0% found this document useful (0 votes)
1 views

44

Uploaded by

JQL contact
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
1 views

44

Uploaded by

JQL contact
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 28

You certainly know the second law of thermodynamics,

Do you know its connection to other laws of physics and


chemistry?
Qiong Ye, Jeremy Cocks, François-Xavier Machu, Qiuping A. Wang

To cite this version:


Qiong Ye, Jeremy Cocks, François-Xavier Machu, Qiuping A. Wang. You certainly know the second
law of thermodynamics, Do you know its connection to other laws of physics and chemistry?. The
European Physical Journal Plus, 2022, 137 (11), pp.1228. �10.1140/epjp/s13360-022-03446-4�. �hal-
03469396v3�

HAL Id: hal-03469396


https://hal.science/hal-03469396v3
Submitted on 8 Nov 2022

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
You certainly know the second law of thermodynamics,

Do you know its connection to other laws of physics and chemistry?

Q. Ye2, J. Cocks1, F.X. Machu1, Q.A. Wang1,2


1
Laboratoire SCIQ, ESIEA, 9 Rue Vésale, 75225 Paris, France
2
IMMM, CNRS UMR 6283, Université du Mans, Ave. O. Messiaen, Le Mans, France

Abstract

Since the discovery of the second law of thermodynamics, skeptics have never ceased to
challenge the absolute validity of the law, leading to a situation where the claim of violation of
the second law has almost become routine today. This situation is also a serious threat to the
fundamental status of thermodynamics. In this work, we bring to light evidences to prove the
absolute validity of the second law as a fundamental law of physics. For this purpose, we
propose a short revisit of the history of the discovery of the second law in order to highlight the
connection between the second law and the first law of energy conservation. We then
demonstrate that the perpetual motion machine of the second kind also violate the first law of
thermodynamics, albeit indirectly, contrary to the common belief. This result confirms the
second law is an inviolable fundamental law of physics, just like the law of energy conservation.
Denying one of these conjoined twin laws is to deny the other. Any presumed violation of the
second law, even a probabilistic one, inevitably violates the laws of energy and mass
conservation, and undermines all fundamental laws of physics and chemistry. We conclude by
summarizing a perspective of the interpretation of the second law taking into account the
multiplicity of paths of random motion.

Keywords: Entropy, second law of thermodynamics, random dynamics, energy, energy


conservation, Carnot engine, perpetual motion machine.

1
1) Introduction

It is beyond any doubt that thermodynamics is among the most important theories of
physics, with very wide applications in almost all domains of natural science and engineering
[1] as well as in social sciences [2]. However, despite the great confidence the scientists and
engineers have in their applications [1][2], serious doubts persist, from the beginning of the
theory until today, about whether thermodynamics is a fundamental physics theory (like for
example classical and quantum mechanics), and whether the thermodynamic laws are exact
rules and universal truth [3][4]. One of the origins of this skepticism is the statistical
interpretation of the behaviors (temperature, pressure etc.) of thermodynamic systems [3][4][5],
especially the statistical interpretation of one of its basic laws : the second law of
thermodynamics (second law for short) [5][6][7]. According to this interpretation, the second
law is only a probabilistic emergence from other more basic laws of mechanics, hence a non-
fundamental rule in the sense that the law is reducible to other laws and, in addition, it does not
always hold [3][4][5][6].

The second law of thermodynamics stipulates that entropy 𝑆 always increases in the
evolution of an isolated system and remains constant when the system reaches thermodynamic
equilibrium. Entropy 𝑆 of a system is defined by its small variation 𝑑𝑆 = 𝛿𝑄/𝑇 [1] when a
small quantity of heat 𝛿𝑄 is absorbed by the system from its surroundings, where 𝑇 is the
absolute temperature of the system. This system which is in an ideal and quasi-equilibrium
undergoes a reversible process [9][10]. For any real (non-ideal) irreversible process, 𝑑𝑆 >
𝛿𝑄/𝑇 [1]. This inequality is the mathematical expression of the second law. When the system
is isolated with 𝛿𝑄 = 0, the second law becomes 𝑑𝑆 > 0 as stipulated above1. According to a
widely accepted interpretation, entropy represents a measure of the uncertainty or disorder in
the movement of the molecules composing any thermodynamic system. This idea is reflected
in Boltzmann's definition of entropy 𝑆 = ln𝑤 [1][11] where 𝑤 is the total number of
microstates (arrangements of the position and momentum of the molecules) in the system. The
greater is 𝑤, the more the molecules are moving in disorder. The second law means that the

1
It is worth mentioning that the entropy is only defined for equilibrium states. It has been extended to non-
equilibrium systems changing sufficiently slowly so as to keep local equilibrium and to apply equilibrium
thermodynamics to each of the sufficiently small partitions [13].

2
system always evolves towards macrostates having a greater number of microstates, i.e.,
towards more and more disorder. A familiar example is the expansion of gas in an isolated
recipient from a part of the container to the whole container, because the number of microstates
in the whole volume is larger than in a part of it. The entropy reaches its maximum when the
gas uniformly fills the volume in an equilibrium state [1].

The statistical interpretation of the second law by Boltzmann [11], a point of view shared
by Maxwell [12], was formulated within classical (Newtonian) mechanics with the help of an
H function depending on the states of the particles in a thermodynamic system, H function
decreases as the system evolves [11][14], contrary to entropy which increases instead. This
formulation encountered serious criticisms, supported by mathematical proofs and irrefutable
arguments based on the time reversibility and recurrent behavior of mechanical motion
[14][15][16]. A reconciliation between the irreversible behavior of entropy in the second law
and the time reversible laws of classical mechanics seems necessary, leading to a statistical
argument stating that the entropy increase is only the consequence of an overwhelmingly large
probability, but the probability of entropy decrease is nevertheless not zero, albeit very small
[11][14]. A thought experiment proposed by Maxwell [17] uses a demonic intelligence able to
follow precisely the movement of all the individual particles in a system, and to sort them
according to their velocities 2 , showing that entropy of the isolated system could indeed
decrease. The pros and cons about this demon continue today from different considerations
based on information [18], quantum property [19], or classical random motion [20]. The
Maxwell demon is still there haunting the second law.

