Você está na página 1de 44

Pensee Journal Vol 76, No.

6;Jun 2014

What Nicolas Leonard Sadi Carnot wanted to tell us in fact


E. B. Starikov 1), 2)

1)
Institute for Theoretical Solid-State Physics, University of Karlsruhe (KIT), Wolfgang-Gaede-Str. 1,
D-76131 Karlsruhe, Germany. E-mail: starikow@tfp.uni-karlsruhe.de)

2)
Department of Chemical and Biological Engineering, Chalmers University of Technology, SE-41296
Gothenburg, Sweden.

Deyr fé, Cattle die,


deyja frændr, Kinsmen die
deyr sjálfr et sama; You yourself die;
ek veit einn, I know one thing,
at aldri deyr: Which never dies:
dómr um dauðan hvern. The judgment of a dead man's life.

Meurent les biens


Meurent les parents
Et toi, tu mourras de même;
Mais je sais une chose
Qui jamais ne meurt:
Le jugement porté sur chaque mort.

DU STIRBST – BESITZ STIRBT


DIE SIPPEN STERBEN.
EINZIG LEBT – WIR WISSEN ES –
DER TOTEN TATENRUHM.

EDDA (76-th Verse of the Hávamál)

Abstract

The logical structure of the conventional “equilibrium” thermodynamics has been carefully analyzed here. An
imperative to attentively and thoughtfully re-read the work by N. L. S. Carnot naturally emerges from such an
analysis. The implications resulting from such re-reading for the whole field of thermodynamics and statistical
physics are discussed in detail.

Keywords: Thermodynamics, Energy, Entropy, Basic Laws

171 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

1. Introduction

After taking the trouble to read through the above abstract and arriving at this point, any attentive reader would
no doubt query, whether the author is ‘stark raving bonkers’ – or he is just in the corresponding initial stage of this
well-known complication …

… Indeed, re-reading Carnot after such colleagues as Clapeyron, Clausius, Lord Kelvin – wouldn’t this really be a
crazy idea? …

But, before applying to the competent medical service, let us first open the book by Hendrick C. van Ness,
entitled “Understanding Thermodynamics” [1]. In his preface, van Ness states that the book at hand is not a standard
textbook in the field – and he justifies an imperative for writing such a non-standard book by hoping to “help the
student over the very difficult ground characteristic of the early stages of an initial course in thermodynamics.” Was
H. C. van Ness, a prominent American chemical engineer, the very first colleague, who could notice such difficulties?
The answer is negative. This looks like a truly old story and even seems to become an endless one … Indeed, a
prominent German engineer, Prof. Dr. Rudolf Plank, “Der Papst von Kältemaschinen” (the Grandpa of chilling
devices) had clearly stated them in his inaugural speech, as he was offered a position of the Karlsruher Technical
University’s Rector in Germany (the KTU of that time is presently known as KIT) [2]. Such difficulties were a field
of thorough studies for a prominent American mechanical engineer at MIT, Joseph Henry Keenan – and his school
[3,4]. In trying to surpass these difficulties, prominent physicists all over the world wrote the books similar to that by
van Ness [5-10]. On the other hand a French engineer Paul Chambadal [11] as well as American engineer Stanley W.
Angrist, together with chemist Loren G. Hepler have noticed (and presented with rigor and humor) the definitely
paradoxical nature of the conventional thermodynamics [12].

… So, well, what are then these difficulties? Where do they come from? … Why the conventional
thermodynamics seems to be paradoxical at all?

The present report attempts to answer the above posers and tries showing the way to overcome them.

2. How many kinds of the conventional thermodynamics are there?

One of the prominent American thermodynamicists, Mark Waldo Zemansky, distinguished among three basic
flavors, trends, or fashions, of thermodynamics [13], namely: its mechanical-engineering, chemical and physical
branches. He published this classification as early as in 1957, but his result seems to be valid up to date …

Furthermore, M. W. Zemansky had pointed out a number of thermodynamic “misnomers”, as he put it, e. g.:
the “enthalpy” (being nothing more than just “heat content”) and the “free energy” (being either Gibbs or Helmholtz
function – or even the both of them at the same time). In connection with this, M. W. Zemansky had even noticed
that “the free energy is not energy but is freely misused” … This standpoint was extensively discussed in the
physical and chemical literature [13], and the conclusion had been drawn that the famous Gibbs function ought to
correspond “rather to entropy than to energy”, whereas the Helmholtz function isn’t entering at all into that

172 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

discussion …

Well, whatever, but M. W. Zemansky clearly stated that “no student can really understand the concept of
entropy when it is introduced in the traditional way by means of the Kelvin-Planck statement of the second law, the
Carnot cycle, and the Clausius theorem”. Here, M. W. Zemansky cited some colleague of him, who pleaded for the
ultimate usefulness of Carathéodory’s mathematical approach in trying to grasp the meaning of entropy notion …
But M. W. Zemansky clearly slighted such an approach, by calling it “dressing a baby boy in a tuxedo” …

With all this in mind, there seems to be a clear imperative to reconsider the very foundations of the
conventional thermodynamics. The present communication is not an attempt to re-construct the thermodynamics
history, for the latter is described and analyzed in detail in the marvelous books by a prominent American engineer
Clifford Ambrose Truesdell [14,15]. Instead, our main point here ought to be a thoughtful re-analysis of the logical
structure of the conventional thermodynamics, to complement the “operational” approach of the prominent
American physicist Percy Williams Bridgman, as concisely formulated by M. W. Zemansky: “The good physicist
has studied everything valuable from thermodynamics”. Herewith we sincerely concur in P. W. Bridgman’s and M.
W. Zemansky’s opinion, as for such a viewpoint. Our humble aim here would therefore be to assist “the good
physicists” in revealing these “valuable” aspects of thermodynamics.

2. 1. Re-reading N. L. S. Carnot’s seminal work

Carnot’s work was translated into different languages, namely: there are three English translations [16-18], one
German [19] and one Russian (in two editions) [20,21]. All the five have been analyzed in detail and commented by
respective professionals in the field.

Moreover, in the English literature there were a lot of attempts to re-read and re-discuss Carnot’s work [22-34].

Hence, it is a good question: Do we really need any re-consideration of this story?

The answer to this seemingly rhetoric question ought nevertheless to be affirmative, to our mind. Why?

1. M. W. Zemansky [13] clearly slighted the attempts to play with the meanings of the French words
“calorique” and “chaleur” used by Carnot. This surely pertains to the works [22-29]. But this is by far not
exhausting the whole original story by Carnot!

2. In effect, most of the standard textbooks on thermodynamics either skip the story, by mentioning Carnot’s
name in connection with his cycle only – or more or less arrogantly slight it. E. g.:

a) In the book by Walter Kauzmann [35] we read, as follows: “…In the year 1824 he [N. L. S. Carnot]
published a monograph ‘Reflections on the Motive Power of Fire’, written in the language of the

173 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

layman and evidently intended as a popular work, in which he pointed out that the production of
mechanical energy in any heat engine results from the transfer of heat from a warm body to a cold one.
(The book was written assuming the caloric theory of heat to be correct). ‘According to this principle’,
Carnot wrote, ‘the production of heat alone is not sufficient to give birth to the impelling power: It is
necessary that there should also be cold; without it the heat would be useless.’ …”. Then Walter
Kauzmann reports in short about the history after Carnot’s book was published. Interestingly, he slights
the well-known re-publication of Carnot’s work by his nephew, Marie François Sadi Carnot, the then
President of French Republic, in 1878 and containing N. L. S. Carnot’s very important hand-written
notes. In effect, the latter aren’t merely showing ‘some doubts’ as for the ‘validity of the caloric theory’.
They do definitely demonstrate N. L. S. Carnot’s firm acceptance of the Energy Conservation Law, as
well as of the atomistic standpoint. SIC! Carnot was thus definitely aware of this Law well before the
famous 1841 publication by Julius Robert von Mayer, so hence, most probably only N. L. S. Carnot’s
sudden departure in 1832 had prohibited him from building up the true conventional thermodynamics…
And the fewest of all the colleagues who really took Carnot’s ideas into a serious detailed scrutiny
were F. Massieu, F. Reech, H. Poincaré, P. Duhem, E. Meyerson [36-41]...

b) In the book by Peter A. Rock [42] even the following deliberation appears: “We shall not outline
Carnot’s original line of reasoning here, because it would be unnecessary complex without the first law.
Rather, we shall briefly analyze the Carnot engine with the same original objective as Carnot …”. And
then the conventional story available in all the standard thermodynamics handbooks follows.

3. There is at least one relatively modern book, namely that by Irving M. Klotz [43], that tries to first of all
analyze the original work by Carnot, and not its graphical and mathematical re-interpretation by Clapeyron.
Indeed, I. M. Klotz notices two general concepts introduced by N. L. S. Carnot himself. These are:

a) “The reversible process, which provides for thermodynamics the corresponding idealization
that ‘frictionless motion’ contributes to mechanics. The idea of ‘reversibility’ has applicability
much beyond ideal heat engines”.

b) “A second conception introduced by Carnot arose from the analogy he drew between the
motive power of a heat engine and of a water wheel. In the heat engine one needs two
temperature levels (a boiler and a condenser, respectively) corresponding to the two levels of
height of a waterfall. In a waterfall, furthermore, the quantity of water discharged by the
wheel at the bottom level is the same as originally entered at the top level. Carnot postulated,

174 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

therefore, a corresponding ‘thermal quantity’, which was conserved as it was carried by the
heat engine from a high temperature to a low one. … … Shortly after the concept of energy
was clarified and generalized into the First Law of Thermodynamics, particularly by
Helmholtz, Mayer, and Joule this ‘thermal quantity’ of Carnot was re-formulated by Clausius
into the concept of ‘entropy’. This conception then provided an alternative basis for the
enunciation of another general principle, the Second Law of Thermodynamics, applicable to
all types of macroscopic natural phenomena instead of just to heat engines”.

As a result, there seems to be an implicit conviction that Carnot’s work itself, without its re-consideration
by Clapeyron, is much too simplistic, for it consists of mere words without any serious math, based upon ancient
spared theories, like ‘Phlogiston’, ‘Calorique’ (as a kind of some “generalized immaterial liquid”, about which a
kind of “myth was dispelled”) etc. … And only owing to the True Genius of Clapeyron, Clausius and Lord Kelvin
could the proper formulation of the Second Law of Thermodynamics have appeared – but, anyway, only after the
pertinent formulation of the First Law of Thermodynamics…

2. 2. The First and the Second Laws of Thermodynamics

First of all, let us note a remarkable logical gap in the deliberations by I. M. Klotz. But this is definitely not
his own failure, for he was doing nothing more than just meticulously expressing the general conviction, more or
less implicitly present in all the thermodynamic books available to us.

Specifically, in the original interpretation of Carnot the ‘thermal quantity’ is conserved, whereas the
entropy by Clausius’ ought to be either equal to zero, or increasing, in accordance with the common knowledge …

Are we pioneering in revealing the logical faults of the conventional classical thermodynamics? Of course,
not! A number of prominent researchers have already expressed the over-all criticism of the classical
thermodynamics’ logical structure [5,44], there were ingenious trials to re-formulate the latter [45], which triggered
really successful attempts to re-build the failing basis [46,47] …

But even with all this in mind, we still have to “bluff” our “ways through the second law” [48], of which
there are for the present hundreds (SIC!) of more or less equivalent definitions [49]…

Rudolf Clausius was definitely aware of this inconsistency himself, so he had continued to look for the
proper solution. And his efforts were boiling down to the two following statements [50]:

1. “The energy of the universe is constant, whereas the entropy of the universe is steadily increasing”.

175 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

2. “The second fundamental theorem, in the form which I have given to it, asserts that all transformations
occurring in nature may take place in a certain direction, which I have assumed as positive, by
themselves, that is, without compensation; but that in the opposite, and consequently negative direction,
they can only take place in such a manner as to be compensated by simultaneous occurring positive
transformations.“

To our sincere regret, Clausius’ investigatory work was in fact practically stopped at this very point, for, as
we know, he had much less time for research after 1870 [51].

As for the first statement, it is a well-known commonplace, which leads to nothing more than a logical
blind alley, although it was constructed by Clausius in order to try underlining the undoubted general significance of
the Second Law of Thermodynamics.

Interestingly, the second statement by Clausius was somehow lost, although it was and remains well
known. Remarkably, it has been recovered most recently [52] in the following context:

“Clausius’ “compensation” is now called thermodynamic coupling. The system including only “positive”
spontaneous process(es) “without compensation” is called uncoupling system, and the system including “negative”
non-spontaneous process(es) “compensated by simultaneous occurring positive” spontaneous process(es) is called
coupling system.”

This reference to some mysterious “coupling” is sounding really solemnly, but remains to be ultimately
helpless, to our mind, for there is no serious analysis of the true physical, chemical, etc., nature of this “coupling”.
Besides, a good question could be posed here: Ought the latter to be connected with something realistic at all?

This is why, in the mean time the standard textbooks on thermodynamics continue seriously discussing the
famous Carnot’s cycle as an ultimate physical reality [53-55], whereas it is nothing more than an ingenious
theoretical widget, devised by N. L. S. Carnot with the sole intention: for nothing more than just to demonstrate
some important idea…

But what is the true Carnot’s idea?

Before answering this important poser, let us first apply to the marvelous book by an American physicist,
Peter Fong [9]. He writes as follows:

“The conventional second law of thermodynamics was introduced by Clausius (1850) and Kelvin (1851)
independently. The concept of entropy was introduced by Clausius (1865), who also summarized the two
thermodynamic laws by saying that the total energy of the universe is a constant and the total entropy of the
universe always increases (SIC! We find here absolutely no mentioning of the second statement by Clausius!). The
historical starting point of thermodynamics was an engineering problem of converting heat into mechanical work.
In this respect Carnot (1824) had actually discovered the essence of the second law decades before it was stated,
but his theory was based on now obsolete caloric theory. It must be borne in mind that before 1850 the caloric
theory was still a reputable theory despite the experiments of Rumford and Davy. The success of Carnot’s theory
and the conflict between the caloric theory and the mechanical theory of heat were reconciled in the 1850’s when

176 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

physicists recognized that there should be two independent laws in thermodynamics – the First and the Second –
not just One, the Carnot principle based on the caloric theory.

