Sunteți pe pagina 1din 13

Non-redundant and natural

variables denition of heat


valid for open systems
Juan Ramn Gonzlez lvarez
juanrga
C:/Simancas 26, E-36208 (Bouzas) Vigo, Pontevedra, Spain
C:/Carrasqueira, 128, E-36331 (Coruxo) Vigo, Pontevedra, Spain
http://juanrga.com
Twitter: @juanrga
2012, May 18, 19:21
Abstract
Although an unambiguous denition of heat is available for closed systems, the question of how best
to dene heat in open systems is not yet settled.
After introducing a set of physical requirements for the denition of heat, this article reviews the
non-equivalent denitions of heat for open systems used by Callen, Casas-Vzquez, DeGroot,
Fox, Haase, Jou, Kondepudi, Lebon, Mazur, Misner, Prigogine, Smith, Thorne, and
Wheeler, emphasizing which physical requirements are not met.
A subsequent section deals with the main objective of this article and introduces a new denition of
heat that avoids the diculties of the existent denitions, providing (i) a complete distinction between
open and closed systems, (ii) non-redundancy, (iii) natural variables for the thermodynamic potentials,
and (iv) a sound, complete, and intuitive generalization of classical thermodynamic expressions.
1 INTRODUCTION 2
1 Introduction
The true nature of heat, as a form of energy that can interconvert to other forms of energy,
was established after much debate in the last part of the 19th century [1]. However, an
unambiguous denition of heat was lacking until G. H. Bryan introduced his denition in
1907 [2, 3]
Q dU W
TD
, (1)
with U and W
TD
being the internal energy and thermodynamic work, respectively. Fourteen
years latter, Born generalized this denition [4] when he considered variations in the total
energy E and total work W
Q dE W. (2)
The Bryan & Born denition (2) is not valid for open systems [1, 4] i.e., for systems
that can interchange matter as well as energy. This restriction has not been a diculty
in the past, because the basic problem of classical thermodynamics is the determination of
the equilibrium state that eventually results after the removal of internal constraints in [an
isolated], composite system [5, 6].
The diculties begin with the extension of classical thermodynamics to irreversible processes.
In modern thermodynamics [1], systems in a nonequilibrium thermodynamic state are divided
into small elements of volume and each element is assumed to be locally at equilibrium [7].
Now, these elements of volume can interchange matter with adjacent elements, which requires
a denition of heat valid for open systems as well.
The importance of a generalization of the closed-systems denition of heat has been empha-
sized many years ago; however, in despite of the existence of some proposals, the question of
how best to dene heat transfer in open systems is not yet settled [8].
The next section gives an introduction to the more basic formalism of modern thermody-
namics, including balance equations for general thermodynamic quantities, and introduces a
set of physical requirements for the denition of heat. The section 3 reviews the available
non-equivalent denitions of heat for simple thermodynamic open systems, emphasizing what
physical requirements are not met.
Section 4 presents a new denition of heat that avoids the main diculties of the existent
denitions. In section 5, the present author introduces the new denition for general thermo-
dynamic systems.
2 Physical requirements for the denition of heat
Consider a general thermodynamic quantity Y, whose density is denoted by y. The usual
local form of the balance equation for this quantity is [1]
y
t
=
Y
J
Y
, (3)
2 PHYSICAL REQUIREMENTS FOR THE DEFINITION OF HEAT 3
where J
Y
is the ux [9] of Y and
Y
the amount of Y produced per unit volume per unit
time. Specic expressions for
Y
for several quantities such as internal energy, amount of
substance, entropy, and others are given in the literature [1].
The change in the amount of Y in a volume V in a time dt can be obtained by integrating
(3) over the whole volume, multiplying both sides by dt, and applying Gauss theorem
dY = d
i
Y + d
e
Y, (4)
in which d
e
Y dt
_
A
J
Y
dA with dA the vector representing an area element is the
change in Y exclusively due to exchanges with the exterior, whereas d
i
Y dt
_
V