As a matter of fact, the signification of the Maxwell demon goes farther than decreasing
entropy. According to Laplace [21], such a demon would be able to know every detail of the
motion of the particles of a system of any size (even the whole universe) if the particles obey

2
…if we conceive a being whose faculties are so sharpened that he can follow every molecule in its course, such
a being, whose attributes are still as essentially finite as our own, would be able to do what is at present
impossible to us. For we have seen that the molecules in a vessel full of air at uniform temperature are moving
with velocities by no means uniform … Now let us suppose that such a vessel is divided into two portions, A and
B, by a division in which there is a small hole, and that a being, who can see the individual molecules, opens and
closes this hole, so as to allow only the swifter molecules to pass from A to B, and only the slower ones to pass
from B to A. He will thus, without expenditure of work, raise the temperature of B and lower that of A, in
contradiction to the second law of thermodynamics.

3
classical mechanics, making it possible to put in a complete equation of motion the data of the
trajectories of all particles from the past to the future3. He then does not need to observe and to
measure the positions and the velocities of the particles and to open (or not) the door of the hole
to sort the particles. As he knows when and how each particle with given velocity moves to the
hole, he can build an automatic door controlled by a program in which he puts the data of the
motion of each particle. He then just take a coffee waiting for the entropy decrease produced
by the door alone. With this demon, the information consideration [18] can no more avoid the
violation of the second law. Worse, for this demon, there is no disorder at all in the system. No
disorder means no entropy. Actually, he does not see the entropy we are describing with the
second law. This gives to the concept of entropy a subjective character depending on the
perception or mental state of the observer. This demon cannot but deepen the doubt about the
fundamentality of entropy, of the second law, and of the theory of thermodynamics.

In this uncertain context of the second law, the statistical interpretation seems the only
refuge of the second law [5]. It has actually become the mainstream interpretation of entropy
increase, despite the firm belief in its fundamental nature claimed by many (see for example
the citation of Edington4 [22]) and the great confidence in the application of thermodynamics
laws in technology (see for example [23][24] and the references there-in). There was even a
statistical theory formulated around a fluctuation theorem [25], allowing systematic violation
of the second law. In this formulation, the second law violation, despite its small probability,
becomes a ubiquitous phenomenon in every non-equilibrium process [25]. Today it is quite
common to see violation of the second law in the literature, supported either by experimental
evidence [26][27] or by theoretical arguments [6][28][29][30], many reports being based on

3
We may regard the present state of the universe as the effect of its past and the cause of its future. An intellect
which at a certain moment would know all forces that set nature in motion, and all positions of all items of which
nature is composed, if this intellect were also vast enough to submit these data to analysis, it would embrace in a
single formula the movements of the greatest bodies of the universe and those of the tiniest atom; for such an
intellect nothing would be uncertain and the future just like the past would be present before its eyes [21].
4
The law that entropy always increases holds, I think, the supreme position among the laws of Nature. If someone
points out to you that your pet theory of the universe is in disagreement with Maxwell's equations - then so much
the worse for Maxwell's equations. If it is found to be contradicted by observation - well, these experimentalists
do bungle things sometimes. But if your theory is found to be against the Second Law of Thermodynamics I can give you
no hope; there is nothing for it to collapse in deepest humiliation.

4
quantum mechanical consideration [27][28]. We also noticed the publication of some very
radical criticisms such as the second law is fallacious based on flawed reasoning from Carnot’s
theorem [31], the entropy of the second law is not a meaningful physical quantity [32], and
entropy is just a mathematical contrivance without solid physics background [33]. There are
also efforts to construct engines challenging Carnot’s efficiency but keeping the validity of the
second law [34][35][36]. Many authors seem to ignore both the unbreakable connection
between the Carnot’s theorems and the second law, and the dangerous consequence of the
breakdown of Carnot’s theorems.

There are many efforts, from both philosophers [4][6] and physicists [7][14][37], defending
the fundamental status of the second law against the statistical interpretation of the second law,
but their defense is often based on conceptual belief and plausible arguments without
convincing (mathematical) proof. The aim of this work is just to contribute to this effort by
proposing a convincing proof of the fundamental nature of the second law as an inviolable iron
rule. This proof is simply the fact that the second law is tightly connected to the laws of
conservation of energy and mass. This work did not invent this proof which is in fact visible in
some episodes of the discovery of the second law [1][7][38] over almost 40 years through the
work of Carnot [9] and Kelvin [17][39] all the way to Clausius [10]. This is why we propose in
what follows a summary of this history in a short form of trilogy, in order to highlight this
unbreakable mathematical link between the second law and the first law of energy conservation.
This link implies that if the second law is violated, the first law is violated as well, which would
be a disaster of natural science and must be avoided. This conclusion is very important from
the conceptual point of view in order to give back to thermodynamics its unquestionable
fundamentality. It is also crucial from the practical point of view for applying thermodynamics
to the study of the world around us and to the development of energy technology with more
confidence and correct vision.

5
Nomenclature

𝑆 – entropy 𝐸 – energy

𝑑𝑆 – entropy change 𝑤 – number of microstates

𝑄 – heat transferred to a system 𝑊 – mechanical work

∆𝑄 or 𝛿𝑄 – small heat transfer 𝑈 – internal energy

𝑅 – heat reservoir 𝜃 – temperature of arbitrary scale

 – efficiency of heat engine 𝑇 – absolute temperature

PMM – perpetual motion machine PMM1 – perpetual motion machine of first kind

PMM2 – perpetual motion machine of second kind

2) Discovery of entropy

It is sometimes said that the second law is an empirical finding5. This is not exactly true.
The discovery of the second law, in its most profound form in terms of entropy, is actually a
result of pure thought experiments and mathematical calculations.

Historically, the discovery of entropy has something to do with a decree of the Academy of
Science of Paris in 1775 declaring they would not examine any proposition concerning
perpetual motion machines [40]. The decree concerned what we call today perpetual motion
machines of the first kind (PMM1)6. Since then, the impossibility of PMM1 began to be taken
for granted on the basis of a common belief that energy and motion cannot be created from

5
https://en.wikipedia.org/wiki/Second_law_of_thermodynamics
6
A dream of many engineers of that time was to construct what is called today the perpetual motion machine of
the first kind, a machine that can produce an unlimited amount of work without consuming heat or other energy.
This machine violates the law of energy conservation. There is another type of perpetual motion machine called
the second kind capable of transforming heat from a single reservoir completely into work. It is straightforward to
show that this engine makes heat flow from a cold to a hot body, hence violate the second law. We will show later
in this paper that this machine turns out to violate the first law of energy conservation as well because we can use
it to fabricate a perpetual motion engine of the first kind.