When thermodynamics was later applied to a wide range of problems in physics and chemistry, we still
used its basic concept (conversion of heat to work) and its mathematical apparatus (entropy) that originated in
engineering. The engineering elements seem to be irrelevant to the physical and chemical problems of interest.
Thus, we are faced with the situation, which Bridgman described as a river rising higher than its source. Obviously
the only way to avoid the difficulties involved in forcing a river to rise higher than its source is to put the source of
the river at the highest point; then the river flows all by itself. This has never been done. In the latter part of the
nineteenth century, many attempts were made to explain the second law in terms of mechanics; none were
successful. Finally, the thermodynamic laws found a statistical interpretation in the microscopic theory by
Boltzmann, which gave rise to a new branch of physics – statistical thermodynamics.

Although the physical problem is thus solved in microscopic theory, the logical structure of classical
thermodynamics as a macroscopic theory remains in an unsatisfactory state. It has been neglected probably in
the belief that no new physical discovery of fundamental importance is to be expected. Carathéodory (1909, 1925)
was the first to try to establish thermodynamics without the engineering concepts, but he stopped halfway.
According to the previous discussion, thermodynamics must be formulated as a theory of equilibrium to begin
with; therefore we start our discussion of the second law by establishing the law of spontaneous processes.

It is natural that the development of science is occasioned by accidents, and scientific breakthroughs are
initiated more by insight than by reasoning. The historical order is usually not the logical order. However, once the
whole picture is in sight, there is no excuse for not having the logical order straightened out.”

We have intentionally presented here the extended detailed citation from P. Fong’s book, for this is the only
explicit formulation of the modern Creed we are aware of, the Creed which doesn’t seem amenable to any kind of
discussion...

Still, we have underlined three points of interest in this citation, which are susceptible for debate, namely:

1. The caloric theory respectable at Carnot’s days is obsolete nowadays.

First of all, let us recall what is in effect this “infamous” caloric theory. It is long known that the ‘Phlogiston’
theory, the conceptual basis of the caloric theory, had and still has a philosophic significance, for at the times it was
“still respectable” they were believing that there were two fluids in effect, namely: “calorique” and “frigorique”, the
one accounting for heat and other one – for cold [56]. This is exactly what was the important starting point of the
idea by N. L. S. Carnot.

The Phlogiston theory cannot be considered absolutely erroneous even nowadays [57-62], so what is then the
valid reason for considering it “obsolete”? Furthermore, this “calorique”-“frigorique” duality closely resembles the
Yin-Yang duality of the ancient Chinese philosophy, cf. ‘Tao Te Ching’ by Lao Tzu, which presents this holistic,
dialectic view of reality. Specifically, Yin-Yan represents the duality of all the existing phenomena – whether hot and
cold, summer and winter, male and female, or – life and death etc. – as opposing manifestations of one and the same

177 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

principle – and not to be viewed as independent. Such an appreciation of Oneness is now widely and well known to be
central to understanding Lao Tzu's poetry and is fundamental to his philosophy. The difficulty for the European and
American researchers at N. L. S. Carnot’s time – and up to the beginning of the XX-th century – was an ultimate
absence of detailed, clear and legible analyses of the Chinese philosophy for non-specialists. Such accounts in
European languages started to appear only in 20-ies-30-ies of XX-th century, just as the “quantum revolution” in
physics took place. Thus, the minds of the leading researchers in natural sciences of those times weren’t in the first line
busy with thermodynamics’ problems. This is why the “operational” standpoint by P. W. Bridgman was – and still
remains – the prevailing way of thinking: even, if there are some logical inconsistencies, we are still capable of using
these concepts to get plausible answers to our posers…

So, what is the problem? …Still, the logical inconsistency surely remains, and it is throughout possible – and
desirable – to re-furbish the old stuff, in aiming at getting better paradigms…

2. The First and the Second Laws of Thermodynamics are two independent laws.

Are they really so? It is possible to show – using the well-recognized formulas of Clausius and Gibbs – and
employing the elementary straightforward math – that the Clausius’ ‘compensation’ ought to be just the
entropy-energy/enthalpy compensation, in fact [63]. In following this train of thoughts, we might conclude that the
First and the Second Laws of Thermodynamics are thus not independent from each other, but they are connected to
each other with a kind of dialectic, Yin-Yang-kind of relationship: Whereas the energy is really conserved in total, its
functional part, capable of promoting some useful work, is consumed. While consuming this useful energy portion,
the progress of the process in question encounters omnipresent and ubiquitous hindrances/hurdles. The particular
case of this rule was already formulated by Isaac Newton – as his Third Law: “For every action, there is an equal
and opposite reaction”. To sum up, there is no reaction without action – and vice versa.

Thus, what N. L. S. Carnot would in fact like to tell us, could most probably sound as follows. Even when
designing an ideal, perfectly working mechanism (in Carnot’s case – the ideal heat engine) with the circular mode of
action, where the starting point is always equal to the final point, the resulting widget must always have some
maximum efficiency (Carnot efficiency), which is anyway less than 100%. If so, then why the latter situation must
always take place? Because the forces promoting the useful work to be done should ALWAYS encounter some
definite hindrances, that ANYWAY thwart the actual promotion. Then, the equilibrium – discussed by Carnot – and
necessary to formulate the true ‘equilibrium thermodynamics’ – ought to be the mutual compensation of the forces
promoting the work in question and those thwarting it. The latter forces could then be designed as entropic ones.

With this in mind, the popular commonplace ‘entropy production’ has nothing to do with the actual driving
force(s) and ought to be revised. The true driving forces should come from the proper energy conversion – the useful,
productive energy might then be produced solely this way. Hence, the last, but not the least – the useful energy in
question is anyway only some definite part of the total energy. The mechanical analogue would be the potential
energy, which is convertible into the kinetic energy (or Vis Viva – the ‘living force’ – using the notation of Gottfried
Leibnitz for the latter). Further, the conversion of the potential energy into the Vis Viva is never free from ubiquitous
hindrances. Therefore, the equilibrium in the process thus running would be reached, when the potential energy is

178 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

completely spent – thus reaches its minimum – through creating the Vis-Viva-type of the process promotion – and
simultaneously overcoming the available hindrances. Now, the picture of any realistic process ought to be complete,
of whatever nature the process involved might be in effect.

Interestingly, the above-sketched representation of the true equilibrium thermodynamics has been worked out
and formulated by Dr. George Augustus Linhart. Meanwhile, although he could manage such a terrific research
achievement, his name remained completely unknown until most recently [64-66].

In returning to the theory by N. L. S. Carnot, we ought to conclude that he was completely right. In effect,
there ought to be only One Basic Law of thermodynamics, which is but similar to the Janus Bifrons mascot of the
ancient Roman mythology – or to the Yin-Yang mascot of the ancient Far-Eastern mythology. The two faces of
Janus, or the Yin and Yang, in the conventional thermodynamics are then – a) the livening force, Vis Viva (in the
modern notation: the useful energy), the source of the productive force causing and promoting any progress – as
well as b) the entropy, or, in other words, all the hindrances/hurdles thwarting the progress, respectively. It is also
possible to put this as follows: the useful energy and the entropy are related to each other as the action and reaction,
respectively, in Newton’s mechanics.

3. Was the conventional statistical thermodynamics the ultimate solution of the logical problems in the
classical macroscopic thermodynamics?

Our answer would be – not completely. Why? The pertinent development of the statistical thermodynamics wasn’t
properly finalized – due to the sudden, but inevitable, departures of James Clerk Maxwell, Josiah Willard Gibbs and
then Ludwig Boltzmann. First of all, neither Ludwig Boltzmann, nor Max Planck – not to speak of their official
followers – could have formally proven the famous formula for the entropy vs. probability (here, only Bernard H.
Lavenda’s detailed considerations on the subject represent a fortunate exception from such an observation [67]),
although the following expression is intuitively correct and well known to be practically true:

S = kln( W). (1)

Still, Dr. G. A. Linhart was, to the best of our knowledge, the very first colleague, who could have formally solved
this problem using intrinsically Bayesian statistical approach – anyway without explicitly applying to the
conventional atomistic representation [64-66]. However, Linhart’s work is “widely unknown”, although it is (at least)
partially published in well-known scientific journals … Thus, the important relationship between entropy and
probability still remains by and large tentative.

The main disadvantage of such a long-standing reputedly tentative nature of the equation S = k ln(W) is its
misdirecting role of serving as a “scientific foundation” of the “fact” that the entropy is in effect the “degree of
disorder” (for the most recent literature on the theme, see, e. g., [68]).

179 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Besides, it is important to note that the very possibility to avoid the conventional explicit microscopic
consideration could be rather helpful, for notions like “extremely large number of atoms/molecules” are in fact
FUZZY, in accordance with the well-known SORITES paradox. Indeed, it’s a practically difficult task to achieve a
universally strict definition of what exactly the “large number” is and hence, operating with such notions and their
derivatives as “thermodynamic limit” isn’t really a productive approach. However, we would prefer to dwell on this
very important conceptual topic elsewhere.

2. 3. Reversibility, equilibrium… What ought to be the true “equilibrium thermodynamics”?

Now, let us revert to the idea of “reversible processes”, of “reversibility” which ought to “have applicability
much beyond the ideal heat engines” [43]. On the one hand, any honestly written textbook of thermodynamics
explicitly states, that in reality there are no processes, which could be perfectly to 100% reversible. But on the other
hand, most of these textbooks consider in detail the N. L. S. Carnot’s widget as an ultimate physical reality, for there
seem to be no other way to introduce the notion of entropy … But why is it just so, as it is?

The point is that the most of colleagues in the field(s) retained – and still retain – the
Clapeyron/Clausius/Kelvin way of treating the ingenious work by N. L. S. Carnot. Moreover, the Carnot’s ingenious
widget seemed (and, interestingly, seems as yet!) to be the only possible way of introducing the important notion of
entropy.

Meanwhile, several serious authors, Hermann von Helmholtz and Max Planck among them [69-72], have
already noticed that all the conventional thermodynamic notions, like entropy and absolute temperature might well
be introduced without cyclic processes, ideal gas and other purely logical-mathematical – but expressly non-realistic
– auxiliaries. This train of thoughts was triggered in France – by such authors, as Reech [36], Massieu [37,38],
Poincaré [39] and Duhem [40] – by Hermann von Helmholtz [69,70] and Max Planck [72] in Germany – and along
with all the latter colleagues – by Josiah Willard Gibbs in the USA [73].

But before getting to the works of the named prominent colleagues, we would like to dwell on the
ambiguities/obscurities of the conventional “equilibrium” thermodynamics. In a particularly blatant form it appears
in the treatises on kinetics. Let us take a marvelous recent book on the subject by R. W. Balluffi, S. M. Allen and W.
C. Carter [74].

The authors start with the assertion that thermodynamics and kinetics should be clearly distinguishable.
They point out as follows:

“Thermodynamics is the study of equilibrium states in which state variables of a system do not change with
time, and kinetics is the study of the rates at which systems that are out of equilibrium change under the influence of
various forces. The presence of the word dynamics in the term thermodynamics is therefore misleading but is
retained for historical reasons.”

180 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

More details about this “Thermodynamics – Thermostatics” paradox could be borrowed from the book by a
prominent American thermodynamicist, Myron Tribus [75]. But what is then this notorious “equilibrium
thermodynamics”? The following answer comes from the book [74]:

“Thermodynamics grew out of studies of systems that exchange energy. Joule and Kelvin established the
relationship between work and the flow of heat, which resulted in a statement of the first law of thermodynamics. In
the Clausius treatise, ‘The Mechanical Theory of Heat’, the law of energy conservation was supplemented with a
second law that defined entropy, a function that can only increase, as long as an isolated system approaches
equilibrium [50]. Poincaré coined the term ’thermodynamiques’ to refer to the new insights that developed from the
first and second laws. Development of thermodynamics in the nineteenth century was devoted to practical
considerations of work, energy supply, and efficiency of engines. At the end of the nineteenth century, J. Willard
Gibbs transformed thermodynamics into the subject of phase stability, chemical equilibrium, and graphical
constructions for analyzing equilibrium that is familiar to students of materials science. Gibbs used the first and
second laws rigorously, but focused on the medium that stores energy during a work cycle. From Gibbs's careful and
rigorous derivations of equilibrium conditions of matter, the modern subjects of chemical and material
thermodynamics were born [73]. Modern theories of statistical and continuum thermodynamics which comprise the
fundamental tools for the science of materials and processes derive from Gibbs's definitive works.

Thermodynamics is precise, but is strictly applicable to phenomena that are unachievable in finite
systems in finite amounts of time. It provides concise descriptions of systems at equilibrium by specifying constant
values for a small number of intensive parameters.” (our underlining)

Therefore, this unintelligible thermodynamics/thermostatics looks like just a sheer misunderstanding, as


compared with the kinetics … An immediate question should then be posed: Well, but how can we profit from the
conventional thermodynamics, if it really seems to be inapplicable to any realistic process??? Nonetheless, further
reading of the book [74] still provides us with some insight into the relationship between thermodynamics and
kinetics. In their explanations the authors deal with the properties of the Gibbs’ potential function, which helps
define the conditions for thermodynamic equilibrium. Then, at some point the authors [74] lament:

“A quandary arises: general statements regarding the approach to equilibrium that are based on
thermodynamic functions necessarily involve extrapolation away from equilibrium conditions. However, useful
models and theories can be developed from approximate expressions for functions having minima that coincide with
the equilibrium thermodynamic quantities and from assumptions of local equilibrium states.”

Well, but the recent works by Jarzynski and Crooks definitely show that there is absolutely no such problem
with the free energy differences [76-78]. Hence, first of all, as the authors [74] conclude – however, without even
mentioning the results by Jarzynski and Crooks:

“This approach is consistent with the laws of thermodynamics and provides an insightful and organized
theoretical foundation for kinetic theories.”

181 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Moreover, here it is also appropriate to recollect the paper by M. W. Zemansky we have already cited above
[13], where he exclaimed:

“…To engineers, the heat engine and the refrigerator are the main excuses for the entire subject of
thermodynamics. Chemists, on the other hand, know little about P-V diagrams, engine cycles, or steady flow
equations and give the impression that the only process of importance in the universe is the chemical reaction. They
love to refer to the free energy without making clear whether they mean the Helmholtz function or the Gibbs function
and without stating what it is a function of, and how many variables are independent. With so much left unsaid and
with unseemly haste, the chemist then proceeds to render the function a minimum for no discernible reason, and,
believe it or not, he always gets the correct answer! This is a mystery compared with which nuclear forces are
child’s play.”

For the present, it is throughout possible to productively comment on M. W. Zemansky’s exclamations.