Y
dV is
the change produced by processes in the interior of the thermodynamic system.
Finally, it must be emphasized that modern thermodynamics does not rely on the use of
imperfect dierentials, because modern thermodynamics is dened over an extended thermo-
dynamic space that includes time [11]. This means that classical thermodynamics expressions
such as (2) are modernized to
dQ dE dW, (5)
with both dQ and dW well-dened and measurable physical quantities. Of course, this modern
denition of heat is also restricted to closed systems. A generalization to open systems will
be given latter in this article.
We consider four basic requirements for any physically admissible denition of heat: (i) a com-
plete distinction between open and closed systems, (ii) non-redundancy, (iii) natural variables
for the thermodynamic potentials, and (iv) a sound, complete, and intuitive generalization of
classical thermodynamic expressions. A detailed discussion of the requirements is as follows.
(i)
By a complete distinction between open and closed systems, we mean that the denition
of heat must consider the important fact that the entropy S and the internal energy U are
extensive thermodynamic quantities and both vary when the mass of the open system changes.
Regarding the internal energy [8]:
It is therefore expected that the usual version of the rst law of thermodynamics
for closed systems, namely dU = dQ + dW, will not be valid for open systems.
As Kondepudi & Prigogine remark [1]: For open systems, there is an additional contri-
bution due to the ow of matter dU
matter