6
nothing [9]; the law of energy conservation was not yet known at that time. Today we have the
mathematical proof that this machine cannot exist because it violates the first law of
thermodynamics, the law of conservation of energy. This decree is the starting point of the
remarkable work of Carnot in the study of the efficiency of thermal engine [9], leading to the
second law many years later. We can at once guess that there must have some vital connection
between the second law and its starting hypothesis. In order to highlight this connection and the
mathematical rigor in it, we will retrace the impressive trilogy from the impossibility of
perpetual motion all the way to the discovery of the second law. The reader will see a summary
of the key elements of the story below. The details of the mathematics are described in
Appendix.

a) From perpetual motion machines to Carnot theorem

Carnot was concerned with heat engines and their efficiency. In 1824, he published the
extraordinary idea of a reversible heat engine7, now called a Carnot engine [1]. This ideal engine
can work in a reversible cycle between two heat reservoirs (R1 and R2), one (R1) at higher
temperature 𝜃1 and the other (R2) at lower temperature 𝜃2 . The engine can work in a forward
cycle, extracting heat 𝑄1 from 𝜃1 , emitting heat 𝑄2 (<𝑄1 ) to 𝜃2 and producing a mechanical
work 𝑊 = 𝑄1 − 𝑄2 ; it can also reverse the cycle, extracting heat 𝑄2 from R2, absorbing work
W, and emitting heat 𝑄1 to R1. The Carnot engine works in both directions with the same
𝑊 𝑄1 −𝑄2 𝑄
efficiency defined by  = 𝑄 = = 1 − 𝑄2 [1]. This is a purely ideal machine which cannot
1 𝑄1 1

exist in reality. It is impossible for a real machine to perform such a reversible operation due to
friction or other form of energy loss. For this reason all real engines are irreversible machines.
By examining this ideal machine, Carnot discovered the efficiency limit for all possible engines;
which he described this in the first part of his theorem, asserting that8:

All irreversible heat engines between the two reservoirs are less efficient than the reversible
engine working between the same reservoirs.

7
Ref [9], pp17-20
8
The maximum motive power achieved by the use of vapor is also the maximum motive power achievable by
any other means. Ref [9], p22
7
Carnot proved this theorem by reductio ad absurdum based on the impossibility of PMM19.
The details of his proof are illustrated in Appendix I. To summarize, this theorem is true if and
only if the PMM1 does not exist. In other words, if an irreversible engine working between the
same reservoirs is more efficient than the Carnot engine, we can then use this irreversible engine
to construct a PMM1 (Figure A1 of Appendix I) to produce energy with no need of energy
source. This engine would solve our big worry about renewable energy source and
environment10.

We will see below that the meaning of the existence of this PMM1 goes much deeper than
simply giving a solution to our energy concerns. This first part of Carnot’s theorem was crucial
for the discovery of the second law. However, the second law would be impossible without the
second part of Carnot’s theorem.

The second part of Carnot’s theorem is in some way implicit in the first part. Since the
Carnot engine is the most efficient of all possible engines, all Carnot engines, each working
with a different substance (air, steam, etc.), must be equally efficient. If two Carnot engines
were not equally efficient, we would be able to replace the first irreversible engine in Carnot's
proof (Appendix I) by this more efficient Carnot engine, in order to make a PMM1. All Carnot
engines, therefore, have the same efficiency regardless of their working substances. This
efficiency, which we call , can only depend on the temperatures 𝜃1 and 𝜃2 of the reservoirs
R1 and R2, respectively. The second part of the theorem states [9]:

All Carnot engines have the same efficiency regardless of their working substances; this
efficiency only depends on the temperatures of the reservoirs: (𝜃1 , 𝜃2 ).

For Carnot, the only interest of his magic engine is to specify the limit of the efficiency of
all possible heat engines. A practical question is how to calculate this limit (𝜃1 , 𝜃2 ) when the
temperatures of the reservoirs are given? Physicists had to wait about thirty years to find the
answer, which was given by Kelvin [39].

9
Ref [9]. pp21-22
10
Ref [9], p22

8
b) Kelvin's efficiency of Carnot’s engine

In 1854, Kelvin found the expression of (𝜃1 , 𝜃2 ) by considering the second part of the
Carnot theorem and using a new scale of temperature called absolute temperature denoted by
𝑇 [39]. The expression is [1]

𝑇2 (1)
= 1−
𝑇1

for a Carnot engine working between two reservoirs of absolute temperature 𝑇1 and 𝑇2 (𝑇1 >
𝑇2 ). This expression can be found in a general way using a combination of two Carnot engines
in series, the first one working between the reservoir R1 at 𝑇1 and the reservoir R2 at 𝑇2 (𝑇1 >
𝑇2 ), the second one working between the reservoir R2 and the reservoir R3 at 𝑇3 (𝑇2 > 𝑇3). The
mathematics is straightforward as shown in Appendix II. Kelvin’s equation (1) is an
unavoidable consequence of the second part of the Carnot theorem that the efficiency only
depends on the temperatures of the two reservoirs.

Now let ∆𝑄1 be the heat absorbed by the Carnot engine from R1, and ∆𝑄2 the heat it
“absorbs” from R2 (it actually emits −∆𝑄2 to R2). After a complete cycle, the engine returns
to R1. The work done during the cycle is given by 𝑊 = ∆𝑄1 −(−∆𝑄2 ). It is obvious that the
𝑊 (−∆𝑄2 ) 𝑇 ∆𝑄1 ∆𝑄2
efficiency  = ∆𝑄 = 1 − = 1 − 𝑇2 , which gives us + = 0. Suppose now we
1 ∆𝑄1 1 𝑇1 𝑇2

increase the number of reservoirs to N, where the ith reservoir (𝑖 = 1,2 … 𝑁) at temperature 𝑇𝑖
gives 𝛿𝑄𝑖 to the engine. Kelvin obtained the following equation [1]
𝑁 𝛿𝑄𝑖 (2)
∑ =0
𝑖=1 𝑇𝑖

But he stopped here in his quest for the second law without reaching the concept of entropy
[7][39].

c) Clausius' perception

The discovery of what is called the second law of increasing entropy finally came to
Clausius who, like Kelvin, also derived Eq.(2) [10]. But he saw something deeper in it. If the
𝛿𝑄
number of reservoirs is very large, Eq.(2) takes the form of a closed line integral : ∮ = 0.
𝑇
𝛿𝑄
This integral over a reversible cycle implies that the quantity represents the variation of a
𝑇

variable 𝑆 which is a state variable. During contact with a reservoir at temperature T, the engine