First of all, nobody apart from engineers could find how to correctly use the classical thermodynamics – to achieve
definite successes (heat engines, refrigerators etc.) – the break-through in this important direction has definitely
started with the works by Joseph Henry Keenan, his numerous predecessors and his school [3,4]. They have enabled
the introduction and successful usage of the exergy notion (this stands for the available, useful energy) and clearly
shown: a) that the latter is intimately connected to the mysterious “free energy/free energies” and b) how to
productively employ such a notion – to render the classical thermodynamic analysis a true science coping with
diverse research and development problems.

That chemists could somehow reach truly correct answers using their – seemingly ‘weird’ – train of
thoughts is by far not so mysterious, as soon as we recall the recent work by Jarzynski and Crooks [76-78].

Moreover, it is possible to show that there is an intrinsic connection between the Helmholtz and Gibbs
functions. The latter interrelationship clearly points out the not less intrinsic interconnection between the First and
the Second Basic Laws of thermodynamics [63].

Remarkably, the chemists together with physicists are definitely aware of the intrinsic difficulties with the
so-called “equilibrium” thermodynamics and intensively try to find the ways out [79-83]. In fact, the activities in
connection with working out the true concept of thermodynamic equilibrium isn’t a recent novelty – they had started
already with the work by Lord Kelvin [84], it entered even a number of late XIX-century’s, as well as
mid-XX-century’s textbooks [85-88] – and continued by other notable chemists, like Leopold Pfaundler in Austria,
August Horstmann in Germany and George Downing Liveing in Great Britain – but was somehow lost, although not
completely forgotten. For the all the interesting details of this instructive story, cf. the most recent marvelous review
by William B. Jensen [89].

Remarkably, the work by W. B. Jensen underlines one of the most significant and important points within
the “chemical” thermodynamics. This pertains to revealing the genuine roots of the sincere attempts by numerous
colleagues to rationalize the notion of entropy. The story starts practically immediately after the introduction of the
latter term by Rudolf Clausius. Indeed, one of the prominent Scottish physicists of that time, Peter Guthrie Tait,

182 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

attempted to re-conceptualize the entropy notion, by ascribing it the physical sense of “available, usable, useful,
lively energy” and not “dissipated, lost, dead energy”. Interestingly, P. G. Tait’s train of thoughts found really
prominent proponents like, e.g., James Clerk Maxwell (although J. C. Maxwell had later on revoked his support to
this idea of P. G. Tait) [89,90].

W. B. Jensen notes equitably:

“The entire hiatus prompted by Tait’s ill advised attempt to redefine entropy can be summarized by the
fundamental relationship G = -T*St, where G is the Gibbs’ free energy of the reaction system, S t is the total entropy
of both the reaction system and its surroundings, and T is the absolute temperature. Essentially, G is identical to
what Tait meant by “Availability” and, like it, tends to a minimum as one approaches equilibrium. This was the term
Tait wished to rename ‘entropy’ – a proposal that entailed not only a change in the meaning and sign of S t – as
originally defined by Clausius – but also, taking T into account, a change in its fundamental physical dimensions as
well. T*S t on the other hand when taken in the Clausius’ original sense, is a good measure of the total energy
dissipation of the system at the temperature in question and tends to a maximum as one approaches equilibrium.”

To sum up, we might add here nothing more than just adhere to W. B. Jensen’s completely justified
conclusion as for the true physical sense of the entropy notion: “… Neither information nor disorder are true causal
agents like energy …”. And there is one more apparently erroneous, but stubbornly vivid, causal agent for entropy,
namely the so-called “arrow of time” [48] … Well, sure, the only true logical, physical, chemical etc. connection for
entropy can be nothing more than energy. But there ought to be definitely a couple of topics more to the story.

1. If we speak of ‘reaching thermodynamic equilibrium’, then what exactly is being equilibrated during
the processes in question? Our answer would be – the equilibration takes place between several types of energy,
while the total energy remain constant – this is just the Law of the Energy Conservation (The First Basic Law of
Thermodynamics). What are these types? From the classical mechanics we know that basically there is potential
energy and kinetic energy. The both could be converted into each other, so that the former could be seen as a kind of
energy inventory, whereas the latter stands for the “Vis Viva”, “the livening force”, that is, something which is
capable of livening the processes up. Therefore, the movement of whatever kind (be it mechanical or chemical – or
even other possible kinds) starts as soon as the energy inventory starts to be “livened up”. This would be a perfect
situation, if nothing more than the mentioned livening-up would happen, for the progress of the movement would be
infinite, all the processes based upon such movements would be efficient to 100%. But exactly here comes the
famous “Second Basic Law of Thermodynamics” and introduces all the omnipresent, all the ubiquitous constraints,
hindrances and faults which render any process finite – and hence – any objective, any aim reachable in principle!
Such negative factors could be united together under the roof of the Entropy Notion. Thus, the available “Vis Viva”
could be thought of as being exhausted by the available sum of the resisting (entropic) factors – and this is why, the
energy inventory can never be used with the outmost efficiency. The best (ever possible) efficiency estimate comes
from the ingenious widget introduced by N. L. S. Carnot.

To sum up, there is an eternal mutual compensation of the available energy by the ubiquitous omnipresent
entropy (constraints, hindrances). The available energy inventories are used to trigger diverse processes all over the

183 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

world. Triggering processes means livening up the potential energy to render it kinetic energy, by means of energy
conversion, whereas the total energy remains to be constant, according to the First Basic Law of Thermodynamics.
Before the process has ever been triggered, there is absolutely no way for constraints and hindrances. As soon as the
triggering takes place, the energy inventory starts to be exhausted, in that the latter is being livened up by producing
the kinetic energy necessary to promote the actual process, to promote the progress of the latter. As soon as the onset
of the progress begins, the omnipresent ubiquitous hindrances and constraints start to increase, with the aim to stop
the developing progress. This is exactly where the Second Basic Law of Thermodynamics starts to be active, for the
over-all increase in all the possible hindrances and constraints could in general be represented by the increase in
entropy. As Rudolf Clausius had ingeniously shown, if we employ now the ideal cyclic widget model of N. L. S.
Carnot, then we get from the original, initial state to the final state ideally identical to the original one. This is why
the total change in entropy will be identically equal to ZERO. Otherwise, at any other point of the process
development the entropy will be GREATER than that in the corresponding initial/final point. This is just the only
possible physical sense of the notorious “entropy increase”…

2. The last but not the least: We seem to have finally arrived at the clear picture of the true conventional
equilibrium thermodynamics – in full accordance to what N. L. S. Carnot and all of his prominent – and not so
prominent – followers would like to communicate us. Now, the natural question arises: What is then the actual role
of Carnot’s widget? Does this successful model have any (implicit or explicit, whatever) additional physical sense?

The answer is definitely negative. This has already been recognized long time ago by such reputable
researchers as H. von Helmholtz, J. W. Gibbs, M. Planck … The next question would then be: Why then are we still
so firmly convinced that there is no other way to introduce the entropy notion aside from the Carnot’s widget? The
answer is definitely difficult and would require a separate paper or, may perhaps, even a separate book… But
nonetheless, there were at least two colleagues, who took their time to describe the notorious introduction of the
entropy and absolute temperature without applying to Carnot’s widget. The first one was a prominent German
engineer, Prof. Dr. Ernst Schmidt (1892 – 1975), who devoted a separate chapter in his thermodynamics textbook to
the entropy without Carnot’s cycle – and made just a volatile reference to a “suggestion of Max Planck” [91]. The
second one was a prominent American engineer, Prof. Dr. Joseph Kestin (1913 – 1993), who translated the German
text by E. Schmidt into English (with the assistance of Schmidt) and published it [92].

Interestingly, Prof. Kestin has published/edited several books devoted to thermodynamics as a whole, as
well as to the Basic Second Law [93-95]. But neither he himself, nor any reader of the books by Schmidt and Kestin
– nobody – has (at the very least!) noticed the above remarkable point (not to speak of discussing it in detail!)… The
objective reason for such a flaw might in principle be the fact, that Ernst Schmidt had in 1933 signed the notorious
“Bekenntnis der Professoren an den deutschen Universitäten und Hochschulen zu Adolf Hitler” … Even if this
ought to be the correct answer … Dear Colleagues! Errare Humanum Est!

3. The interrelationship between the conventional thermodynamics, system’s size and time. That the
conventional thermodynamics ought to be “thermostatics” in effect is nothing more than another “hiatus”. The
further “hiatus” is “the inapplicability of the conventional thermodynamics to describe the systems of finite size”.

184 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Why?

The answer to the latter problem is clearly given by the theories of Jarzynski, Crooks and other colleagues
[76-78,96,97], which are known to be properly applicable to all the known systems, up to the single-molecular ones
[98]. Sure, with all this in mind, one has to be really careful when operating with the terms like “thermodynamic
limit”, for strictly speaking the latter cannot be universally defined, due to the fuzziness of the notion “big number of
particles” or, in other words – owing to the well known SORITES paradox.

The root of all these difficulties seems to be a drastic over-interpretation of the N. L. S. Carnot’s widget.
And there have already been several notable attempts to overcome this difficulty. Consequently, we have to consider
two more important points on this theme.

3. Non-equilibrium thermodynamics

First of all, the seminal theory by Lars Onsager [99, 100] must be considered, for he may equitably be
called the Father of the Non-Equilibrium Thermodynamics.

Onsager’s theory was a huge leap forward, with respect to conventional classical thermodynamics, in that
he considered the real – irreversible – processes, which are “away from equilibrium”, as he described them, but we
might anyway refine this designation to put it as “away from N. L. S. Carnot’s widget”.

The real processes are indeed real, for they are connected with diverse changes of the corresponding
system’s state(s) with time – and this is just how they are observable in effect. Meanwhile, the conventional
thermodynamics is firmly based upon the fact: the time as an explicit variable is eliminated in the N. L. S. Carnot’s
widget.

The good question: Well, if the time is (at least) implicit in thermodynamics, then how it could be possible
to consistently interpret the ‘physics’ of the N. L. S. Carnot’s widget, which is permanently in equilibrium? Peter
Landsberg showed in his book on thermodynamics [101] the perfectly logical pathway to cope with such a
conceptually blind alley, namely:

“… Thermodynamics is among the most abstract branches of physics. There are two main reasons for this.
First, the theory contains results, which are within wide limits independent of particular systems. … Secondly, in so
far as equilibrium states are concerned, space and time enter only in a rudimentary manner. Space coordinates
occur merely implicitly to describe boundaries between parts of systems, which are often homogeneous. Time
ordering enters through the principle of the increase of entropy, but, for the greater part, it is not involved at all.
Thermodynamics is in fact a study of general relations, which refer to an abstract multidimensional phase space,
and, in so far as the time coordinate is absent, nothing happens in thermodynamics. The quasi-static process, for
example, is strictly speaking a curve in this phase space, the term process being merely a reminder that actual
physical process can often be made to follow such a curve to a high degree of approximation. …”
This is a perfect description of the situation at the time, when Lars Onsager was working on his seminal

185 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

theory – as well as of our actual situation nowadays. Indeed, the N. L. S. Carnot’s widget ought to be taken for
nothing more than a kind of approximation for some realistic cyclic/reversible, etc., process (that is, a kind of
linear/non-linear vibration) developing in time and space and described by a stable limit cycle. And the actual
physical stability of the latter is ensured by the conventional equilibrium thermodynamics (still, the question as to
“what are then the physical agents equilibrating each other and thus ensuring the stable equilibrium?” remains
largely unanswered)…

Moreover, according to the general theory of dynamical processes (mathematically – the theory of
non-linear differential equations) apart from the limit cycles, there might be other types of attractors in the phase
space, the stability of which should anyway be proven: These are namely fixed points, strange attractors … In
following this train of thoughts, the question immediately arises about the true generality of the conventional
thermodynamics based upon the N. L. S. Carnot’s widget. From the standpoint of dynamical systems’ theory in the
phase space the Carnot’s widget ought to represent just a very crude and rather restricted approximation …

Well then, but where the irreversibility described by the entropy ought to come from? It may perhaps be
possessed of some cosmological backgrounds, so, this ensures then the generality of the whole story isn’t it? Here it
should be appropriate to recall the famous utterance, the throughout debatable philosophical maxim, by Rudolf
Clausius: “The energy of the universe always remains constant, the entropy of the universe is always increasing” …

And Peter Landsberg clarified the situation as follows [102]:

“… THE COSMOLOGICAL ARROW OF TIME DOES NOT DOMINATE

It has been suggested that thermodynamic irreversibility is due to cosmological expansion. It is said that it
causes the darkness of the night sky and the red shift of the spectral lines and it is therefore a real phenomenon,
which can cause other effects such as irreversibility. In a sense cosmology contains all subjects because it is the
story of everything, including biology, psychology, and human history. In that single sense it can be said to contain
an explanation also of time's arrow. But this is not, what is meant by those, who advocate the cosmological
explanation of irreversibility. They imply that in some way the time arrow of cosmology imposes its sense on the
thermodynamic arrow. I wish to disagree with this view. …”

Therefore, the actual irreversibility ought to be driven by physically definite forces originating from here
upon Earth, causing/promoting the pertinent observable dynamical processes/flows. So, what are these driving
forces? Lars Onsager, before deriving the general form of his famous reciprocal relations, gives a clear and definite
answer [99]:

“Among the relations to be derived many have been proposed before, but some will be new. An important
group among these relations can be summarized in a variation-principle, which is nothing but an extension of Lord
Rayleigh’s “principle of the least dissipation of energy”; we shall retain the name for the extended principle.
According to this theorem the rate of increase of the entropy plays the role of a potential.”

186 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

In other words, the driving force for the irreversible processes under study ought to be of entropic nature.
The latter idea seems to have been formulated by Lars Onsager himself and for the first time ever – a careful
analysis of all the works cited in his papers [99,100] supports drawing such a conclusion.

To our regret, Lars Onsager is not describing in detail himself, how he could come to exactly such an idea
of all others possible. He refers us to the work by Hon. John William Strutt, Baron Rayleigh [103-105]. In his
preliminary report [103] Baron Rayleigh clearly states that his theory is about rather small deviations from some
mechanical equilibrium and uses the results of William Thomson (Lord Kelvin) and Peter Guthrie Tait on the
“slightly disturbed equilibrium” [106], by writing the relevant mathematical expressions for the kinetic and potential
energies – and thus assuming nothing more than the conservative mechanical systems. But in his famous book on
the theory of sound [104,105], Baron Rayleigh goes a huge step forward: He introduces the dissipation function, F,
aside from the kinetic energy, T, and the potential energy, V, with the function F physically corresponding to
“another group of forces, whose existence it is often advantageous to recognize specially, namely those arising from
friction or viscosity. ...