dU = dQ + dW + dU
matter
. (6)
A similar extension is expected for the second law. Eectively, the DeDonder [12] entropic
term for open systems is [1, 4, 13]
d
e
S =
dQ
T
+ (d
e
S)
matter
. (7)
(ii)
A denition would not be devoid of distinctive physical meaning. A hypothetical denition of
heat as dQ
void
dU will be rejected because merely replaces one letter by another.
3 REVIEW OF AVAILABLE DEFINITIONS OF HEAT FLUX 4
(iii)
The denition of heat would use natural variables for the thermodynamic potentials instead of
non-natural variables. For example, the internal energy U is not a thermodynamic potential
in a temperature, volume, composition, (T, V, N), space.
(iv)
By sound, complete, and intuitive generalization, we mean that the denition of heat for open
thermodynamics systems would reduce to the corresponding classical expressions for closed
systems in a simple and natural way.
In the next section, we will verify that the available denitions of heat do not satisfy all the
requirements, conrming the non-equivalence found by Smith [8]:
In irreversible thermodynamics [...] there exist several denitions of heat ux
[...] although many authors employ non-equivalent denitions (see e.g. deGroot
and Mazur 1962)
3 Review of available denitions of heat ux
We begin with one-component elements of volume, at rest, that can interchange internal
energy and matter, without chemical reactions
N
= 0, and that verify the generalized Gibbs
equation [4, 7]
Tds = du dn, (8)
for thermodynamic temperature T, density of entropy s, density of internal energy u, chemical
potential , and mole unit per unit volume n.
The entropy ux J
S
is given by
J
S
=
J
U
J
N
T
, (9)
where J
U
and J
N
are the ows of internal energy and matter, respectively. This will be our
starting point in the review of the dierent denitions of heat ux proposed up to now. More
general systems will be considered in section 5.
A rst denition of heat ux is J
Q
DM J
U
. Using this denition, (9) can be rewritten as
J
S
=
J
Q
DM J
N
T
. (10)
The ux J
Q
DM is used by DeGroot & Mazur [13], Fox [14] and Jou, Casas-Vazquez, &
Lebon in irreversible thermodynamics [15]; by Jou, Casas-Vazquez, & Lebon in extended
thermodynamics [15] in this case with equation (10) generalized to an extended thermody-
namic space; and is the standard in the kinetic theory of diluted gases [13, 16].
This rst denition of heat ux does not satisfy the requirements (i), (ii), and (iv) presented
in the section 2. A detailed discussion is given next.
3 REVIEW OF AVAILABLE DEFINITIONS OF HEAT FLUX 5
Eectively, the ux J
Q
DM does not completely distinguish between open and closed systems.
Substituting J
Q
DM in the balance equation for the internal energy of a homogeneous thermo-
dynamic system, at rest and in absence of external eld, integrating over the volume V of
the system, and multiplying by dt, we obtain [13]
dU = dQ
DM
+ dW, (11)
where dW = pdV, with p denoting pressure. This expression disagrees with the rst law
for open systems (6) because (11) is missing a dU
matter
term for matter exchange.
The underlying physical reason why J
Q
DM does not distinguish between open and closed
systems, at the energetic level, is that the internal energy U is an extensive thermodynamic
quantity and, therefore, J
U
= J
Q
DM does not dierentiate between changes in the internal
energy due to ows of mass transferring the internal energy per particle in the ow and
changes due to genuine heat ows.
The ux J
Q
DM is redundant. Any instance of J
Q
DM in the equations could be substituted by
J
U
without physical or mathematical changes. Just compare (9) and (10). At the same time,
any instance of heat ux in the text of the above references [1316] could be reverted to
internal energy ux without any appreciable change. It is superuous to introduce this
concept of heat in the formalism. This defect is more evident when J
Q
DM is compared with
the other denitions considered in this work, which are non-redundant and really introduce a
physical concept of heat ux dierent from the physical concept of internal energy ux.
The formalism introduced by DeGroot & Mazur is not backward compatible with classical
thermodynamics. Indeed, they apply J
Q
DM to thermoelectric phenomena only after redening
internal energy see their equation XIII.32 in presence of electromagnetic elds [13]. Their
redenition is not compatible with the usual meaning of internal energy as the energy of a sys-
tem at rest in absence of external elds [4]. Their redenition is not considering the Coulomb
interaction energy between particles within the system, for example. As is now well-known,
the van der Waals equation for the internal energy contains a term due to the interactions
between molecules in a gas. Their redenition of the well-established concept of internal
energy is clearly motivated by their need to identify J
Q
DM with J
U
see the corresponding
redenition of heat ux in XIII.33 [13].
There are no objective reasons for using two dierent concepts of internal energy at once: one
in classical thermodynamics where internal energy includes the interaction energy between
particles and another concept when using J
Q
DM and dQ
DM
in modern thermodynamics.
These three disadvantages of J
Q
DM are a motivation for the search of improved denitions of
heat ux for open systems.
A second denition of heat ux is J
Q
C J
U
J
N
. Now the entropy ux (9) takes the simple
form
J
S
=
J
Q
C
T
. (12)
This denition of heat ux is used by Callen [5]. Misner, Thorne, & Wheeler use a
weird variant [17, 18] in curved spacetime thermodynamics in this case with (9) formulated
in curved spacetime.
This heat ux J
Q
C does not satisfy the requirement (i) presented in the section 2. Eectively,
3 REVIEW OF AVAILABLE DEFINITIONS OF HEAT FLUX 6
if we integrate (12) over the area A of an isothermal system, and multiply by dt we obtain
d
e
S =
dQ
C
T
, (13)
which disagrees with the DeDonder entropic term for open systems (7), because (13) is
missing the (d
e
S)
matter
term for matter exchange.
The underlying physical reason why J
Q
C does not distinguish between open and closed systems,
at the entropic level, is that the entropy S is an extensive thermodynamic quantity and,
therefore, (12) does not dierentiate between changes in the entropy due to ows of mass
transferring the entropy per particle in the ow and changes due to genuine heat ows.
Precisely Callen introduces his J
Q
C = TJ
S
in analogy [5] with the Q = TdS of
classical thermodynamics. The problem is that the classical Q = TdS is not valid for open
systems [1, 4, 12, 13], and the same physical defect is inherited by J
Q
C = TJ
S
.
It is interesting to remark a kind of complementarity between J
Q
DM and J
Q
C. If we only
consider the J
U
term, as J
Q
DM does, we have problems with the internal energy; at the other
hand, if we add the J
N
term, as J
Q
C does, we have problems with the entropy. A new
denition of heat that considers the extensive character of both entropy and internal energy
is evidently needed.
In a rst step of their study of open systems, Kondepudi & Prigogine apply a change
of variable, (u, n) (T, n), to the generalized Gibbs equation (8). The entropy, internal
energy, and chemical potential functions are substituted by s = s(T, n), u = u(T, n), and
= (T, n), respectively. As a consequence, the entropy ux (9) is modied to
J

S
=
J

U
J
N
T
. (14)
In a second step, Kondepudi & Prigogine [1] obtain the next identity
= u
n
Ts
n
. (15)
Above s
n
(s/n)
T
is a partial molar entropy and u
n
( u/n)
T
a partial molar energy.
Substituting the identity in the entropy ux (14) they nally obtain
J