9
𝛿𝑄
receives heat 𝛿𝑄 as well as undergoing a variation of S: 𝑑𝑆 = . After a complete cycle, the
𝑇
𝛿𝑄
engine comes back to its initial state, and the total variation of 𝑆 is zero: ∮ 𝑑𝑆 = ∮ = 0 [1].
𝑇

𝛿𝑄
Clausius didn't stop here with 𝑑𝑆 = for a reversible process. He had the brilliant idea of
𝑇

using the same analysis as above for a real but irreversible engine, of efficiency ′, cycling
between the same reservoirs R1 and R2, absorbing ∆𝑄1 and ∆𝑄2, and doing work 𝑊′ during a
𝑊′ (−∆𝑄2 )
cycle. We have ′ = ∆𝑄 = 1 − . Then he used the first part of Carnot’s theorem which
1 ∆𝑄1

states that the efficiency of an irreversible engine is less than that of a reversible engine i.e.
𝑇
′ < , and using Kelvin's limit  = 1 − 𝑇2, he wrote
1

∆𝑄2 𝑇2 (3)
1+ <1−
∆𝑄1 𝑇1
∆𝑄1 ∆𝑄2 𝛿𝑄
which means + < 0. For a large number of reservoirs on the cycle, he wrote ∮ < 0,
𝑇1 𝑇2 𝑇

which yields (see Appendix III) [1][10]:

𝛿𝑄 (4)
𝑑𝑆 ≥
𝑇

for both irreversible (>) and reversible (=) processes. Clausius coined the term entropy for this
variable S because he wanted to find a Greek word that is parallel and analogous to the Greek
word energy11 [41]. He certainly felt some intrinsic link between energy and entropy, himself
having just at that time proposed the first law of thermodynamics about energy conservation
[42].

3) If we backtrack the trilogy

By summarizing the trilogy of discovery of the second law, we wanted also highlight the
infallible mathematics going from the impossibility of perpetual motion machine all the way to
the second law. These rather straightforward mathematical deductions show that, if perpetual

11
I prefer going to the ancient languages for the names of important scientific quantities, so that they may mean
the same thing in all living tongues. I propose, accordingly, to callthe entropy of a body, after the Greek word
'transformation.' I have designedly coined the word entropy to be similar to 'energy,' for these two quantities are
so analogous in their physical significance, that an analogy of denomination seemed to me helpful [41].

10
machine is impossible, then there must be the Carnot theorem, and then the second law follows;
you cannot get other thing different. And inversely, if the second law is broken, i.e., Eq.(4)
breaks down, you can backtrack all the mathematical calculations of Clausius from Eq.(2) to
Eq.(4), to prove Eq.(3) untrue. Eq.(3) stems from Eq.(1) and the first part of Carnot theorem.
Its falseness implies either the falseness of Eq.(1) or that of the first part of Carnot theorem.

Now suppose Eq.(1) is false, in view of the infallible mathematics of Kelvin using Figure
A2 (Appendix II), the second part of Carnot theorem must be untrue. In other words, different
Carnot engines using different working substances may have different efficiencies. According
to the discussion in the section 3a), we will be able to construct a PMM1 using two Carnot
engines of different efficiencies.

If we suppose Eq.(1) and the second part of the Carnot theorem are true but the first part of
the theorem is false, so there may be engines more efficient than Carnot engine. Then we can
use these more efficient engines, as shown in Figure A1 of Appendix I, to fabricate PMM1,
which will give us infinite quantity of clean energy we need to preserve our pet planet.

Up until now, we have only talked about PMM1. The reader may wonder what would
happen if the perpetual motion machine of the second kind (PMM2) is involved. PMM2 is a
machine transforming heat from a single heat reservoir into mechanical work with no other
effects. Unlike PMM1, PMM2 does not violate the first law [1], because it transforms a heat Q
into mechanical work W of same quantity of energy (W=Q). It only violates the Kelvin-Planck
expression of the second law [17][39] stipulating, simply, that the perpetual machine PMM2 is
impossible. This machine is widely regarded as an example of the violation of the second law
without violating the first law of energy conservation [1]. In what follows, we will prove that
PMM2, despite its conservation of energy in a cycle, indirectly violates the first law of energy
conservation through its violation of the second law. For this purpose, we will show first how
it violates the second law in allowing heat flow from colder body to hotter body.

As shown in Figure A4 of Appendix IV, we can combine a PMM2 pumping heat 𝑄3 from
a colder reservoir R2 at 𝑇2 and does a work 𝑊3 = 𝑄3 , with a Carnot engine which uses 𝑊3 in
order to pump a heat 𝑄2 from the reservoir R2 and to give the 𝑄1 = 𝑄2 + 𝑄3 to a hotter
reservoir R1 at 𝑇1 (𝑇1 > 𝑇2 ). In this cycle, the hotter reservoir R1 gets the heat 𝑄1 from the
colder reservoir R2 without leaving any side effect, leading to the total entropy variation ∆𝑆 =
𝑄1 𝑄
− 𝑇1 < 0, a violation of the Clausius's expression of the second law ∆𝑆 ≥ 0 for the isolated
𝑇1 2

11
system composed of the two engines and the two reservoirs. In this case, we can construct an
engine with an efficiency ′ larger than the efficiency of the Carnot engine , i.e., ′ >  =
𝑇
1 − 𝑇2. This engine, violating the first part of the Carnot theorem, can be of course used in
1

Figure A1 of Appendix I to construct PMM1 (see Appendix IV, Figure A5) which violates the
first law of energy conservation. To summarize, the violation of the second law by PMM2
necessarily entails the existence of PMM1 providing free energy. Well, PMM1 and free source
of energy come into being! So much the better, granted! So what!

4) So what if the second law is violated?

The most important implication of the violation of second law is not the existence of
perpetual motion machine (PMM) as a source of infinite energy. There is another much deeper
meaning behind the PMM : energy becomes something you can create or annihilate as much as
you please. It loses at once the attribute of a conservative fundamental quantity of motion,
regardless of its form, thermal, mechanical, electromagnetic, gravitational, and so on. The first
mathematical consequence is the collapse of the first law of thermodynamics 𝑑𝑈 = 𝛿𝑄 + 𝑑𝑊.
Why? Because after a cycle of the PMM in the Figure A1 of appendix I for instance, 𝑑𝑈 =
𝛿𝑄 = 0, but 𝑑𝑊 ≠ 0; this is the work done by the perpetual engine without consuming heat.
Needless to say, the whole theoretical edifice of the thermodynamics will collapse.