... But although in an important class of cases the effects of viscosity are represented by the function F, the
question remains open, whether such a method of representation is applicable in all cases. I think it probable that it
is so; but it is evident that we cannot expect to prove any general property of viscous forces in the absence of a strict
definition, which will enable us to determine with certainty, what forces are viscous and what are not. In some cases
considerations of symmetry are sufficient to shew that the retarding forces may be represented as derived from a
dissipation function. At any rate, whenever the retarding forces are proportional to the absolute or relative velocities
of the parts of the system, we shall have equations of motion on the form ...

... We may now introduce the condition, that the motion takes place in the immediate neighborhood of a
configuration of thoroughly stable equilibrium; T and F are then homogeneous quadratic functions of the velocities
with coefficients, which are to be treated as constant, and V is a similar function of the coordinates themselves,
provided that (as we suppose to be the case) the origin of each coordinate is taken to correspond with the
configuration of equilibrium. ...”

And this is exactly where the development of Lars Onsager’s theory starts. But what were the reasons led
Lars Onsager to start developing his theory? He reported all the reasons in question to a conference of the
Scandinavian Natural Scientists’ Society in 1929, in his lecture entitled: “Simultaneous Irreversible Processes”. The
Norwegian-to-English translation of the introduction to that lecture was published relatively recently [107]. In effect,
Lars Onsager has clearly stated, that:

a) several authors (including such prominent authors as Lord Kelvin and Hermann von Helmholtz) have
derived certain relations among two or more irreversible transport processes taking place simultaneously (e. g., heat
conduction, electrical conduction and diffusion) using apparently thermodynamic reasoning. A number of other
authors have used somewhat different approaches (W. Nernst, E. D. Eastman);

187 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

b) common for all the earlier theories is a consideration of a cyclic process that cannot be carried out
reversibly. It is therefore impossible to base such theories on the two laws of thermodynamics. On this ground
Ludwig Boltzmann has pointed out that by purely thermodynamic reasoning one cannot deduce more than certain
inequalities;

c) the proper rigorous derivation of the reciprocal relations could be possible by adopting the statistical
standpoint like that developed in Einstein’s theory of fluctuations. And it is necessary to make only one additional
assumption, viz. the past and the future are on the same footing, in the sense, that any dynamically possible
trajectory might also be traversed in the opposite direction.

Indeed, with all this in mind, Lars Onsager took the results by Lord Rayleigh and boldly threw away the
very potential energy by substituting it with the “rate of entropy increase”. ... To the best of our knowledge, the only
detailed analysis of Lars Onsager’s logical train of thought had been presented by the prominent German theoretical
physicist Richard Becker. In particular, he wrote [108]:

“Classical thermodynamics deals only with reversible changes. The meaning of this restriction has been
demonstrated in Sect. 6 by the example of a simple Carnot engine. Reversible processes have to be performed
"infinitely slowly".

Any real process occurs with a finite velocity and, therefore, is necessarily irreversible. For instance, an
exchange of heat between two bodies A and B is possible only if A is warmer than B, or a piston between two gas
containers moves only if the pressure in the two containers differs. In both cases the actual process is associated
with an increase of entropy. It is a quite strange situation that thermodynamics deals only with reversible processes,
which conserve the entropy of a closed system, whereas the entropy increases in all actual processes.

For the subject dealt with in the following section we have to emphasize a new and characteristic point of
view. We have seen that increase of entropy and irreversibility always occur simultaneously. The statement that
one phenomenon causes the other is, therefore, nothing new, physically. But sometimes such a statement introduces a
more lively and convincing formulation of the basic laws.

In our case, for instance, we can say: Either "the entropy increases because an irreversible process is
taking place" or "the irreversible process occurs because it is associated with an increase of entropy". In the earlier
days the first formulation was more popular. Nowadays the second point of view has been adopted widely. Here the
tendency of entropy to increase is thought to cause the irreversible process.

Sometimes one even speaks of a "force" due to increase of entropy, which pushes the irreversible change.
When assuming such a kind of picture we expect that the irreversible change might be faster, if the associated
increase of entropy is larger. This leads to the suggestion that the velocity of the irreversible process is directly
connected with the corresponding change of entropy.”

188 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Interestingly, such an introduction of some physically unclear, truly magic “entropic driving forces” is in
full accordance with the suggestion by Peter Guthrie Tait (first enthusiastically adopted and then clearly omitted by
James Clerk Maxwell, as we have already discussed). Richard Becker points out as well, that such a train of
thoughts does not introduce any new physical idea, but is nothing more than just facilitating mathematical
description of some realistic events. And as to the “more lively and convincing formulation of the basic laws”... Well,
by now we have several hundreds of these formulations, but somehow such liveliness is absolutely not convincing
[49]... To our mind, it ought to be much more productive to carefully re-consider the logics of the conventional
trains of thoughts [109].

Indeed, when using of the ingenious ideas by Baron Rayleigh we are still situated in the infinitesimal
vicinity of equilibrium, and it is sheer impossible to study both the physical reasons and the actual stages of
approaching the equilibrium in such a way. This is why, to our mind, it is surely much too early to speak about any
clear-cut and final formulation of the “non-equilibrium thermodynamics”. Meanwhile, the mathematical work in this
direction is going on steadily and successfully (cf. [110,111] and the references therein). But, howbeit, the huge
benefit provided by Lars Onsager’s seminal work consists anyway in clearly demonstrating the insufficient
formulation of the conventional, so-called “equilibrium thermodynamics” – and – successfully invoking stimuli to
look for the proper physical re-formulation of the conventional thermodynamics. And one of such stimuli is
connected with the conventional logical train of thought clearly shown by Richard Becker [108]: we know that
entropy is somehow connected with irreversibility and we have the following two logical choices: either
irreversibility causes entropy production, or, vice versa, the former is a direct consequence of the latter. Remarkably,
there is, in fact, the third logical choice as well: Namely, some common (hidden) factor(s) might in principle
underlie both irreversibility and entropy [112]...

With all this in mind, we know that Lars Onsager made skillful use of statistical thermodynamics. Using the
latter enables us to dive into the microscopic (hidden) factors underlying the macroscopic observations. Therefore, it
is of crucial importance to analyze in detail the relevance and effectiveness of statistical thermodynamics as our
working instrument.

4. Statistical interpretation of thermodynamics

Whereas Eq. (1), the so-called Boltzmann-Planck equation, represents the well-known story about the bold
guess by Ludwig Boltzmann as for the interconnection between the notions of thermodynamic entropy and
probability, and therefore opens a way to the statistical representation of the Second Basic Law of thermodynamics,
the story about the attempts to introduce the statistical interpretation of its first basic law is much less discussed.
Nonetheless, the latter story seems to be of considerable interest and importance for our present discussion.

The attempts to establish a statistical interpretation start with the work by such prominent physicists as
Niels Bohr, Hendrik Anthony Kramers and John Clarke Slater [113]. They employ the well-known result by Einstein,
who was able to give a particularly simple derivation of the Planck's heat radiation law under the assumption that the
physical behavior of an atom in a given stationary state is being guided by probability laws. As a result, they came to

189 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

the conclusion that at the First Basic Law of thermodynamics, the Energy Conservation and Transformation Law,
might at some particular circumstances get obsolete at the microscopic level. In other words, one might in principle
imagine some microscopic situations when the latter law becomes obsolete, but such situations ought to be
extremely improbable.

Specifically, by the time when the publication [113] was being prepared, the seminal atomic theory of Niels
Bohr was already well recognized, but certain unsolved theoretical problems were still remaining, like the
difficulties when explaining radiation intensities and studying in detail the atoms possessed of a more complicated
electron structure than the hydrogen atom. Moreover, in its then state the Bohr’s theory could not be successful at
explaining such features essentially related to wave processes as wave interference and wave diffraction. The
classical wave theory was pretty successful at explaining them, whereas the novel quantum approach had come to a
kind of standstill. This is why the ingenious idea by Bohr, Kramers and Slater was to build a conceptual bridge
between the classical and quantum approaches as follows.

The classical theory dictates that atoms should lose some amount of energy by transferring it to the
radiation energy, while the latter one could then be absorbed by some other atom. And, according to the suggestion
by Bohr, Kramers and Slater, the interaction among the atoms ought to be based upon some kind of “virtual
radiation field” not capable of carrying any portion of energy. Therefore the energy released by the radiating atom
might in principle be lost, transferred to nothing. Vice versa, if some atom gets energetically excited, it is then not
absorbing the energy coming from elsewhere, from outside, but it creates the energy from virtually nothing…

This story is instructive for our current discussion owing to the two following points.

1. The above suggestion by Bohr, Kramers and Slater was carefully checked in several
experiments by independent laboratories, but it could not get any experimental support,
so that the work on the atomic theory was continuing along other directions (cf.
[114-122] and many other works).

2. And of extreme, of special interest for us was the reaction of Max Planck at this suggestion
[123].

Indeed, Max Planck has commented the relevant situation as follows:

“ … Auf den tiefen Ernst der hier vorliegenden Schwierigkeiten wirft ein bezeichnendes Licht der Umstand,
daß neuerdings von berufenster Seite sogar der Vorschlag gemacht worden ist, die Annahme der genauen
Gültigkeit des Prinzips der Erhaltung der Energie zu opfern – ein Ausweg, der wohl mit gewissem Recht ein
verzweifelter genannt werden darf, und der allerdings bald durch besondere Versuche als unzugänglich,
erwiesen werden konnte.”

190 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

“ … The deep seriousness of the difficulties described above casts a revealing light on the fact that most
recently a suggestion has even come from the utterly competent side to sacrifice the assumption of the exact
validity of the principle of conservation of energy – a definite way out, which probably may with some
justification be considered a desperate move, but could soon be proven to be inaccessible by special
experiments.”

Interestingly, in his work about the Second Basic Law published practically at the same time Max Planck
asserted forcefully contrariwise, namely as follows [72]:

“Jeder in der Natur stattfindende Prozeß verläuft in dem Sinne, daß die Summe der Entropien aller an dem
Prozeß beteiligten Körper vergrößert wird. Im Grenzfall – für einen reversiblen Prozeß – bleibt diese
Summe unverändert. Oder kürzer ausgedrückt: Die Entropie ist ein Maß für die Wahrscheinlichkeit (§ 3).

Damit ist der Inhalt des zweiten Hauptsatzes der Thermodynamik erschöpfend bezeichnet und zugleich den
Übergang zu der statistischen Definition der Entropie vorbereitet.”

“Every natural process takes place in the sense that the sum of the entropies of all bodies involved in the
process is increasing. In the limiting case – for a reversible process – this sum remains unchanged. Or, in
putting it more succinctly: The entropy is a measure of the probability (§ 3).

Herewith the contents of the second basic law of thermodynamics are exhaustively characterized and, at the
same time, the move to the statistical definition of entropy is made possible.”

Remarkably, in the introduction part of this same communication – before all the paragraphs including the
mentioned § 3 – Planck analyzed and criticized the work by Constantin Carathéodory. It is of considerable interest
for our present discussion to cast a detailed look at the logics of Max Planck. Here it would definitely be appropriate
to recall the conclusion by M. W. Zemansky as for the hard digestibility of the Kelvin-Planck’s statement of the
second law for students – as to the mathematical over-complications of the original Carathéodory’s approach [13], as
well as the assertion by P. Fong that Carathéodory ‘had stopped halfway’ [9]. Finally, it is extremely important to try
answering the poser of why Max Planck considered the statistical interpretation of the First Basic Law ‘a desperate
move’, whereas the statistical interpretation of the Second Basic Law – ‘an exhaustive solution’.

Constantin Carathéodory, an outstanding world-class mathematician, has published only two works on the
basics of the conventional thermodynamics [124,125]. To analyze the logical nuclei of the latter and provide it with
the clear rationale on par with the classical mechanics was in effect Carathéodory’s very aim. And he could achieve
this crucial goal in a generally comprehensive and abstract way. Carathéodory’s effort was met by the most of the
colleagues enthusiastically and triggered some further refinements and revisions [126-129]. It is important to
underline here that Carathéodory was by far not alone in his endeavor.

Substantially similar reports were published by a Hungarian mathematician-physicist, Farkas Gyula (Julius
Farkas, in German) [130], as well as by a Russian physical chemist, Nikolai Nikolaevich Schiller [131]. Remarkably,

191 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

although the both of the reports came out prior to those by Carathéodory, the report by Farkas remained unnoticed
until very recently [132]. And, as for Schiller’s report, it was originally published in Russian and consequently was
noticed for the first time only by Tatjana Ehrenfest-Afanassiewa, who published a detailed logical analysis of the
Carathéodory’s work vs. the Kelvin-Planck’s trains of thoughts [133]. She had also published a short note about her
coming across the Schiller’s work and accompanied her Russian note by a concise German translation. But, anyway,
all this wasn’t helpful in disseminating that important work by Schiller.

According to Tatjana Ehrenfest-Afanassiewa the main conceptual (both mathematical and physical) point
here consists in what follows. Conventionally, the system in question may be considered consecutively trespassing
some states quite similar to the equilibrium ones, if the pertinent real process is sufficiently slow. This is just what
they usually describe by such a mathematically explicable phrase: “The quasi-static processes are limiting cases of
the realistic infinitely slow processes”. In view of the usual over-interpretation of N. L. S. Carnot’s widget, the term
“quasi-static” is frequently substituted by the word “reversible”, by completely forgetting the fact that Carnot’s
‘widget process’ is indeed reversible, but only because its starting and final points are coincident – due to Carnot’s
ingenious idea, but not owing to some ‘mysterious physics’ underlying it. This is why, concluded Tatjana
Ehrenfest-Afanassiewa, from the beginning on it is much more physically appropriate to keep the identifiers
‘reversible’ and ‘quasi-static’ separate from each other.

Furthermore, any realistic process ought to go through the pertinent succession of the non-equilibrium
states. With this in mind Tatjana Ehrenfest-Afanassiewa suggests using the notation “non-static processes” to
describe such events, for the term “irreversible processes” usually employed to denote them isn’t appropriate at all,
as the very work by Carathéodory is in effect devoted to analyzing the processes’ irreversibility. Nonetheless it
should be noted that Caratheodory introduced the term “quasi-static process” himself [124], and in this connection
Tatjana Ehrenfest-Afanassiewa equitably stated that what he was actually considering ought to be conceived rather
as a “quasi-process”.