S
=
J
Q
KP
T
+s
n
J
N
(16)
for a heat ux J
Q
KP J

U
u
n
J
N
. This J
Q
KP is used by Kondepudi & Prigogine [1], by
Haase [4], and by Smith [8].
This third denition of heat ux does not satisfy the requirements (iii) and (iv) presented in
the section 2. A detailed discussion is given next.
It is evident that the ux J
Q
KP cannot be dened in the thermodynamic space of natural vari-
ables. We can work in the natural energy-composition space when obtaining the entropy ux
(9) from the uxes of mass and internal energy, but we are forced to switch to a temperature-
composition space if we want to obtain the entropy ux from the heat ux and the ux of
mass (16).
4 NEW DEFINITION OF HEAT FLUX 7
There are no objective reasons for introducing a change of variables before using a concept
of heat ux.
The backward incompatibility with the classical expressions for closed systems can be shown
as follows. Using J
Q
KP in the balance equation for the internal energy of a homogeneous
thermodynamic closed system, at rest and in absence of external eld, integrating over the
volume V of the system, and multiplying by dt we obtain
d

U = dQ
KP
+ d

W, (17)
which introduces a departure from the ordinary rst law for closed systems dU = dQ pdV,
because dQ
KP
= TdS and d

W = pdV see also the boxed equation 2.2.12 in [1].


It is worth to mention that Kondepudi & Prigogine redene their heat ux J
Q
KP in presence
of electromagnetic elds see the equation 15.4.20 in [1]. Initially they consider a denition
of heat that satises J

U
= J
Q
KP + u
n
J
N
; however, this changes to J

U
= J
Q
KP + u
0
n
J
N
in
presence of elds, where u
0
n
is computed in absence of elds. Their more general denition
implies a weird mixture of eld-dependent and eld-less terms, even in the same equations!
Indeed, the above J

U
is the ux associated to u
n
, not to u
0
n
. Moreover, their treatment is full
of typographical errors and inconsistencies [19].
There are no objective reasons for using three dierent concepts of internal energy at once:
the ordinary U in classical thermodynamics, and both

U and

U
0
when using J
Q
KP in irreversible
thermodynamics.
After showing that the available denitions of heat ux are redundant, do not properly describe
the thermodynamics of open systems, and do not use natural variables for the thermodynamic
potentials introducing a departure from many classical thermodynamic expressions, we will
propose a new denition without such defects.
4 New denition of heat ux
We will continue considering the simple system of the section 3 for the sake of comparison
with the previous denitions. More general systems will be considered in the next section.
Our starting point is (9) again. A new denition of heat can be obtained by considering
thermal properties in the chemical potential that were not included in the previous denitions
J
Q
DM and J
Q
C recall that J
Q
DM ignored the J
N
term, whereas J
Q
C included this term without
any analysis. No change of variables like that in J
Q
KP is needed here. Our central element
will be the Euler equation associated to (8)
Ts = u + p n. (18)
Notice that we continue working in a thermodynamic space (u, n) of natural variables, because
= (u, n), s = s(u, n), p = p(u, n), and T = T(u, n) in the Euler equation.
Using (18), the entropy ux (9) can be rewritten as
J
S
=
J
Q
T
+
_
s
n
_
J
N
, (19)
4 NEW DEFINITION OF HEAT FLUX 8
with a new heat ux dened by
J
Q
J
U

_
+
Ts
n
_
J
N
. (20)
This new denition of heat ux satises all the physical requirements (i) (iv) presented in
the section 2. A detailed discussion is given next.
It is evident that this new heat ux is non-redundant and that introduces a concept of heat
ux dierent from the concept of internal energy ux. Moreover, the physical interpretation
is intuitive. For instance, the parenthesized term in (19) is a molar entropy [20]. Therefore,
(19) is telling us that J
Q
is proportional to a thermal entropy ow J

S
where the entropy
transfered through a mass ow has been subtracted
J
Q
= TJ

S
T
_
J
S

_
s
n
_
J
N
_
. (21)
Notice that TJ
S
= J
Q
only holds for closed systems, which means that J
Q
allows a physical
distinction between open and closed systems at the entropic level. Eectively, integrating (19)
over the area A of an isothermal system, and multiplying by dt, we obtain the DeDonder
entropic term (7) for open systems with
(d
e
S)
matter

_
S
N
_
d
e
N. (22)
The new J
Q
also allows a physical distinction between open and closed systems at the energetic
level because for open systems J
U
= J
Q
. This other advantage can be shown by using J
Q
in the balance equation for the internal energy of a homogeneous thermodynamic system, at
rest and in absence of external eld; by integrating over the volume V of the system and
multiplying by dt. The nal result is the rst law of thermodynamics for open systems (6)
with
dU
matter