This is not all. As well known, energy conservation is not only one of the cornerstone of
thermodynamics, it is also a key pillar of the whole edifice of physical and chemical science,
as well as a fundamental rule guaranteeing the stability of all natural systems. For example, our
pet planet is in stable revolution around the sun because its total mechanical energy remains
constant (conserved) when the pair earth-sun is a good isolated system, at least for a period
sufficiently long for our intelligence to develop, fortunately. Now if the energy of the isolated
earth-sun was no longer a constant and can arbitrarily change (maybe due to a PMM in the
system modifying arbitrarily the energy of the earth without changing anything else), the earth
would have fallen in the sun or flight to the border of Milky Way since long. Please make the
same reasoning for nuclei, atoms, molecules and any energy conservative system in Nature; do
you think that they can be there as stable components of our universe? Imagine also that you
want to make a solid state synthesis; you put the prepared sample in the oven and increase the
temperature to 1000 K. But one day you realize that the sample is freezing! Why? Because the
heat is flowing from the colder sample into the hotter oven, a direct consequence of the violation

12
of the second law and of the energy conservation! Moreover, from a fundamental point of view
taking into account the Einstein’s formula for energy-mass equivalence12, if energy loses its
attribute of a conservative quantity, the mass can no longer be a conserved quantity, so that the
mass conservation is no longer a fundamental law of nature, at least conceptually. It is well
known that most, if not all, fundamental equations in physics and chemistry are based on the
energy and mass conservation. If energy and mass were no more conserved quantities, we
would see the disappearance of Newtonian equation, Maxwell’s equations, chemical reaction
equations, stoichiometry formula, the laws of diffusion of matters and heat, Einstein's equation
of general relativity, and, finally, Schrödinger equation! After these losses, what does remain
in physics and chemistry?

Indeed, when the violation of the second law, together with the time reversibility, is
regarded as isolated phenomenon, it looks like something plausible and interesting even
exciting to investigate. However, when this violation is coupled to the violation of the laws of
energy and mass conservation, we quickly understand to what extent the second law of
thermodynamics is an inviolable iron rule. Can we imagine a violation of energy and mass
conservation, even probabilistic with infinitesimal likelihood, as the fluctuation theorem stated
for second law violation [25][26]? Can we imagine that energy and mass are no more
conservative quantities? Simply we cannot! That would be a fatal menace for the whole edifice
of physics and chemistry; the whole nature would collapse into nil even if the violation of the
second law and of energy-mass conservation took place during only a fraction of a second.

If one suggests the violation of the second law, he should first explain how and why both
energy and mass become non-conservative quantities, in both classical and quantum systems!
He must also provide an explanation why and how the universe and the world around us can
continue to exist while all the physics and chemistry laws have disappeared.

Now we understand better the comment of Edington about the absolute validity of the
second law (footnote p.2) [22].

12
In special relativity theory, we have the famous formula 𝐸 = 𝑚𝑐 2 , where 𝐸 is the energy of a matter, 𝑚 its mass,
and 𝑐 the velocity of light in vacuum.

13
5) Concluding remarks and perspectives

The objective of this work is to provide a proof of the fundamental nature of the second law
as an inviolable iron rule. We have first reviewed the uncertain situation concerning the
fundamentality of the second law and thermodynamics in general, in order to highlight that the
statistical interpretation of the second law based on classical mechanics is the main origin of
the impressive confusion in the understanding of the second law and the concept of entropy. In
order to reach our objective, we have first summarize the history of the discovery of the second
law by Carnot, Kelvin and Clausius in three episodes, to highlight an unbreakable connection
between the second law and the first law of energy conservation. We have then proved that, the
perpetual motion machine of the second kind, regarded most textbooks [1] as a process violating
the second law but not the first law, also necessarily violate the first law of energy conservation,
albeit indirectly. This result confirms the unbreakable link between the two laws.

Now we see that entropy increase is not an isolated phenomenon on its own in nature. It is
an unavoidable consequence of the conservation of energy. The violation of the second law,
even a statistical one, necessarily leads to the fatal failure of the laws of energy and mass
conservation, and to the conceptual breakdown of the whole physics and chemistry science. In
view of the incontestable universal validity of energy and mass conservation, it can be
concluded that any claim of violation of the second law, be it classical or quantum, is necessarily
questionable. If a theory, well confirmed both theoretically and empirically in its domain, leads
to the violation of the second law, the only plausible explanation is either the theory is
incorrectly applied to thermodynamics, or the "entropy" violated is simply not the entropy of
the second law of thermodynamics.

To our eyes, this conclusion is important from the conceptual viewpoint for
thermodynamics to be regarded as one of the most fundamental theory of physics. It is also
important from practical viewpoint for scientists and engineers to be more confident in the
application of thermodynamics to the understanding of the world at different scales up to the
universe, to the prediction of its future evolution, and to the development of technology
especially for the production of sustainable energy [23][24][34][35][36].

It is worth noticing that this conclusion emphasizes once more the discordance between the
second law of thermodynamics and its statistical interpretation based on classical mechanics
[1][11][12]. Although the statistical interpretation is widely accepted by both the philosophy

14
and physics community [4][5], it is always challenged by the irrefutable criticisms [5][14] based
on the Liouville theorem [43], the Poincare recurrence theorem [15] and the time symmetry of
classical mechanical motion [43][44]. The first law – second law connection revealed in this
work with mathematical proof constitutes another hard obstacle on the pathway from classical
mechanics to the second law.

One of the possible interpretation of the second law is to consider the quantum mechanical
behavior of the particles of thermodynamic systems [19][45]. But this approach has its own
uncertainty because, as mentioned above, there are more and more propositions of violation of
the second law in quantum regime [27][28][29]. The last word needs further investigation.
Anyway, quantum interpretation is not all because a great number of systems around us obeying
the second law are not quantum. In our opinion, it is of first necessity to interpret the second
law within classical mechanics before addressing its quantum origin. We would like to mention
here a candidate formalism in this direction. This formulism considers an essential difference
between the classical mechanics motion, which is regular, deterministic, and characterized by
the uniqueness of path of each particle [43], and the thermodynamic motion, which is random,
indeterministic, and characterized by the multiplicity of paths of each particle between an initial
state and a final one. For this kind of motions, the fundamental principle of least action and the
equations of motion, working only for single path motion [43], must be extended in such a way
to include many paths and their probability to occur. Although many results have been achieved
within this multiple path approach through either informational consideration [46][47] or
generalization of the basic principle of classical mechanics [48], further effort is necessary to
be able to finally formulate complete interpretation of the second law [49][50].