Further, the work [133] analyzes the notion of the infinitesimally small heat amount, dQ, accepted by the
system when it carries out some work on its surrounding, but first of all, it is placed into the expression for the First
Basic Law:

dQ = dU + dA, (2)

where dU and dA stand for the infinitesimal changes in the amounts of internal energy and work,
respectively. Now, to mathematically analyze changes in the system’s state, we introduce some state variables, x1 …
xN and assume that dQ can finally be cast as some proper function of the latter of the as follows:

dQ = Y1dx1 + … + YNdxN, (3)

where Yi = Yi(x1 … xN), i = 1 … N, stand for some functions of the system’s state variables determined by
the structure of the system under study.

192 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Eqs (2,3) are properly expressing the First Basic Law in that any dQ ¹ 0 means that the system acquires
some non-zero amount of heat from the surrounding, when the system works on the latter. The word “acquires”
should be understood in the algebraic sense, that is, the system might both “accept the heat from” and “transfer the
heat to” its surrounding, so that both dQ > 0 and dQ < 0 are equally possible, respectively. In effect, Eqs (2,3)
constitute the physical contents of the Caratheodory’s first axiom [124].

Besides, the processes with dQ = 0 stand for the so-called adiabatic ones. Consequently, adiabatically
isolated systems should have no heat exchange with its respective surroundings. And the Caratheodory’s second
axiom accordingly states [124]:

“In beliebiger Nähe jedes Zustandes eines Systems von Körpern gibt es Nachbarzustände, die vom ersten
Zustand aus nicht auf adiabatischem Wege erreichbar sind.”

“In any close proximity to any state of a system of bodies, there are neighboring states that are not
reachable from the first state by an adiabatic way.”

It is exactly this Carathéodory’s axiom that was forcefully attacked by Max Planck [72], but nonetheless,
this axiom is nowadays fully acknowledged to be the true logical basis for the definition of the entropy notion (see
[134] and the references therein). And Max Planck said in this regard:

“… Das Thomson’sche Prinzip zeigt sich dem Carathéodorys ohne weiteres überlegen. Denn während das
Problem des perpetuum mobile zweiter Art unzählige Male experimentell behandelt worden ist, hat wohl noch
niemand jemals Versuche angestellt in der Absicht, alle Nachbarzustände irgendeines bestimmten Zustandes auf
adiabatischem Wege zu erreichen. Indessen ist diese Überlegung nicht ausschlaggebend, da doch auch das zweite
Verfahren zur Verfügung steht und da im vorliegenden Falle die zahlreichen experimentellen Bestätigungen der
einzelnen Folgerungen des zweiten Wärmesatzes vorliegen, welche sich aus dem Carathéodory’schen Prinzip genau
so ergeben wie aus dem Thomson’schen.

Ja, man könnte mit Rücksicht auf diesen Umstand vielleicht geneigt sein, der ganzen hier zur Diskussion
gestellten Frage im Grunde nur eine formale Bedeutung beizumessen. Dies wäre ja nur dann richtig, wenn die
Thermodynamik ein in sich vollkommen fertig abgeschlossenes isoliertes Gebiet im Rahmen der physikalischen
Wissenschaft darstellen würde. Das ist aber bekanntlich keineswegs der Fall. Im Gegenteil bildet die sogenannte
allgemeine Thermodynamik in dem System der gegenwärtigen Physik, welche durchweg auf atomistischer
Grundlage aufgebaut ist, einen speziellen Teil, einen gewissen Grenzfall, der sogar streng genommen niemals mit
absoluter Vollkommenheit verwirklicht ist.

Wenn man von diesem Standpunkt aus die Frage betrachtet, gewinnt sie ein ganz anderes Gesicht. Dann
verliert der zweite Wärmesatz seine prinzipielle Bedeutung, er erscheint nur mehr als ein statistischer Satz, gültig
nicht für die Eigenschaften eines einzelnen Körpersystems, sondern nur für die Mittelwerte aus den Eigenschaften
einer sehr großen Anzahl von makroskopisch identischen Exemplaren des Körpersystems. Die Schwankungen der
einzelnen Werte um die Mittelwerte sind umso beträchtlicher, je weniger Freiheitsgrade das System besitzt.”

193 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

„ ... The Thomson's principle readily proves to be superior to the Carathéodory’s. While the problem of
perpetual motion of the second kind countless times has been treated experimentally, probably no one has ever
experimented with the intention to reach all neighboring states of any particular state by an adiabatic way. This
consideration might seem to be not fully conclusive, but another viewpoint is also available. Indeed, in the present
case all the numerous experimental confirmations of the Second Law consequences result from the Carathéodory's
principle just in the same manner as from the Thomson's principle.

Yes, with respect to the latter fact, one might be basically inclined to attach just a formal meaning to the
whole discussion on the question posed here. This would only be correct if the thermodynamics ought to be a
perfectly formulated, complete in itself, isolated area within the physical science. But this is known to be definitely
not the case. On the contrary, within the structure of the current physics, which is entirely built upon the atomistic
basis, the so-called general thermodynamics serves as a specific part, as a certain limiting case, which is, strictly
speaking, still not built-up with the absolute perfection.

If you look at the problem from the above standpoint, everything gains a completely different appearance.
Then the second set of heat loses its fundamental importance, it appears only as a statistical record, not valid for the
properties of a single body system, but only for the mean values of the properties of a very large number of
macroscopically identical copies of the body system. The fluctuations of the individual values around the means are
all the more significant, the fewer degrees of freedom has the system.“

Logically seen, in full recognition of the tremendous importance and conceptual difficulties connected to
the conventional thermodynamics, Max Planck did perform here the same ‘desperate jump’, just as Bohr, Kramers
and Slater, by doffing the general applicability of the Second Basic Law and suggesting the statistical interpretation
of the latter. Whereas the suggestion by Bohr, Kramers and Slater as concerns the First Basic Law could finally be
overthrown by the numerous careful experiments, Max Planck was absolutely on solid ground, for it shouldn’t be a
trivial task to experimentally properly check the validity of the Carathéodory’s second axiom. So that, we seem to
have no more ways out than just to follow the conventional Kelvin-Plank’s train of thoughts … The only remaining
thing would presumably be the parting words by Charles-Maurice de Talleyrand-Périgord: Il définissait la statistique
comme “l'addition correcte de chiffres faux.” (Statistics is the correct addition of the false numbers – and this ought
to be the best possible expression for the essence of the SORITES paradox!) … Most probably, this is just why the
notion of entropy still remains a ‘Protean Concept’ [135] – but the situation is by far not so forlorn.

Tatjana Ehrenfest-Afanassiewa commented this conceptual inconsistence between the Kelvin-Plank and
Carathéodory’s approaches in her work [133] published in Russian. Specifically, she stated as follows.

Carathéodory has surely gone valuable steps in defining the “quantity of heat” and the “absolute temperature”,
but he based these notions upon the non-static processes in willing to consider the realistic processes. However, it is
also throughout possible to place the proper grounds under the whole thermodynamics of the “quasi-static”
processes without taking the non-static processes into account – up to the point where it is necessary to define the
internal energy U as a function of the state variables x1 … xN – such an additional possibility is also of extreme value,

194 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

for the very physical essence of the Second Basic Law becomes clear only after the separation of the quasi-static
approach from the non-static approach.

Interestingly, Tatjana Ehrenfest-Afanassiewa underlined in this regard that Nikolai Nikolaevich Schiller
[131] had introduced the holonomic property without invoking the irreversible processes, unlike Carathéodory – and
thus Schiller could be considered to have come nearer to the proper, that is, to the more general, understanding of the
Second Basic Law’s physical essence than Carathéodory.

Still, Carathéodory had introduced the absolute temperature as the general integrating divisor for energy –
and was thus capable of truly generalizing the notion of entropy, whereas Max Planck didn’t considered this
achievement as something really special [136]. He just dealt with the ideal gas and divided dQ by T, thus declaring
the result to be entropy, and verified that the result is truly a full differential. To prove the generality of the latter
result, he just combined the ideal gas with some other arbitrary system, made the combined system to follow the
circular process and required that there is no heat from the outwards. Mathematically, he performed the contour
integration of the total entropy (the gas plus the system) and got zero as the result, as anticipated. He stated that, as
the total entropy of the ideal gas after the cyclic process is anyway equal to zero, then the resulting total entropy of
the system should also be equal to zero. And, as we haven’t specified the system before taking it into consideration,
we get therefore the general result about the entropy of any arbitrary system.

If this would really be the case, then due to the possibility of considering the ideal gas, the nature might
produce nothing more than adiabatic systems. Or – another kind of drawing conclusion – as we obtain some result
for the ideal gas, it could be generalized for any kind of heat exchange processes. These both trains of thoughts are
clearly not valid.

The Max Planck’s logical failure could be described as follows [133]: As soon as we have no additional
information about our system under study, we cannot guarantee that it will without fail run through the trajectory of
the cyclic process in exactly the same way as the ideal gas does this. And, most probably, it is exactly such a kind of
underestimate that made Planck to place the entropy increase property to the absolute foreground of the Second
Basic Law, without thinking in detail about what should this increase mean physically. And this is, most probably,
why he focused all of his attention on the impossibility of the perpetuum mobile, by rejecting the Carathéodory’s
achievements... But, as we noticed already: Dear Colleagues! Errare Humanum Est!

On the other hand, one cannot deny the importance and undoubted successes of the conventional statistical
mechanics, one cannot deny the atomistic/molecular picture of the matter – but one has to rework the conventional
logical frameworks – to refine the whole field…

Finally, there is even a bit more to the story.

At approximately the same time as Carathéodory’s work proceeded, the mathematical working tools of
thermodynamics had also been discussed in detail in the book by a prominent American mathematician William F.
Osgood [137]. The readership can immediately recognize that the actual mathematics of thermodynamics ought to

195 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

be readily digestible for good high-school graduates. Meanwhile, Osgood wasn’t entering any detailed discussion of
the physical logics, by sending his readership away to the discussions of the relevant physics and logics in the
appropriate high-quality handbooks [39,138,139].

Thus, the true problem of the conventional thermodynamics is not its complicated mathematical toolset and not
the absence of the proper handbooks, but rather logical complications. Here we would like to cite the clever words by
Otto Redlich [140].

“ … In the last 100 years, several authors noticed that the variables now called generalized coordinates and
forces belong to two classes, differing from each other and from all other variables. But repeated attempts to
characterize these classes were unsuccessful. Consequently, the concept of work, defined as the integral of a
generalized force with respect to its generalized coordinate, never had a solid foundation. On the basis of the present
discussion, work can be introduced in a clear and unambiguous manner. Temperature can be defined in the
conventional way. The path to the first and the second law has been shown by Carathéodory. Serious mistakes have
been the curse of thermodynamics. They have caused the uncertainty and uneasiness that often have emerged. They
have been repressed by the habit-forming procedures of applying thermodynamics. Wrong definitions of extensive and
intensive quantities, of generalized coordinates and forces, and defective definitions of work have been quite common.
A historical accident – namely, Tolman's unawareness of a prior definition – has greatly contributed to a widespread
confusion of extensive and intensive quantities with coordinates and forces. The so-called zero-th law going back to
Carathéodory is not only unnecessary but actually misleading. The problems of the relationship between the physical
sciences and mathematics become more manifest in thermodynamics than in any other part of science. The wonderful
efficiency and elegance of so many mathematical tools, starting from the concept of the limit, should not induce us to
see in them more than tools. The theoretician as well as the experimenter must be the master of his tools, not their slave.
It has been the purpose of the present discussion to show that thermo-dynamics can be developed in a clear and
consistent conceptual structure. We may conclude that thermodynamics can be understood.”

We can only add here that not only the Zero-th Law, but also the so-called Third Basic Law (Franz Simon: “Es ist
unmöglich, eine Substanz vollständig ihrer Entropie zu berauben.” – “It is impossible to completely deprive any
substance of its entropy.”) belongs to the definite legacy issues (Altlasten) of physics. The point is that when
calculating the total entropy using the conventional integral formula, we have to define some proper constant of
integration [140]. There are two Basic Laws in thermodynamics. Max Planck was aware of some intrinsic
interrelationship between them (cf. the preface to the 1st edition of his thermodynamics handbook [136]), but could not
manage to study the nature of this relationship. This is what we now ought to do…

5. Conclusions

To sum up, according to the actual common standpoint, thermodynamics in its classical sense is nothing
more than a scientific discipline based upon the general laws of nature and employing the mathematical tool of unit
total differentials. Its use is only possible in the study of idealized equilibrium systems, as well as in the study of
quasi-isolated or quasi-closed systems and quasi-equilibrium processes.

196 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

There are virtually no objections to the first statement, whereas the second one slights the actual field of
thermodynamics – viz. – substantially.

The reasons for such a substantial slighting are fully impartial and historically justified: The true
thermodynamics seems to have gone astray in the whirlpool of the physical revolution at the beginning of twentieth
century, as the minds of the scientific vanguard were fully concentrated on the quantum physics, relativity theory etc.
etc. … This isn’t meant to diminish the mentioned giant achievements of the natural science – this is solely to
remind that revolutionary crowds might sometimes be truly inattentive …

Indeed, the very starting point of the whole story was the fact that N. L. S. Carnot had triggered developing
the true and quite general “mechanical theory of heat”, but could not put this work to the end himself for objective
reasons. He managed to qualitatively show the “back side” of the energy conservation and transformation law, which
he was very well aware of (although at the time of his active research there were still no publications about this
law!). Slighting the clear analogies put forward by N. L. S. Carnot – namely those connected with waterfalls – as the
“primitive leftovers”, “useless remnants” of the old demising phlogiston theory – aren’t the proper estimates. First of
all, as we have seen above, the latter theory wasn’t at all false and primitive. Secondly, N. L. S. Carnot tried to
present a qualitative picture of a true mechanical event having to do with thermal events. He was looking for the
driving forces of the latter processes – or, “actions”, using the language of mechanics – and even managed to
approach the pertinent “reactions”. His ingenious cyclic model had clearly demonstrated the effect of such
“reactions” which do not allow performing the desired “action” with all the 100% of the possible efficiency –
independently of all the striving to improve the state of affairs by suggesting the thoroughly idealized device, which
is anyway impossible to produce in reality. The open questions implicitly posed, but, to our sincere regret, not
explicitly answered by N. L. S. Carnot himself are in fact as follows:

a) What are the true physical reasons for the “actions” in connection with heat processes?

b)What are the true physical reasons for the “reactions” in connection with the above-mentioned
“actions”?

