experience and insight. In any case, the procedure outlined in this paper resolves the basic d i f f i c u l t y : finding a set — any set — of pertinent governing equations. (Note, though, that the procedure does not yield the functions of state, but only t e l l s what their variables will be. But simply knowing what they will be greatly simplifies the determination, via experimentation, of the functions themselves.) The most notable difference between the equations developed here and the standard reaction equations, is the appearance, for each formation reaction, of the variable in the governing equations. The new variable enters as an integrating factor which is needed because the states are far from equilibrium. This is in contrast to traditional theories (e.g., 64, Chapter X), wherein it is assumed that all states, regardless of composition, can be connected reversibly. The traditional theories have proven valuable— i.e., the assumption "works out" — for predicting equilibrium compositions; see (_2, Appendix B.5). But at best, "empirical modi f i c a t i o n s " of these traditional theories have been required to a

£

3

3

3

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

13.

GAGGiOLi A N D S C H O L T E N

Nofiequilibrium

Processes

225

obtain good results for reacting mixtures. This failure of the traditional theories can be explained from the results obtained in this paper, in that the traditional theories regard the a f f i n ­ ity — an "equilibrium" variable — as the driving force for reac­ tion, whereas the results here (viz. Eq. 78) show that ε g must also be included. Thus the theory given here includes not only the traditional states of equilibrium composition, but also the highly nonequilibrium states which are encountered in "real" chem­ i c a l reactions.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

Closure It has been illustrated how, by replacing the reversibility concept by the more general concept, ideality, a thermokinematic and a thermokinetic state principle can be derived which is applic­ able to nonequilibrium states. The key to replacing reversibility by ideality is to base the second law on availability; the key to implementing this for the purpose of deducing property relations is the introduction of an integrating factor, in analogy to tem­ perature. These state principles along with the extensive proper­ ty balance equations provide a system of governing equations. Abstract In thermodynamics, entropy has been defined only for equili­ brium states. In the present approach this is overcome by replac­ ing the concept of r e v e r s i b i l i t y in favor of a more general concept, called ideality, which includes reversibility as a special case. There are two keys to this generalization, whereby a rational definition for entropy of nonequilibrium states is obtained. F i r s t , a more general second law is postulated, based on the property a v a i l a b i l i t y . (The availability at any state of a system reflects the extent to which it could affect any other system.) The second key is the introduction of other integrating factors, in addition to temperature, in order to deduce the fundamental differential property relationship ( i . e . , Gibbs equation.) Another feature of the present theory is that it provides a formalism for deducing a complete mathematical representation of a phenomenon. In particular, beginning with the balance equations for (1) the pertinent constraints (e.g. volume, alone, for "simple compressible flows"), (2) the energy and (3) the a v a i l a b i l i t y , the needed complementary equations are deduced; namely, the equations reflecting i n e r t i a l effects (e.g., the momentum equation), and the constitutive relations. In this paper, the theory is applied to two specific cases. The f i r s t case is simple compressible flows to i l l u s t r a t e the theory in familiar circumstances. (Where some assumptions inherent in the traditional governing equations also come to light.) The second case is chemically reactive flows; here the added generality of ideality over r e v e r s i b i l i t y is need­ ed, since the states visited are not at internal equilibrium. As

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.

226

THERMODYNAMICS:

SECOND

LAW

ANALYSIS

desired, the governing equations obtained distinguish themselves from those currently employed, by the appearance of relations involving additional integrating factors, besides temperature. Other highly nonequilibrium phenomena for which the added generality of the proposed theory could be f r u i t f u l , inasmuch as the contemporary models are not satisfactory, are multiphase metastable flows and viscoelastic flows.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on October 16, 2014 | http://pubs.acs.org Publication Date: May 28, 1980 | doi: 10.1021/bk-1980-0122.ch013

Acknowledgment The authors would l i k e to acknowledge the counsel of Professor Edward F. Obert. Those who are familiar with (3) will recognize its influence, which was crucial. Literature Cited 1. Obert, E.F. and Gaggioli, R.A. Thermodynamics. McGrawNew York, Second Edition, 1963. 2. Gaggioli, R.A. and Scholten, W.B. "A Thermodynamic Theory for Non-equilibrium Processes I. Thermokinematics." Marquette University College of Engineering Report, Marquette University, Milwaukee, Wisconsin, Nov., 1970. 3. Hatsopoulos, G.N. and Keenan, J.H. Principles of General Thermodynamics. John Wiley & Sons. New York, 1965. 4. Sokolnikoff, I.S. Tensor Analysis Theory and Applications to Geometry and Mechanics of Continua. John Wiley & Sons, New York, Second Edition, 1964. 5. Lodge, A.S. Elastic Liquids. Academic Press, New York, 1964. 6. deGroot, S.R. and Mazur, P. Non-Equilibrium Thermodynamics. North-Holland, Amsterdam, 1962. 7. Wepfer, W.J. "Applications of the Second Law to the Analysis and Design of Energy Systems." Ph.D Dissertation, University of Wisconsin-Madison, 1979. 8. Moran, M. and Gaggioli, R. "Generalized Dimensional Analysis." U.S Army Mathematics Research Center Report No. 927, 1968. Hill,

RECEIVED

October

17,

1979.

In Thermodynamics: Second Law Analysis; Gaggioli, R.; ACS Symposium Series; American Chemical Society: Washington, DC, 1980.