Data availability statement


The data that support the findings of this study are available from the corresponding author
upon reasonable request.

15
References
[1] I. Müller, W.H. Müller, Fundamentals of thermodynamics and applications, Springer-
Verlag, Berlin, 2009
[2] V.N. Pokrovskii, Thermodynamics of Complex Systems: Principles and applications,
IOP Publishing, Bristol, 2020
[3] C. Callender, Taking thermodynamics too seriously, Stud. Hist. Phil. Mod. Phys., 32
(2001) 539
[4] M. te Vrugt, P. Needham, G. J. Schmitz, Is thermodynamics fundamental?
arXiv:2204.04352v1
[5] J. L. Lebowitz, Boltzmann's entropy and time's arrow, Physics Report, 46 (1993) 32
[6] T. Müller, The reversibility objection against the Second Law of thermodynamics
viewed, and avoided, from a logical point of view, Stud. Hist. Phil. Mod. Phys., 66 (2019)
52
[7] J. Uffink, Bluff your way in the second law of thermodynamics, Stud. Hist. Phil. Mod.
Phys., 32 (2001) 305–394
[8] D.Z. Albert, Time and chance, Haward University Press, Cambridge, 2000
[9] S. Carnot, Réflexion sur la puissance motrice du feu et sur les machines propres à
développer cette puissance (Reflection on the motive power of fire), Chez Bachelier
Libraire, Paris (1824)
https://archive.org/details/reflectionsonmot00carnrich/page/n7/mode/2up?view=theater
[10] R. Clausius, Über eine veränderte Form des zweiten Hauptsatzes der mechanischen
Wärmetheorie. Annalen der Physik. xciii 1854; 12, 481–506,
doi:10.1002/andp.18541691202, Retrieved 24 March 2014. Translated into
English: Clausius, R. (July 1856). On a Modified Form of the Second Fundamental
Theorem in the Mechanical Theory of Heat. London, Edinburgh, and Dublin
Philosophical Magazine and Journal of Science. 4th. 2: 86. Retrieved 24 March 2014.
Reprinted in: Clausius, R. (1867). The Mechanical Theory of Heat – with its
Applications to the Steam Engine and to Physical Properties of Bodies. London: John
van Voorst. Retrieved 19 June 2012. editions: PwR_Sbkwa8IC.
[11] L. Boltzmann, Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen.
Sitzungsberichte Akademie der Wissenschaften 1872; 66, 275-370. English
translation: L. Boltzmann, Further Studies on the Thermal Equilibrium of Gas

16
Molecules. The Kinetic Theory of Gases. History of Modern Physical Sciences. 1.
pp. 262–349.
[12] J.C. Maxwell, Theory of Heat. London, New York, Longmans, Green (1871); New
York: Dover, (2001) ISBN 0-486-41735-2
[13] De Groot, S.R., Mazur, P. Non-equilibrium Thermodynamics, North-Holland,
Amsterdam (1962).
[14] T.Y. Wu, Boltzmann's H theorem and the Loschmidt and the Zermelo paradoxes,
International Journal of Theoretical Physics, 14 (1975) 289, doi:10.1007/BF01807856.
[15] H. Poincaré, Sur le problème des trois corps et les équations de la dynamique (On the
problem of three bodies and the equations of the dynamics), Acta Math. 13 (1890) 1
[16] H. Poincaré, Thermodynamique, Gautier-Villars, Paris, 1908, p.450
[17] W. Thomson, Kinetic theory of the dissipation of energy, Nature. 9 (1874) 441–444;
and The sorting demon of Maxwell, Nature, 20 (1879) 126
[18] C.H. Benette, demons, engines, and the second law, Scientific American, 1987, p.108
[19] K. Robertson, The demons haunting thermodynamics, Physics Today, 74 (2021) 44
[20] Sophia Chen, A reboot of the Maxwell demon thought experiment – in real life,
WIRED, Science, 14th October 2022,
[21] Pierre-Simon Laplace, Essai philosophique sur les probabilités (Philosophical essay on
the probability), John Wiley & Sons, New York (1902)
[22] A.S. Eddington, The Nature of the Physical World, Cambridge University Press, 1928,
p.74
[23] B. Mei, P. Barnoon, D. Toghraie, C.H. Su, H.C. Nguyen, A. Khan, Energy, exergy,
environmental and economic analyzes (4E) and multi-objective optimization of a PEM
fuel cell equipped with coolant channels, Renewable and Sustainable Energy Reviews,
157 (2022) 112021
[24] P. Barnoon, M. Ashkiyan , D. Toghraie, Embedding multiple conical vanes inside a
circular porous channel filled by two-phase nanofluid to improve thermal performance
considering entropy generation, International Communications in Heat and Mass
Transfer, 124 (2021) 105209
[25] D. J. Evans, D. J. Searles, Advances in Physics, 51 (2002) 1529-1585

17
[26] Wang, G. M., Sevick, E. M., Mittag, E., Searles, D.J. & Evans, D.J. Experimental
Demonstration of Violations of the Second Law of Thermodynamics for Small Systems
and Short Time Scales. Physical Review Letters, 89 (2002) 050601
[27] P. A. Camati, J. P. S. Peterson, T. B. Batalhão, K. Micadei, A. M. Souza, R. S. Sarthour,
I. S. Oliveira, and R. M. Serra, Experimental Rectification of Entropy Production by
Maxwell’s Demon in a Quantum System, Phys. Rev. Lett. 117 (2016) 240502
[28] A Rivas, M.A. Martin-Delgado, Topological Heat Transport and Symmetry-Protected
Boson Currents, Scientific Reports, 7 (2017) 6350. See also : An apparent macroscopic
violation of the second law of thermodynamics in a quantum system is discovered,
https://phys.org/news/2017-09-apparent-macroscopic-violation-law-
thermodynamics.html
[29] M. Kostic, Challenges to the second law challengers, Thermodynamics Symposium,
2016; AAAS Pacific Division 97th Annual Meeting, University of San Diego, San
Diego, CA, June 14-17, 2016.
See https://sites.google.com/site/professorkostic/the-second-law-of-
thermodynamics/challanges-to-the-second-law-challangers and the references there-in
[30] See https://www.eoht.info/page/Violations%20of%20the%20second%20law, and
references there-in.
[31] J. Egashira, Thermodynamics criticism: Carnot cycle judged from physics, Physics
Essays, 23 (2010) 569
[32] S.F. Zhang, Entropy: A concept that is not a physics quantity, Physics Essays, 25
(2012) 172
[33] K. Mayhew, Entropy: An ill-conceived mathematical contrivance? Physics Essays, 28
(2015) 352
[34] E. Panarella, Experimental proof-of-principle of heat recovery and recirculation in a
reciprocating steam engine. Applicability of the technology to present electricity
generating power plants and estimation of the yearly world energy saving and reduction
of greenhouse gas emission, Phys. Essays, 35 (2022) 115
[35] M. de Peretti, A. P. Olson and E. Panarella, Energy breakeven in thermonuclear fusion.
An advanced concept for efficient energy input with positive implications for energy
saving and the geoscience of climate change, Physics Essays 34 (2021) 4