_
+
TS
N
_
d
e
N. (23)
This new heat ux J
Q
is able to distinguish between open and closed systems completely,
because J
Q
considers the extensive thermodynamic character of both the entropy S and the
internal energy U. Eectively, (19) distinguishes changes in the entropy due to ows of mass
from changes due to genuine heat ows, whereas J
U
= J
Q
+ ( + Ts/n)J
N
distinguishes
between changes in the internal energy due to genuine heat ows and those due to ows of
mass. Notice that for simple closed systems J
Q
= J
Q
DM = J
Q
C.
Moreover, J
Q
is dened in a thermodynamic space of natural variables and allows for a smooth
generalization of classical thermodynamic expressions to open systems. For closed systems
(d
e
S)
matter
= dU
matter
= 0 and the ordinary rst law for closed systems dU = dQ pdV is
recovered because dW = pdV and dQ = TdS.
The main advantages of J
Q
over the previous denitions are (i) complete distinction between
open and closed systems, (ii) its non-redundancy, (iii) its use of natural variables, and (iv)
that it properly generalizes classical thermodynamic expressions.
In the following section, we will present the new general denition of heat. We will also show
further advantages, such as its ability to deal with general thermodynamic systems without
any need to redene the standard concept of internal energy.
5 HEAT FOR GENERAL THERMODYNAMIC SYSTEMS 9
5 Heat for general thermodynamic systems
Instead considering a simple system as we did in the section 3, we will consider now a generic
thermodynamic system. We start our analysis with a multi-component element of volume
that can interchange internal energy, matter, and a collection of work coordinates [4]
whose densities are z
j
. This thermodynamic element of volume veries a generalized Gibbs
equation [4, 7]
Tds = du

j
dz
j

k
dn
k
, (24)
for thermodynamic temperature T, density of entropy s, density of internal energy u, and
chemical potential
k
and mole unit per unit volume n
k
of component k. The coecients

j
, conjugate to the densities z
j
, are work coecients [4]. Examples of work coecients
and work coordinates are given in the literature for specic thermodynamic interactions, such
as those due to transfer of density of charge q across a potential dierence , change of
density of electric dipole moment p in presence of an electric eld E, and change of density
of magnetic dipole moment m in presence of a magnetic eld B [1, 4]

j
dz
j
= dq E dp B dm+ . (25)
The balance equation for entropy has the same form as (3), but now with a generalized density
of production of entropy

S
=

j

j

z
j

k

k

N
k
T
+J
U

_
1
T
_

j
J
Z
j

_

j
T
_

k
J
N
k

_

k
T
_
(26)
and with a generalized entropy ux
J
S
=
J
U

j

j
J
Z
j

k

k
J
N
k
T
. (27)
In the above expressions, J
U
, J
Z
j
, and J
N
k
are the ows of internal energy, work coordinate
j , and component k, respectively; whereas
U
,
z
j
, and
N
k
are their corresponding densities
of production. For instance, the density of production of component k by chemical reactions
can be expressed in terms of the reaction velocities v

and the corresponding stoichiometric


coecients
k
as
N
k

k
.
Our starting point is (27). Once again, the new denition of heat for general systems can be
obtained by considering thermal properties in the chemical potential. The generalized Euler
equation associated to (24) is
Ts = u + p

j
z
j

k
n
k
. (28)
Notice that we are working in a thermodynamic space (u, {z
j
}, {n
k
}) of natural variables,
albeit more general.
We nd now a diculty which was absent in the one-component system, the chemical poten-
tials are interrelated and we cannot isolate the thermal properties of the chemical potential
5 HEAT FOR GENERAL THERMODYNAMIC SYSTEMS 10
of component k from the generalized Euler equation. We can solve this diculty if we use
(28) to obtain the next identity
1
R
_
Ts u p +

j
z
j
_
=
k
n
k
, (29)
with R

k
the total number of components. This identity will be our central element
for the denition of heat ux for general systems. Using (29), the density of production of
entropy (26) can be rewritten as