What is here but as plain as the noses on our faces, ought to be the clear – although purely qualitative –
pronouncement of N. L. S. Carnot that these “actions” and “reactions” are definitely compensating each other, or, as
he had even put it, ‘equilibrating each other’. And it is exactly this equilibration, which puts every realistic process
to its logical and actual end. And this is just the true physical sense of the so-called, notorious, “thermodynamic
equilibrium”.

The next very important methodological step of the Carnot’s train of thoughts was to formulate his story
using the mathematical language. This was first achieved by a prominent French engineer Benoît Paul Émile
Clapeyron, who had also presented the graphical images of the Carnot’s ingenious widget. But Clapeyron had not
enough time to fruitfully develop the theory by Carnot, for he finally had to depart.

197 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

The Clapeyron’s interpretations had become accessible to Rudolf Clausius and William Thomson (Lord
Kelvin), whose lines of research were fully independent of each other, but they both had drawn very similar
conclusions. First of all, they had absolutely correctly recognized that the whole story is rotating about the sheer
impossibility of the perpetuum mobiles, but – from the beginning on – revealed two independent laws of
thermodynamics, contrary to what N. L. S. Carnot was suggesting (in fact, his initial idea was – to reformulate it in our
modern language: a single general law with two mutually opposite aspects). Interestingly, the both of them were in
detail investigating the notion of the “thermal equilibrium” – and Clausius had even explicitly introduced the notion of
“compensation”, but they couldn’t pursue their work for quite different reasons, but anyway due to their natural
departures – and this is why the both thermodynamics laws were – and still are – remaining basically independent from
each other…

…Still, the work on thermodynamics was continuing during all the final years of the XIX-th century. This can
clearly be following from the then literature. For instance, the seminal book by an outstanding French professor of
physics from Nancy, M. Prosper-René Blondlot [138], the second edition of which was published at the beginning of
the XX-th century and translated into German. Well, the name of M. Prof. Dr. Blondlot might sound not quite
persuasively in our days, due to his definite mistake in connection with the so-called “N-rays”, but he was referring in
detail to the relevant works by such eminent colleagues of him, as Henri Poincaré [39], Jonas Ferdinand Gabriel
Lippmann [141]. Furthermore, noteworthy with respect to our present discussion are also the relevant works by a
prominent French mathematician, Joseph Louis François Bertrand [142] and an outstanding French engineer,
examinateur a l’École polytechnique, Université du Paris-Saclay, Jules Moutier [143], for all the named authors were
persistently trying to clearly and explicitly answer the two implicit questions by N. L. S. Carnot…

Interestingly, in the introduction to his book, M. Prof. Dr. Bertrand had told an interesting and instructive
story about the term ‘Vis Viva’ (‘livening force’) which sounds as follows.

“…Un vieux professeur m'a raconté que, il y a cinquante ans environ, un étudiant, qui déjà se croyait un
maître, avait choisi pour sujet de thèse, à la Faculté des Sciences de Paris, les applications du principe des forces
vives. La démonstration du principe fut la première question qu'on lui adressa; il parut fort surpris. On ne peut pas,
dit-il, démontrer un axiome. Les juges, fort surpris à leur tour, lui refusèrent le diplôme.

L'étonnement serait moindre aujourd'hui. Un grand nombre de savants, instruits avec moins de peine et
devenus intolérants, traiteraient volontiers d'ignorants ceux que de plus sérieuses études conduisent à faire des
réserves.

Il est sur le principe des forces vives que reposent les travaux admirés auxquels on a donné le nom, fort mal
choisi, de Théorie mécanique de la chaleur. Le travail interne des molécules d'un corps ne dépend, dans une
transformation quelconque, que de l'état initial et de l'état final. Telle est la base de la théorie. On allègue le principe
des forces vives et l’on passe outre.

Le principe des forces vives ne rend l'assertion évidente qu'à la condition de fermer les yeux à des difficultés

198 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

très sérieuses.

Les actions mutuelles des molécules doivent s'exercer suivant la droite qui les joint et dépendre de la seule
distance. A priori, l'évidence est douteuse. La chaleur, dit-on, est un mouvement des molécules matérielles. L'idée est
ancienne. Partout où se trouve une suffisante vitesse, disait Descartes, dans les parties des corps terrestres, il y a du
feu. Sans disconvenir de l'assertion, est-il permis de voir dans ses conséquences une théorie de la chaleur? Un corps
chaud, par sa présence, échauffe les corps voisins. Il accroît donc la force vive de leurs molécules. Mais jamais on n'a
vu un mouvement, par son seul voisinage, en influencer un autre; il faut que des forces interviennent.

D'où viennent ces forces? La réponse n'est pas douteuse: les parties de l'éther, violemment agitées, comme
dirait Descartes, sont la cause de l'action.

Les molécules matérielles agissent donc sur l'éther et l'éther sur elles. Ces actions, dont on ignore la
grandeur et la loi, interviennent dans tous les phénomènes; elles semblent s'imposer dans les raisonnements. On ne les
mentionne même pas. Le principe des forces vives suffit à tout. Ces forces remplissent-elles au moins les conditions
sans lesquelles on ne peut l'appliquer?

Rien a priori ne le rend vraisemblable.”

“An old professor told me once, that about fifty years ago there was a student at the Faculty of Sciences in
Paris who considered himself being already a master enough to try defending his PhD about the applications of the
‘livening force’ principle. And what is the proof of principle – this was the first question he encountered; and he
seemed surprised. Well, but we can’t prove an axiom, he said. The jurors were surprised in turn and had refused him
the diploma.

The astonishment would be lower nowadays. Many scholars are being educated according to the motto – the
less trouble the better – and thus become really intolerant. They willingly prefer to be treated as ignorant, while
postponing all the more serious studies to some indefinite future.

It is the principle of the ‘livening force’ that represents the true basis of the scientific branch bearing the
(badly chosen) name: ‘Mechanical Theory of Heat’. Indeed, the internal working of the molecules of a body during
any transformation depends upon the initial and the final states. And this is the basis of the whole theory. This ought to
boil down to the principle of the ‘livening force’ – but no heed is being paid to the latter.

In effect, the ‘livening force’ principle might render such an assertion obvious, if one turns blind eyes to the
very serious difficulties.

The interactions among the molecules must be exercised along the straight line joining them and depend
solely on distances between them. Then, a priori, the following statement is questionable. Heat, they say, is a
movement of material molecules. This idea is rather old. Wherever there is a sufficient speed, said Descartes, meaning
the terrestrial bodies, there is fire. Without disagreeing with such an assertion, is it permissible to see in its

199 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

consequences a theory of heat? A hot body, just by its presence, is capable of warming up its neighbors. Hence, the
momentum of their molecules increases. But there has never been any movement, which is due to nothing more than
just the neighbors somehow influencing each another; some forces must yet operate.

And where these forces might then come from? The answer is straightforward: the parts of ether, violently
agitated, as Descartes would say, are the cause of action.

The molecules of the material should therefore act on the ether and vice versa. Such actions, apart from the
system’s size and the corresponding law, ought to be involved in all phenomena; they seem to prevail in reasoning. So,
we should not even mention this. Indeed, the principle of ‘livening force’ does enclose all of this. But do such forces at
least meet the conditions without which they couldn’t be applied?

Nothing could be rendered a priori probable.”

The above-listed are most probably the conceptual difficulties that prevented many colleagues from finding
the true interconnection between the macroscopic thermodynamics and the classical mechanics. These same
difficulties seem to have finally caused P. G. Tait to reduce the notion of “livening force” to just the mechanical kinetic
energy [141, page 17] – “… and the energy produced is measured by half the product of the mass into the square of the
velocity produced in it. This active form is called Kinetic Energy, and it is the double of this to which the term Vis Viva
(erroneously translated as Living Force) has been applied.” And – most probably, with this in mind – he had come to
the idea of rendering the entropy responsible for the actual driving force, as we have already discussed above (but now
we regret to have no possibility to ask him directly!). Well, on the other hand, P. G. Tait has earlier [142, page 72] stated
on this theme as follows:

“In all these cases the potential energy involved, whether it depends upon molecular forces, as in a spring, or
upon external forces, as gravity, is of the same species as that of a raised weight; and the only form of kinetic energy
contemplated is that of visible motion. And here there is constant transformation from one of these forms to the other,
and back again, forever, without loss by dissipation, as the process is in each case exactly reversible. They give us,
therefore, little insight into the more complex phenomena to which we proceed. They are all summed up in the law of
conservation of Vis Viva, which we have already seen to be merely a different form of statement of one of Newton's
discoveries. But in the ordinary text-books, the loss of Vis Viva in the impact of imperfectly elastic bodies is asserted,
and its amount calculated; not a hint being given that the so-called loss is merely a transformation, partly, no doubt,
into the potential form of distortion of the impinging bodies, but mainly into the kinetic form – heat. The same
textbooks also assert that there is no loss of Vis Viva in the impact of perfectly elastic bodies. This is, of course, true,
but not in the sense in which it is asserted, since in the case of impact of perfectly elastic bodies, a portion of the Vis
Viva of each would be changed, in general, into vibrations of the body itself, and would, therefore, not appear as part
of the Vis Viva of the body considered as moving as a whole. Take, for instance, the case of a bell and its clapper, both
supposed perfectly elastic.”

Indeed, after reading the above citations it becomes much clearer, why P. G. Tait was viewing the entropy as
the actual driving force in thermodynamics – he made absolutely no difference between the total energy and the Vis

200 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Viva in the system under study, by considering a purely Hamiltonian case of conservative systems, that is, without any
dissipation (hence, in the systems he considered the potential energy is transferred into the kinetic energy without any
perceptible loss). But the Newton’s discoveries he mentioned ought to contain one more important element – the
action-reaction duality – and, hence – the eternal dissipation of the Vis Viva… Moreover, both Rudolf Clausius and
Lord Kelvin maintained the ‘energy dissipation’ as the event behind the ‘entropy production’… Most probably, these
are the latter two considerations that had led Carathéodory to introducing his (fiercely mooted) second axiom about the
‘adiabatic non-reachability’ (which in effect introduces and maintains the absolutely essential role of entropy in any
surmisable process)…

Interestingly, the priority of introducing the notion of ‘Potential Energy’ could clearly be ascribed to the
father of N. L. S. Carnot – Lazare Nicolas Marguerite Carnot [143, page 37] who had discussed the notions of ‘force
vive et force vive latente’ as follows:

“58. Sous le nom simple de forces ou puissances, ou de forces proprement dites, on comprend les quantités de
mouvement et les forces motrices; ou si l'on veut, les forces de percussion et celles de pression, parce qu'elles sont
soumises aux mêmes décompositions et aux mêmes lois. Mais lorsqu'on veut désigner une force vive, on ajoute
toujours l'épithète qui la caractérise, c'est-à-dire le mot vive.

59. Nous venons de voir que la force vive peut se présenter ou sous la forme M*u 2 d'une masse par le carré
d'une vitesse, ou sous la forme PH d'une force motrice par une ligne. Dans le premier cas, c'est la force vive
proprement dite; dans la second, on pourroit lui donner la dénomination particulière de force vive latente.”

“58. Under the simple name of forces or powers, or actual forces, it includes the momenta and driving forces;
or, if you wish, the impact forces and those of pressure, because they are subject to the same decomposition and the
same laws. But when you would like to designate a living force, it always adds the epithet that characterizes it, that is
to say, the word ‘living’.

59. We have seen that the momentum can be expressed in the form the M*u2 as a mass by the square of speed,
or in the form PH as a driving force by a line distance. In the first case, it is the actual driving force; in the second, we
might give it a special name latent momentum.”

And it is exactly here that the intrinsic professional interconnection between the father and the son Carnot [34]
could be clearly followed.

Therefore, P. G. Tait is right in that the Vis Viva is, on the one hand, not just some philosophically elusive
‘Living Force’. But Vis Viva is also much more than solely a duplicate mechanical kinetic energy – in fact, it is
‘Livening Force, or Livening Energy’, that is, the very part of the total energy, which is capable of livening up the
real processes, that is, the ‘Livening Energy’ is just the Useful Energy, capable of producing the desired work. And
the Vis Viva is definitely not conserved – it is dissipated, as Rudolf Clausius and Lord Kelvin had equitably stated –
for it is spent when producing the useful work. And why is it always spent? The answer is clear – because of the
ubiquitous, omnipresent hindrances and hurdles. The Vis Viva is thus spent to overcome the latter – and, if the bank

201 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

of Vis Viva could be large enough to successfully overcome all the hindrances and hurdles available – the process
ought to successfully reach its desired final point. To this end, the actual bank of Vis Viva is nothing more than the
corresponding amount of Potential Energy.

And what N. L. S. Carnot would actually like to tell us was that – to produce some useful work – firstly, we
need enough Potential Energy – and, secondly, the latter will be converted into Vis Viva – and, thirdly, while
performing the useful work, the Vis Viva will be dissipated into two directions: a) to promote the desired process
and b) to equilibrate all the hindrances and hurdles available on the way to the desired result, whereas the total
energy is always perfectly conserved. With this in mind, we have not two, three, or even four laws of
thermodynamics, but just two interdependent laws – or – two different sides of one and the same law of
thermodynamics: a) the energy conservation/transformation and b) the energy realization/dissipation.

A propos, there are two marvelous modern books analyzing the numerous difficulties with the entropy notion
[144] and the true essence of the energy and entropy notions [145]. That “… In the latter part of the nineteenth
century, many attempts were made to explain the second law in terms of mechanics; none were successful. Finally,
the thermodynamic laws found a statistical interpretation in the microscopic theory by Boltzmann, which gave rise
to a new branch of physics – statistical thermodynamics …” [9] was definitely the result of under- and/or over-
interpretation of N. L. S. Carnot’s work and – therefore – not taking into account the intrinsic dialectic
interrelationship between the energy and the entropy.

Now, as for the statistical interpretation of thermodynamics – was it just a plain success or solely a sheer
despair? To our mind, the correct answer would be – the both. Why?

Indeed, as we have seen above, one of the two thermodynamics laws (the 1 st one) cannot be considered a kind
of statistical regularity, whereas another one (the 2nd one) – could nonetheless be made to obey the statistical
regularities. How could then we combine this result with the clear dialectic interrelationship between the two laws in
question?