18
[36] Sungguen Ryu, Rosa López, Llorenç Serra, David Sánchez, Beating Carnot efficiency
with periodically driven chiral conductors, Nature Communication, 13 (2022) 2512
[37] M.M. Kostic, The elusive nature of entropy and its physical meanings, Entropy, 16
(2014) 953; see also challanges-to-the-second-law-challangers
[38] W.M. Saslow, A history of thermodynamics, a missing manual, Entropy, 22 (2020) 77
[39] W. Thomson, Mathematical and Physical Papers, Vol. I, Cambridge: Cambridge
University Press (1882), pp.232-291
[40] Académie royale des sciences, Histoire de l'Académie royale des sciences : suivi
des Mémoires de mathématique et de physique tirés des registres de l'Académie royale
des sciences, 1775, Paris, Imprimerie royale, 1778, 1 vol., 66 p. et 575 p., in-4o
[41] C. C. Gillispie, The Edge of Objectivity: An Essay in the History of Scientific Ideas,
Princeton University Press. 1960, p. 399
[42] R. Clausius, Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus
für die Wärmelehre selbst ableiten lassen, Annalen der Physik, 1850; 79, 368–397, 500–
524; English Translation: On the Moving Force of Heat, and the Laws regarding the
Nature of Heat itself which are deducible therefrom. Phil. Mag. 2 (1851) 1–21, 102–119
[43] V.I. Arnol'd, Mathematical methods of classical mechanics, Springer (1978)
[44] Q.A. Wang, Q. Ye, No mysterious motor driving time forward - Multiple paths of random
motion toward time arrow, preprint
[45] P. Ball, Physicists rewrite the fundamental law that leads to disorder, Quanta Magazine,
May 26 (2022) 1
[46] E.T. Jaynes, The minimum entropy production principle, Ann. Rev. Phys. Chem., 31
(1980) 579
[47] K. Ghosh, L. Agozzino, P.D. Dixit, K.A. Dill, The maximum caliber principle for
nonequilibria, Ann. Rev. Phys. Chem., 71 (2020) 10.1
[48] T. Lin, R. Wang, W. P. Bi, A. El Kaabouchi, C. Pujos, F. Calvayrac, Q. A. Wang, Path
probability distribution of stochastic motion of non-dissipative systems: a classical
analog of Feynman factor of path integral, Chaos, Solitons and Fractals, 57 (2013) 129,
arXiv:1310.0411
[49] S. Davis and S. Gonzales, Hamiltonian formalism and path entropy maximization, J.
Phys. A: Math. Theor., 48 (2015) 425003

19
[50] Q.A. Wang, A. El Kaabouchi, From Random Motion of Hamiltonian Systems to
Boltzmann H Theorem and Second Law of Thermodynamics -- a Pathway by Path
Probability, Entropy, 16 (2014) 885, arXiv:2010.07697

20
Appendix
This is a summary of the history of the discovery of the second law of thermodynamics over
thirty years starting from 1824.

I) Carnot theorem from a thought experiment

Figure A1 illustrates a combination of an irreversible real engine and a reversible Carnot


engine, both working between a hot reservoir R1 at temperature 𝜃1 and a cold reservoir R2 at
temperature 𝜃2 . The irreversible engine produces a work 𝑊 ′ = 𝑄1 − 𝑄2 in absorbing a heat 𝑄1
𝑊′
from the hot R1 and emitting a heat 𝑄2 to R2. Its efficiency is given by ′ = [1]. The
𝑄1

reversible engine is working in the inverse direction using a work 𝑊 to absorb the heat 𝑄2 from
𝑊
R2 and give a heat 𝑄1 to R1. Its efficiency is given by  = 𝑄 . After a cycle, the two reservoirs
1

recover their initial states respectively because they didn't obtain nor lose heat.

Figure A1: Illustration of a PMM1 (red line) composed of a real irreversible engine having
𝑊′ 𝑊 𝑄
efficiency ′ = and a Carnot engine having the efficiency  = 𝑄 = 1 − 𝑄2. If ′ > , the
𝑄1 1 1

PMM1 is able to produce the work ∆𝑊 = 𝑊 ′ − 𝑊 after each cycle without consuming heat,
so that the energy ∆𝑊 is created from nothing.

21
Now Carnot supposed that the irreversible engine has a larger efficiency than the reversible
𝑊′ 𝑊
engine, ′ > , or > 𝑄 , leading to 𝑊′ > 𝑊. In this case, the reversible engine can use a part
𝑄1 1

of 𝑊′ to produce the work 𝑊, so that after a cycle, the only thing changed is the extra work
∆𝑊 = 𝑊 ′ − 𝑊 > 0 produced by the ensemble of two engines (red line in Figure A1). This
means that the ensemble of two engines is a perpetual motion machine of the first kind (PMM1)
producing the work ∆𝑊 from nothing, which Carnot thought impossible according to a
common belief that mechanical work and motion cannot be created from nothing (a rudimental
idea of energy conservation). Conclusion: ′ can never be larger than ; which is the first part
of Carnot theorem ′ ≤  [9].

From the same reasoning, we can prove that all reversible engines must have the same
efficiency regardless of their different working substances. This is because if a reversible engine
was more efficient than another one, we could combine them in the same way as in Figure A1
with the more efficient Carnot engine replacing the irreversible one; this would make a PMM1
(absurd). Conclusion: all Carnot engines must have the same efficiency depending only on the
temperatures of the reservoirs 𝜃1 and 𝜃2 , which is the second part of Carnot theorem [9].