S
=

j

j

z
j

k

k

N
k
T
+J
Q

_
1
T
_

j
J
Z
j

_

j
T
_

k
J
N
k

_

k
T
_
(30)
and the entropy ux (27) rewritten as
J
S
=
J
Q
T
+

k
_
s
Rn
k
_
J
N
k
, (31)
with
k
(1/Rn
k
)p

j
(z
j
/Rn
k
)
j
and a generalized heat ux
J
Q
J
U

j
J
Z
j

k
_

k
+
Ts
Rn
k
_
J
N
k
. (32)
This extension to general systems of the new denition of heat ux continues satisfying all
the physical requirements (i) (iv) presented in the section 2.
Notice that the factor (s/Rn
k
) has units of molar entropy, which implies that (
k
+Ts/Rn
k
)
has units of molar energy. The new heat ux J
Q
for general systems continue considering the
extensive thermodynamic character of both the entropy S and the internal energy U. This
allows a direct extension of the analysis for simple systems presented in the section 4 to
multi-component generalized thermodynamic systems. Eectively, (31) is telling us that J
Q
is proportional to a thermal entropy ow J

S
where the entropy transfered through the mass
ows has been subtracted
J
Q
= TJ

S
T
_
J
S

k
_
s
Rn
k
_
J
N
k
_
. (33)
The corresponding generalizations of the DeDonder entropic term (22) and of the dU
matter
in (23) are easy to obtain.
Integrating the denition (32) of the new heat ux J
Q
over the area A of an initially homoge-
neous system, multiplying by the time interval dt needed to achieve a nal equilibrium state
from the initial equilibrium state, downgrading from modern to classical thermodynamic
space [11], and considering the relation between the total energy and the internal energy, we
obtain a generalization of the classical Bryan & Born denition of heat (2) to open systems
Q dE W

k
_
TS
RN
k
_

e
N
k
. (34)
Evidently (34) reduces to (2) for closed systems. Using (31), we can also obtain the proper
generalization of the classical Clausius theorem to open systems
dS
Q
T
+