The first (to the best of our knowledge) interesting suggestion as to possible ways out of the above-mentioned
logical blind alley came from Johannes Diderik van der Waals as early as in 1911 [146]. In his short note Prof. Dr.
van der Waals had discussed the interrelationship between the probability and entropy notions, as it appeared from
the considerations by Boltzmann and Gibbs. And he clearly stated as follows:

“… Nur wenn wir aus der wahrgenommenen Serie oder aus dem untersuchten Gebiete schließen, daß wir die
Wahrscheinlichkeit a priori unrichtig abgeschätzt hatten, und daß, zum Beispiel, die Punkte auf der Fläche nach
einer gewissen Regel angeordnet worden sind, würden wir aus den beobachteten Ereignissen eine Schlußfolgerung
über die noch nicht beobachteten ableiten können. Um zu untersuchen, ob wir zu einem derartigen Schlusse
berechtigt sind, muß bekanntlich die Bayes’sche Regel herangezogen werden. ... Nur glaube ich gezeigt zu haben,
daß es ohne die Annahme eines solchen Prinzips unmöglich ist, die Wahrscheinlichkeitsbetrachtungen mit der
kontinuierlichen Geltung des Entropieprinzips in Einklang zu bringen.”

202 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

“… Only if we might conclude from some experimentally observed series or from some examined areas that our
a priori probability estimation turns out to be incorrect, or that, for example, the points on the surface have been
arranged according to a certain rule, we would be able to derive a conclusion about the events not yet observed
from the already observed events. To investigate whether we are entitled to such a conclusion, the well-known
Bayesian rule ought to be used. ... Anyway, I guess I could have shown that, without the adoption of such a principle,
it is impossible to bring the probability considerations in line with the solid validity of the entropy principle.”

But, as we know, this ingenious suggestion van der Waals had in effect fallen on deaf ears – in the literature
one might find only a volatile comment by Myron Tribus and Edward C. McIrvine [147], but not in his earlier
detailed review paper together with Robert B. Evans in 1963 [148]:

“… From Maxwell’s time on many leading investigators pondered the relation between observation and
information on the one hand and the second law of thermodynamics on the other. For example, in 1911 J. D. van der
Waals speculated on the relation between entropy change and the process of reasoning from cause to effect. In 1929
Leo Szillard commented on the intimate connection between entropy change and information. In 1930 G. N. Lewis
wrote: “Gain in entropy means loss of information; nothing more”. Until Shannon came on the scene, however,
there was no measure of information, so that the discussions could not be quantitative. …”

As to the real value of the notorious ‘Maxwell’s demon’, the stories like that particular writing by Leo
Szillard together with all of what it had triggered, etc. – and, especially, as for the complete absence of any actual
conceptual interrelationship between the thermodynamic entropy and the Shannon’s information measure: Dear
readership, please refer to the marvelous book by the late Prof. Dr. Stephen Jay Kline [149].

As to the contribution by J. D. van der Waals, that was in no way ‘just a speculation’ – but a clearly
formulated, substantiated, justified and fully competent suggestion to employ the Bayesian approach in deriving the
relationship between the entropy and probability notions. The only serious and effectual obstacle on the way to the
practical embodiment of that ingenious suggestion by van der Waals was the fact that the Bayesian approach to the
probability notion wasn’t really ‘trendy’ – not only at his time – but even approximately until the time of after the
Second World War – it had to succumb to the ‘frequentist’ train of thoughts…

As for the story of Prof. Dr. Gilbert Newton Lewis, he was really one of the most serious investigators of
the true sense of the entropy notion. He had a lot of undoubtedly serious accomplishments in the fields of general
physical chemistry – and chemical thermodynamics, in particular. The latter line of his work is clearly documented
in detail and very well recognized [150-151]. Still, his own deliberations concerning the possible ways of revealing
the true interconnection between the probability and entropy notions sound very interesting and conversable, but
somehow helpless [152, Chapter 6, pp. 135-162]…

Nonetheless, one of the G. N. Lewis’ actual apprentices, Dr. George Augustus Linhart, had managed not
only to fruitfully employ the Bayesian approach in formally deriving the famous Eq. 1 – the well-known
Boltzmann-Planck relationship, but also to recognize the true logical structure of the conventional thermodynamics.
To our regret, G. A. Linhart was successful in publishing only a number of papers about the first line of his research

203 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

(see [153-158] and the references therein), whereas the most important second line remained in a number of
practically unattainable preprints (fortunately, something could nonetheless be found, at least) – until very recently
(see [63-66] and the references therein)... Well, the true story of G. A. Linharts’ personal and professional life ought
to deserve a separate monograph, which is presently in preparation…

And – the last, but anyway not the least important topic – should be the non-equilibrium thermodynamics.
The seminal work by Lars Onsager clearly requests for its corresponding further development, because it solely
describes the behavior of the systems in the nearest proximities to their thermodynamic equilibrium states. There is a
very active research work underway in this direction (see, e. g., [110-111] and the references therein), so that only
one technical point ought to be taken into account: One ought to consider, how realistic systems might manage
reaching their equilibrium states in view of the dialectic correlation between the energy and entropy notions. The
so-called “entropy production forces” being the ‘true driving forces’ of the corresponding processes ought to have
nothing to do with the actual physical, chemical, biological etc. reality.

To sum up, here are the possible ‘road-map’ directions of the research work in the field of thermodynamics,
as far as we might recognize them:

1. The conventional statistical thermodynamics ought to be reformulated without explicit application to


microscopic/atomistic hypothesis – in order to avoid encountering the SORITES paradox.
This work could be based upon the results of G. A. Linhart and it is already underway [64].

2. A proper re-building of the Lars Onsager’s seminal theory, with taking into account the mutual
compensation of the pertinent driving forces and hindrances/hurdles, that is, the
energy-entropy compensation. This work is also ongoing and will be reported in detail
elsewhere. Here we would only like to mention that the compensation ought in effect to mean
the competition among many factors (factor sets) opposing each other – so that the relevant
mathematical machinery ought to be connected with the theory of games.

204 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

References

1 Van Ness HC, Understanding Thermodynamics 1969, Courier Dover Publications, NY, USA.

2 Plank R, Begriff der Entropie. Grenzen der Gültigkeit des zweiten Hauptsatzes, Zeitschrift des Vereines
Deutscher Ingenieure 1926 70: 841.

3 Keenan JH, Thermodynamics 1941, John Wiley & Sons, Inc., NJ, USA.

4 Hatsopoulos GN, Keenan JH, Principles of General Thermodynamics 1965, John Wiley & Sons, Inc., NJ,
USA.

5 Bridgman PW, Nature of Thermodynamics 1941, Harvard University Press, MS, USA.

6 Pippard AB, Elements of Classical Thermodynamics 1957, Cambridge University Press, UK.

7 Buchdahl HA, Twenty Lectures on Thermodynamics 1975, Elsevier Science & Technology Books, the
Netherlands.

8 Buchdahl HA, The Concepts of Classical Thermodynamics 2009, Cambridge University Press, UK.

9 Fong P, Foundations of Thermodynamics 1963, Oxford University Press, UK.

10 Silver RS, An Introduction to Thermodynamics 1971, Cambridge University Press, UK.

11 Chambadal P, Paradoxes of Physics 1973, Corgi Children’s, UK.

12 Angrist SW, Hepler LG, Order and Chaos 1973, Pelican London, UK.

13 Zemansky MW, Fashions in Thermodynamics, Am. J. Phys. 1957 25: 349.

14 Truesdell CA, The Tragicomic History of Thermodynamics, 1822-1854 1980, Springer, NY, USA.

15 Truesdell CA, Bharatha S, The Concepts and Logic of Classical Thermodynamics as a Theory of Heat
Engines 1989, Springer, NY, USA.

16 Reflections on the Motive Power of Heat. Edited by R. H. Thurston. John Wiley & Sons, NY, USA; Chapman

205 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

& Hill, London, UK, 1897.

17 The Second Law of Thermodynamics. Translated and edited by W. F. Magie. Harper & Brothers Publishers,
NY, USA, London, UK, 1899.

18 Reflections on the Motive Power of Fire. Edited with an introduction by E. Mendoza. Dover Publications, NY,
USA, 1960.

19 Betrachtungen über die bewegende Kraft des Feuers und die zur Entwickelung dieser Kraft geeigneten
Maschinen. Übersetzt und herausgegeben von W. Ostwald. Verlag von Wilhelm Engelmann, Leipzig, Germany,
1892.

20 Размышления о движущей силе огня и о машинах, способных развивать эту силу. Перевод С. Э. Фриша.
Под редакцией и с примечаниями В. Р. Бурсиана и Ю. А. Круткова. М.-П.: Госиздат, USSR, 1923.

21 Второе начало термодинамики. Под редакцией и с предисловием А. К. Тимирязева. М.-Л.: ГТТИ, USSR,
1934.

22 Callender HL, Carnot’s Principle, Proc. Phys. Soc. London 1910 23: 153.

23 Larmor J, On the Nature of Heat, Proc. Roy. Soc. London 1918 94: 326.

24 Lunn AC, The Measurement of Heat, Phys. Rev. 1919 14: 1.

25 La Mer VK, N. L. Sadi Carnot, Am. J. Phys. 1949 109: 598.

26 La Mer VK, Some Current Misinterpretations, Am. J. Phys. 1954 22: 20.

27 La Mer VK, Some Current Misinterpretations, Am. J. Phys. 1955 23: 95.

28 Kuhn TS, Carnot’s version of “Carnot’s cycle”, Am. J. Phys. 1955 23: 91.

29 Hirshfeld MA, On ‘Some Current Misinterpretations’, Am. J. Phys. 1955 23: 103.

30 Mendoza E, A Sketch for a History of Early Thermodynamics, Phys. Today 1961 14: 32.

31 Thomsen JS, Hartka TJ, Strange Carnot’s Cycles, Am. J. Phys. 1962 30: 26; 30: 338.

32 Frank FC, Reflections on Sadi Carnot, Phys. Educ. 1966 1: 11.

33 Pitteri M, Classical Thermodynamics of Homogeneous Systems Based upon Carnot’s General Axiom, Arch.

206 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

Rat. Mech. Anal. 1982 80: 333.

34 Pisano R, Gillispie ChC, Lazare and Sadi Carnot: A Scientific and Filial Relationship 2013, Springer NY,
USA.

35 Kauzmann W, Thermodynamics and Statistics 1967, W. A. Benjamin, Inc., New York, Amsterdam.

36 Reech F, Théorie générale des effets dynamiques de la chaleur, Journal de Mathématiques Pure et Appliquées
1853, 1: 357.

37 Massieu F, Sur les Fonctions Caractéristiques des Divers Fluides et Sur la Théorie des Vapeurs, Comptes
Rendus 1869 69: 858–62, 1057–61.

38 Massieu F, Thermodynamique: Mémoire sur les fonctions caractéristiques des divers fluides et sur la théorie
des vapeurs. Académie des Sciences de L'Institut National de France, 1876.

39 Poincaré H, Thermodynamique 1892, Georges Carré, Éditeur, Paris, France.

40 Duhem P, Traité d'Énergétique ou de Thermodynamique générale, t. I et II, 1911, Éditions Jacques Gabay, Paris,
France

41 Meyerson E, Identity and Reality 1964, George Allen & Unwin Ltd, London, UK.

42 Rock PA, Chemical Thermodynamics 1983, University Science Books, CA, USA.

43 Klotz IM, Introduction to Chemical Thermodynamics 1964, W. A. Benjamin, Inc., New York, Amsterdam.

44 Born M, Natural Philosophy of Cause and Chance 1949, Clarendon Press, Oxford, UK.

45 Tisza L, Generalized Thermodynamics 1966, MIT Press, MS, USA.

46 Callen HB, Thermodynamics and an Introduction to Thermostatistics 1985, John Wiley & Sons, NY, USA.

47 Landau LD, Lifshitz EM, Pitaevskii LP, Statistical Physics. Part 1. 1980, Butterworth-Heinemann, UK.

48 Uffink J, Bluff your way in the second law of thermodynamics, Stud. Hist. Phil. Mod. Phys. 2001 32: 305.

49 Daudrich D, Der Zweite Hauptsatz der Thermodynamik und seine Macht 2002,
First-Minute-Taschenbuch-Verlag, Emsdetten, Germany, ISBN 3-932-805-33-X:
http://www.storyal.de/story2005/hauptsatz.htm.

207 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

50 Clausius R. Mechanical Theory of Heat. 9th memoir. 1867, John van Voorst, London, UK.

51 Wolff SL, in: Die großen Physiker. Von Aristoteles bis Kelvin, herausgegeben von Karl von Meyenn 1997. pp.
289-302. Verlag C. H. Beck, München, Germany.

52 Wang J, Modern Thermodynamics 2011, Science Press Beijing, China; Springer Berlin, Heidelberg, Germany.

53 Mahan BH, Elementary Chemical Thermodynamics 1964, W. A. Benjamin, Inc., New York, Amsterdam.

54 Brzustowski ThA, Introduction to Principles of Engineering Thermodynamics 1969, Addison-Wesley


Publishing Company, Inc., Reading, MS, USA.

55 Sears FW, Salinger GL, Thermodynamics, Kinetic Theory, and Statistical Thermodynamics 1998, Narosa
Publishing House PVT. Ltd., New Delhi, India.

56 Chalmers TW, Historic Researches 1949, Morgan Brothers Ltd., London, UK.

57 Kuhn ThS, The Caloric Theory of Adiabatic Compression, ISIS 1961 49: 132.

58 Kuhn ThS, Sadi Carnot and the Cagnard Engine, ISIS 1961 52: 567.

59 Hermanns R, Phlogiston-Theorie: Ein Thema für den Chemieunterricht heute? – Praxis der
Naturwissenschaft – Chemie 1986 35: 28.

60 Freise V, Das Phlogiston: Irrweg oder klassische Theorie? – Praxis der Naturwissenschaft – Chemie 1986 35:
32.

61 Psillos S, Transition from Caloric Theory to Thermodynamics, Stud. Hist. Phil. Mod. Phys. 1994 25: 159.

62 Ehrlichson H, Sadi Carnot, ‘Founder of the Second Law of Thermodynamics’, Eur. J. Phys. 1999 20: 183.

63 Starikov EB, Valid Entropy-Enthalpy Compensation: Its True Physical-Chemical Meaning, J. Appl. Solution
Chem. Model. 2013 2: 240.

64 Starikov EB. Many Faces of Entropy or Bayesian Statistical Mechanics, ChemPhysChem 2010 11: 3387.

65 Starikov EB. George Augustus Linhart – as a “Widely Unknown” Thermodynamicist, World J. Condensed
Matt. Phys. 2012 2: 101.

66 Starikov EB. ‘Entropy is anthropomorphic’: does this lead to interpretational devalorisation of


entropy-enthalpy compensation? Monatshefte für Chemie 2013 144: 97.

208 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

67 Lavenda BH, Statistical Physics: A Probabilistic Approach. 1991 John Wiley & Sons Inc., NY, USA.

68 Atkins P, The Laws of Thermodynamics: A Very Short Introduction 2010, Oxford University Press, UK.

69 Hermann von Helmholtz, Studien zur Statik Monozyklischer Systeme, Berliner Berichte am 6. März 1884.

70 Hermann von Helmholtz, Wissenschaftliche Abhandlungen, Band 3 1895, p. 121, Johann Ambrosius Barth,
Leipzig, Germany.

71 Budde E, Über integrierende Divisoren und Temperatur, Ann. Phys. 1892 281: 751.

72 Planck M, Über die Begründung des zweiten Hauptsatzes der Thermodynamik, Berliner Berichte am 27.
Dezember 1926.

73 Gibbs JW, The Scientific Papers: Thermodynamics 1906, Longmans, Green, London, New York, Bombay.

74 Balluffi RW, Allen SM, Carter WC, Kinetics of Materials 2005, Wiley Interscience Inc., Hoboken, NJ, USA.

75 Tribus M, Thermostatics and Thermodynamics 1961, D. Van Nostrand Inc., Princeton, NJ, USA.

76 Jarzynski C, Non-equilibrium Equality for Free Energy Differences, Phys. Rev. Lett. 1997 78: 2690.

77 Jarzynsky C., Equilibrium Free Energy Differences from Non-equilibrium Measurements: A Master Equation
Approach, Phys. Rev. E 1997 56: 5018.

78 Crooks GE, Non-equilibrium Measurements of Free Energy Differences for Microscopically Reversible
Markovian Systems, J. Stat. Phys. 1998 90: 1481.

79 Leff HS, Thermodynamic Entropy: The Spreading and Sharing of Energy. Am. J. Phys. 1996 64: 1261-1271.

80 Lambert FL, Entropy Is Simple, Qualitatively. J. Chem. Educ. 2002 79: 1241-1246.

81 Lambert FL, Shuffled Decks, and Disorderly Dorm Rooms – Examples of Entropy Increase? Nonsense! J.
Chem. Educ. 1999 76: 1385-1388.

82 Lambert FL, Disorder – A Cracked Crutch for Supporting Entropy Discussions, J. Chem. Educ. 2002 79:
187-192.

83 Jensen WB, Entropy and Constraint of Motion, J. Chem. Educ. 2004 81: 639.

209 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

84 Thomson W, On a Universal Tendency in Nature to the Dissipation of Mechanical Energy, Phil. Mag. 1852 4:
256-260.

85 Steward B, The Conservation of Energy 1874, Appleton: New York, USA.

86 Liveing GN, Chemical Equilibrium the Result of Dissipation of Energy 1885, Deighton, Bell: Cambridge, UK.

87 Nash LK, Elements of Statistical Thermodynamics 1965, Addison-Wesley, New York, USA.

88 Bent H, The Second Law 1965, Oxford Press, New York, USA.

89 Jensen WB, George Downing Liveing and the Early History of Chemical Thermodynamics, Bull. Hist. Chem.
2013 38: 37-51.

90 Maxwell JC, Theory of Heat 2001, Dover Books, Mineola NY, USA.

91 Schmidt E, Einführung in die technische Thermodynamik 1950, Springer-Verlag, Berlin, Göttingen,


Heidelberg, Germany.

92 Schmidt E, Thermodynamics. Principles and Applications to Engineering 1949, At the Carendon Press,
Oxford, Great Britain.

93 Kestin J, A Course in Thermodynamics (2 Volumes, 3rd edition) 1979, Hemisphere Publishing Corporation,
Washington, London.

94 The Second Law of Thermodynamics. Benchmark Papers on Energy, Volume 5. Edited by Joseph Kestin 1976,
Dowden, Hutchinson & Ross Inc., Stroudsburg USA.

95 Kestin J, Dorfman JR, A Course in Statistical Thermodynamics 1971, Academic Press, New York, London.

96 Oono Y, Paniconi M, Steady State Thermodynamics, Prog. Theor. Phys. Suppl. 1998 130: 29-44.

97 Hatano T, Sasa S, Steady-State Thermodynamics of Langevin Systems, Phys. Rev. Lett. 2001 86: 3463-3466.

98 Trepagnier EH, Jarzynski C, Ritort F, Crooks GE, Bustamante CJ, Liphardt J, Experimental test of Hatano
and Sasa’s non-equilibrium steady-state equality, Proc. Natl. Acad. Sci. USA 2004 101: 15038-15041.

99 Onsager L, Reciprocal Relations in Irreversible Processes. I, Phys. Rev. 1931 37: 405-424.

100 Onsager L, Reciprocal Relations in Irreversible Processes. II, Phys. Rev. 1931 38: 2265-2279.

210 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

101 Landsberg PT, Thermodynamics: With Quantum Statistical Illustrations 1961, Wiley Interscience Publishers,
Hoboken, NJ, USA.

102 Landsberg PT, Thermodynamics, Cosmology, and the Physical Constants, in: The Study of Time III 1978 pp.
115-140; Fraser JT, Lawrence N, Park D, Eds.; Springer Science + Business Media, New York, USA.

103 Hon. John William Strutt (Baron Rayleigh), Some General Theorems Relating to Vibrations, Proc. Math. Soc.
London 1873 4: 357-368.

104 Hon. John William Strutt (Baron Rayleigh), The Theory of Sound. Volume I. 1877, McMillan & Co., London,
UK, (§§ 80 - 89).

105 Hon. John William Strutt (Baron Rayleigh), The Theory of Sound. Volume I. Second Edition. Revised and
Enlarged. 1894, McMillan & Co., London, New York, (§§ 80 - 89).

106 Sir William Thomson (Lord Kelvin), Tait PG, Treatise on Natural Philosophy. Volume I. 1867, Cambridge
University Press, Cambridge, UK. (§§ 337-338).

107 The Collected Works of Lars Onsager (with Commentary) 1996, Hemmer PC, Holder H, Ratkje SK, Eds.,
World Scientific, Singapore, New Jersey, London, Hong Kong.

108 Becker R, Theory of Heat 1967, Springer-Verlag. Berlin, Heidelberg, New York. (p. 344).

109 Lvov I, Thermodynamics: Logical Analysis. What is Energy? 2014, LAP Lambert Academic Publishing,
Saarbrücken, Germany.

110 Haken H, Synergetik. Eine Einführung 1990, Springer-Verlag, Berlin, Heidelberg, New York, London, Paris,
Tokyo, Hong Kong.

111 Jeschke G, Mathematik der Selbstorganisation 2009, Verlag Harri Deutsch GmbH, Frankfurt am Main,
Germany.

112 Starikov EB, ‘Meyer-Neldel Rule’: True History of its Development and its Intimate Connection to Classical
Thermodynamics, J. Appl. Sol. Chem. Model. 2014 3: 15-31.

113 Bohr N, Kramers HA, Slater JC, Über die Quantentheorie der Strahlung, Z. f. Phys. 1924 XXIV: 69-87.

114 Bothe W, Geiger H, Über das Wesen des Comptoneffekts; ein experimenteller Beitrag zur Theorie der
Strahlung, Z. f. Phys. 1925 XXXII: 639-663.

115 Ellis CD, Wooster WA, The Average Energy of Disintegration of Radium E, Proc. Roy. Soc. London A 1927

211 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

117: 109-123.

116 Meitner L, Orthmann W, Über eine absolute Bestimmung der Energie der primären -Strahlen von Radium
E, Z. f. Phys. 1930 LX: 143-155.

117 Sargent BW, The Maximum Energy of the -Rays from Uranium X and Other Bodies, Proc. Roy. Soc.
London A 1933 132: 659-673.

118 Ellis CD, Mott NF, Energy Relations in the -Ray Type of Radioactive Disintegration, Proc. Roy. Soc.
London A 1933 141: 109-123.

119 Shankland RS, An Apparent Failure of the Photon Theory of Scattering, Phys. Rev. 1936 49: 8-13.

120 Jacobsen C., Correlation between Scattering and Recoil in the Compton Effect, Nature 1936 138: 25.

121 Bothe W, Maier-Leibnitz H, Eine neue experimentelle Prüfung der Photonenvorstellung, Z. f. Phys. 1936 CII:
143-155.

122 Shankland RS, The Scattering of Gamma-Rays, Phys. Rev. 1936 50: 571.

123 Planck M, Physikalische Gesetzlichkeit, Naturwissenschaft 1926 14: 249-261.

124 Carathéodory C, Untersuchungen über die Grundlagen der Thermodynamik, Math. Ann. 1909 67: 355-386.

125 Carathéodory C, Über die Bestimmung der Energie und der absoluten Themperatur mit Hilfe von reversible
Prozessen, Berliner Berichte am 8. Januar 1925.

126 Born M, Kritische Betrachtungen zur traditionellen Darstellung der Thermodynamik, Phys. Z. 1921 22:
218-224; 249-254; 282-286.

127 Ehrenfest-Afanassiewa T, Zur Axiomatisierung des zweiten Hauptsatzes der Thermodynamik, Z. Phys. 1925
33: 933-945; 1926 34: 638.

128 Ruark AE, The Proof of the Corollary of the Carnot’s Theorem, Phil. Mag. 1925 49: 584-585.

129 Landé A, Axiomatische Begründung der Thermodynamik durch Carathéodory, in: Handbuch der Physik.
Band IX. Theorien der Wärme, Henning F, Ed., 1926, Verlag von Julius Springer, Berlin, Germany.

130 Farkas J, Vereinfachte Ableitung des Carnot-Clausius’schen Satzes, Mathematische und Physikalische Blätter
aus Ungarn 1895 IV: 7-11.

212 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

131 Шиллер Николай Николаевич, Основные законы термодинамики 1902, Типография Императорского
университета Св. Владимира. Акционерного О-ва печатного и издательского дела Н. Т. Корчак-Новицкого,
Киев, Российская Империя. (N. N. Schiller had in fact started to consequently publish his thermodynamic results
in the Russian periodicals from the year 1897 on, but those earlier original publications of him seem to be
inaccessible right away).

132 Martinas K, Brodszki I, Thermodynamics of Gyula Farkas – A New (Old) Approach to Entropy, Per.
Polytech. Ser. Chem. Eng. 2000 44: 17-27.

133 Афанасьева-Эренфест ТА, Необратимость, односторонность и второе начало термодинамики,


Журнал Прикладной Физики 1928 5: 3-30.

134 Lieb EH, Yngvason J, The Physics and Mathematics of the Second Law of Thermodynamics, Phys. Rep. 1999
310: 1-96.

135 Balian R, Entropy, a Protean Concept, Séminaire Poincaré 2003 2: 13-27.

136 Planck M. Vorlesungen über Thermodynamik 1922 Walter de Gruyter, Berlin, Leipzig, Germany.

137 Osgood WF, Advanced Calculus 1925, The McMillan Co., New York, London. (Chapter XVIII, p. 449-…).

138 Blondlot R, Introduction à l’étude de la thermodynamique 1888, Gauthier-Villars, Paris, France; Einführung in
die Thermodynamik, mit Zusätzen und Verbesserungen des Autors 1913, Verlag von Theodor Steinkopff, Dresden und
Leipzig, Germany.

139 Buckingham E, An Outline of the Theory of Thermodynamics 1900, The McMillan Co., New York, London.

140 Herrmann F, Job G, Altlasten der Physik 2002, Aulis Verlag Deubner, Köln, Germany.

141 Tait PG, Heat 1892, McMillan & Co., London, New York.

142 Tait PG, Sketch of Thermodynamics 1877, David Douglas, Edinburgh, UK.

143 Carnot LNM, Principes fondamentaux de l’équilibre et du mouvement 1803, Imprimerie de Crapelet, Paris,
France.

144 Neswald ER, Thermodynamik als kultureller Kampfplatz. Zur Faszinationsgeschichte der Entropie 1850-1915
2006, Rombach Verlag KG, Freiburg i. Br., Berlin, Germany.

145 Coopersmith J, Energy: the Subtle Concept 2010, Oxford University Press, New York, USA.

213 office@penseejournal.com
Pensee Journal Vol 76, No. 6;Jun 2014

146 van der Waals JD, Jr., Über die Erklärung der Naturgesetze auf statistisch-mechanischer Grundlage, Physik.
Zeitschr. 1911 XII: 547-549.

147 Tribus M, McIrvine EC, Energy and Information, Sci. Am. 1971 225: 179-188.

148 Tribus M, Evans RB, The Probability Foundations of Thermodynamics, Appl. Mech. Rev. 1963 16: 765-769.

149 Kline SJ, The Low-Down on Entropy and Interpretive Thermodynamics 1999, DCW Industries, Inc., Palm
Drive CA, USA.

150 Lewis GN, Outlines of a New System of Thermodynamic Chemistry, Proc. Am. Acad. Arts Sci. 1907 43:
259-293.

151 Lewis GN, Randall M, Thermodynamics and the Free Energy of Chemical Substances 1923, McGraw-Hill
Book Company, Inc. New York, London.

152 Lewis GN, The Anatomy of Science 1926, Yale University Press, New Haven, USA, Humphrey Milford,
Oxford University Press, London.

153 Linhart GA, The Relation Between Entropy and Probability. The Integration of the Entropy Equation, J. Am.
Chem. Soc. 1922 44: 140-142.

154 Linhart GA, Correlation of Entropy and Probability, J. Am. Chem. Soc. 1922 44: 1881-1886.

155 Linhart GA, Note on the Absorption of Oxygen by Sheets of Rubber, J. Phys. Chem. 1932 36: 1908-1911.

156 Linhart GA, The Application of the Law of Mathematical Probability to the Behavior of Gases in their
Pressure-Volume-Temperature Relations, J. Phys. Chem. 1932 37: 645-653.

157 Linhart GA, Interpretation of the Pressure-Volume-Temperature Relations of Single and Composite Gases, J.
Phys. Chem. 1933 38: 1091-1097.

158 Linhart GA, Penetration of Solar and Cosmic Rays into Fresh Water Lakes, J. Phys. Chem. 1935 40: 113-119.

214 office@penseejournal.com

Você também pode gostar