k
_
S
RN
k
_

e
N
k
. (35)
REFERENCES AND NOTES 11
Summarizing: The new heat ux J
Q
J
U

j

j
J
Z
j

k
(
k
+ Ts/Rn
k
)J
N
k
provides (i)
a complete distinction between open and closed systems, (ii) non-redundancy, (iii) natural
variables for the thermodynamic potentials, and (iv) a sound, complete, and intuitive general-
ization of classical thermodynamic expressions such as Clausius TdS Q and the Bryan
& Born denition (2) to open systems.
Acknowledgements
I thank Christopher Game for calling my attention to Bryans denition of heat and for
his invaluable assistance on the preparation of this manuscript.
References and notes
[1] Modern Thermodynamics 1998: John Wiley & Sons Ltd.; Chichester. Kondepudi, D.
K.; Prigogine, I.
[2] Thermodynamics, an introductory treatise dealing mainly with rst principles and their
direct applications 1907: B. G. Teubner; Leipzig. Bryan, G. H.
[3] IUPAC recommendation for the signs work is considered positive if increases the energy
of the system is used in this article as in most of the literature. See, for instance,
[1, 4, 5, 8, 15].
[4] Survey of fundamental laws 1971: In Physical Chemistry, An Advanced Treatise; Volume
I Thermodynamics; Academic Press, Inc.; London; Wilhelm Jost (Editor). Haase, R.
[5] Thermodynamics and an Introduction to Thermostatistics; Second Edition 1985: John
Wiley & Sons Inc.; New York. Callen, Herbert B.
[6] Callen uses the old term closed instead of the more modern and adequate term
isolated for referring to systems that cannot interchange energy or matter with the
surrounds [1, 4].
[7] This is an excellent approximation for systems where there are not large gradients and/or
fast processes. When the local approximation does not hold, the formalism of extended
thermodynamics is needed. See [15] and references cited therein.
[8] Denition of Heat in Open Systems 1980: Aust. J. Phys., 33, 95105. Smith, D. A.
[9] We use here the term ux for J
Y
because readers must be more familiar with this
old terminology. However, recent ocial recommendation by IUPAP and IUPAC propose
the term ux density for J
Y
[10].
Some references are completely inconsistent regarding the terminology. For instance,
DeGroot & Mazur consider that J
e
is both the energy ux and the energy ux
per unit surface and unit time [13], which makes no sense. Callen initially denes
ux in agreement with recent IUPAP/IUPAC terminology, but latter redenes ux
REFERENCES AND NOTES 12
as the x, y, or z component of a current density [5], ignoring the dierent units.
Kondepudi & Prigogine [1] oer similar inconsistencies.
The recent ux density term by IUPAP and IUPAC is still open to objections. Last
recommendations propose the more adequate and unambiguous term areic ux [10].
[10] Scientic quantities 2012: Knowledge 1111221558v1. Gonzlez lvarez, Juan
Ramn.
[11] Classical thermodynamics is dened in a timeless thermodynamic space associated to
idealized, innitely slow, reversible processes [1]. This forces the use of imperfect
dierentials , for quantities such as heat and work. Modern thermodynamics is dened
in an extended thermodynamic space without such idealizations, which avoids the use
of imperfect dierentials [1].
[12] The DeDonder equation d
e
S = dQ/T, associated with modern thermodynamics, would
not be confused with the Clausius equation dS = Q/T of classical thermodynamics.
In the rst place, the DeDonder equation is for d
e
S, whereas the Clausius equation is
for dS. In general, d
e
S = dS, which implies that the Clausius equation is only valid for
reversible processes, whereas the DeDonder equation is valid for irreversible processes
as well.
In the second place, dQ in the DeDonder equation uses perfect dierentials, whereas
the Clausius equation relies on imperfect dierentials such as Q. See [11] for details.
[13] Non-equilibrium thermodynamics 1984: Courier Dover Publications, Inc.; New York.
DeGroot, Sybren Ruurds; Mazur, Peter.
[14] Non-equilibrium thermodynamics 2001: In Encyclopedia of Chemical Physics and Phys-
ical Chemistry; Volume I Fundamentals; IOP Publishing, Ltd.; Bristol; John H. Moore;
Nicholas D. Spencer (Editors). Fox, Ronald F.
[15] Extended Irreversible Thermodynamics; fourth edition 2010: Springer; New York. Jou,
David; Casas-Vzquez, Jos; Lebon, Georgy.
[16] Statistical dynamics; matter out of equilibrium 1997: Imperial College Press; London.
Balescu, Radu.
[17] Gravitation 1973: W. H. Freeman & Co.; San Francisco. Misner, C.; Thorne, K.;
Wheeler, J.
[18] Misner, Thorne, & Wheeler [17] consider only simple thermodynamic systems and
absence of chemical reactions. They begin by introducing a heat-ux 4-vector q
with components (0, q
1
, q
2
, q
3
) dened in the rest-frame of the element of volume; they
vaguely dene the spatial components as energy per unit time crossing unit surface
see their 22.16.b, without specifying what energy. They dene an entropy 4-vector
as s su +q/T their 22.16.e and postulate the second law of thermodynamics in
the form
4
s 0 their 22.16.g, with
4
the four-divergence.
This is all weird and incorrect. In the rst place, they do not unambiguously dene what
is the energy associated to their heat-ux. In the second place, their 4-vector s cannot
be an entropy because has units of entropy per unit area per unit time; s is, in reality,
an entropy-ux 4-vector [9]. In the third place, the authors mix local and material ows
when dene s as their heat-ux 4-vector plus an convective term. As a consequence, the
REFERENCES AND NOTES 13
internal energy ux 3-vector associated to the entropy ux 3-vector is measured locally,
at a given spacetime point, whereas the heat-ux 3-vector associated to their four-vector
q cannot be measured locally. Compare this with Callen and his use of a local heat-ux
3-vector, which is locally measured [5].
[19] A detailed analysis of their work reveals that 15.4.20 and 15.4.23, dening the heat
ux and the source of heat respectively, contain u
k
instead u
0
k
; both denitions contain
dummy

i
symbols that have to be eliminated. The authors give 15.5.12 in terms of an
electric current I which they dene in the text; however, when restating the density of
production of entropy A15.1.13 in terms of their J
Q
KP they introduce a I
k
which they
nowhere dene we can guess that I
k
represents the contribution of ion k to the electric
current. They use everywhere for the electric eld, except in table 15.1 where
they use . Several expressions in the same chapter contain typos such as /dt. Their
15.5.11 is dimensionally incorrect, but following the text we can guess that J

k
has to
be substituted by J
U
. In section 16.3 they start using I
e
for electric current, but after
nine equations suddenly change to the older notation I, which then rename as current
density. A T is missing in 10.1.15 for electrical conduction. Both the gure 10.3 and
the equation 10.1.11 confound chemical and electrochemical potentials; etcetera.
[20] Molar quantities are dened as Y/N. IUPAC recommends the notation Y
m
for molar
quantities. Notice that Y/N = y/n.

S-ar putea să vă placă și