II) Kelvin's efficiency

It is quite logical that the efficiency  of Carnot engine depends only on the temperature of
the reservoirs. But what is the form of the function (𝜃1 , 𝜃2 )? Kelvin is concerned with this
question, and made a calculation [39] which can be illustrated by the connection of two Carnot
engines as shown in Figure A2. The first engine works between two reservoirs R1 at 𝜃1 and R2
at 𝜃2 (𝜃1 > 𝜃2 ), absorbing 𝑄1 from R1, emitting 𝑄2 to R2 and doing a work 𝑊1 = 𝑄1 − 𝑄2 ;
𝑄
the efficiency is 1 = 1 − 𝑄2. The second engine works between the above reservoir R2 and a
1

third reservoir R3 at 𝜃3 (𝜃2 > 𝜃3 ), absorbing 𝑄2 from R2, emitting 𝑄3 to R3, and doing a work
𝑄
𝑊2 = 𝑄2 − 𝑄3; its efficiency is 2 = 1 − 𝑄3. On the other hand, these two engines connected
2

in series can be considered as a third engine, working between the hot reservoir R1 at 𝜃1 and
the cold reservoir R3 at 𝜃3 , absorbing 𝑄1 from R1, emitting 𝑄3 to R3, and doing a work 𝑊3 =
𝑊3 𝑄
𝑊1 + 𝑊2 = 𝑄1 − 𝑄3 ; its efficiency is 3 = = 1 − 𝑄3.
𝑄1 1

22
Figure A2: Illustration of Kelvin's idea to connect two Carnot engines in series to compose a
third Carnot engine (red line) working between 𝜃1 and 𝜃3 .

𝑄 𝑄
The mathematics is the following: from the three efficiencies 1 − 1 = 𝑄2 , 1 − 2 = 𝑄3
1 2

𝑄
and 1 − 3 = 𝑄3 , it is straightforward to write (1 − 1 )(1 − 2 ) = 1 − 3 . As all Carnot
1

engines have the same efficiency depending only on the temperature of the reservoirs, we can
write 1 − 1 = 𝑔(𝜃1 , 𝜃2 ) , 1 − 2 = 𝑔(𝜃2 , 𝜃3 ) and 1 − 3 = 𝑔(𝜃1 , 𝜃3 ) , leading to
𝑔(𝜃1 , 𝜃2 )𝑔(𝜃2 , 𝜃3 ) = 𝑔(𝜃1 , 𝜃3 ), where 𝑔(∙) is any continuous function. The only way to reach
𝑓(𝜃 ) 𝑓(𝜃 ) 𝑓(𝜃 )
this last equation is to let 𝑔(𝜃1 , 𝜃2 ) = 𝑓(𝜃2 ) , 𝑔(𝜃2 , 𝜃3 ) = 𝑓(𝜃3 ) and 𝑔(𝜃1 , 𝜃3 ) = 𝑓(𝜃3 ) with
1 2 1

certain function 𝑓(𝜃). Now let us simply take 𝑓(𝜃) as a temperature scale, say, (𝜃) = 𝑇 , which
𝑇 𝑇
turns out to be the absolute temperature. Kelvin finally got 1 − 1 = 𝑇2 , 1 − 2 = 𝑇3 et 1 −
1 2

𝑇3 𝑇
3 = 𝑇 , or in general, for a Carnot engine working between 𝑇1 and 𝑇2 (𝑇1 > 𝑇2 ):  = 1 − 𝑇2
1 1

[39].

23
III) Clausius’ discovery

Clausius' work [10] starts with a Carnot engine working between 𝑇1 and 𝑇2 . He writes  =
𝑄 𝑇 𝑄 𝑇 𝑄1 𝑄
1 − 𝑄2 = 1 − 𝑇2 , or 𝑄2 = 𝑇2, and 0 = − 𝑇2. With a change of notation 𝛿𝑄1 = 𝑄1 , and 𝛿𝑄2 =
1 1 1 1 𝑇1 2

𝛿𝑄1 𝛿𝑄2
−𝑄2 , he obtained + = 0. Notice that 𝛿𝑄 is a heat absorbed by the engine from a
𝑇1 𝑇2

reservoir (see Figure A3). Obviously, if the engine is in contact with N reservoirs during a cycle,
𝛿𝑄𝑖
and absorbs 𝛿𝑄𝑖 from a reservoir at 𝑇𝑖 , we necessarily have ∑𝑁
𝑖=1 = 0 (Figure A3).
𝑇𝑖

Figure A3: Examples of the cycle of Carnot engine in the 𝑇 − 𝑆 diagram. The horizontal
segments represent the contact of the engine with the reservoirs to absorb heat. The
vertical segments represent adiabatic compression (left sides) and expansion (right sides).
Left : a cycle starting from and ending at an initial state A after two contacts with 𝑇1 and
𝛿𝑄1 𝛿𝑄2
𝑇2 , leading to + = 0. Right : a cycle starting from and ending at the initial state A
𝑇1 𝑇2

after the contacts with N reservoirs at different temperature 𝑇𝑖 , each giving 𝛿𝑄𝑖 to the
𝛿𝑄𝑖
engine, leading to ∑𝑁
𝑖=1 = 0.
𝑇𝑖

𝛿𝑄
Clausius understood at once that is the variation of a variable of state, he denoted by S
𝑇
𝛿𝑄
(𝛿𝑆 = ) and coined the name entropy, because when the engine comes back to its initial state
𝑇
𝛿𝑄𝑖
after a cycle and a series of variations , this variable also comes back to its initial value with
𝑇𝑖
𝛿𝑄𝑖
zero variation ∑𝑁
𝑖=1 = ∑𝑁
𝑖=1 𝛿𝑆𝑖 = 0. If N is vary large, we can write a closed line integral
𝑇𝑖

∮ 𝑑𝑆 = 0 over a cycle.

24
Clausius made the above analysis with Carnot engine. What does happen with a real
𝑄
irreversible engine working between the same reservoirs with ′ = 1 − 𝑄2? Considering ′ <
1

𝑇2 𝑄2 𝑇2 𝑄1 𝑄2
 and  = 1 − 𝑇 , we get 1 − 𝑄 ≤ 1 − 𝑇 or 𝑇 − 𝑇 < 0. The same formal change as above
1 1 1 1 2

𝛿𝑄1 𝛿𝑄2 𝛿𝑄𝑖 𝛿𝑄


yields + < 0. With N reservoirs during a cycle, we have ∑𝑁
𝑖=1 < 0 or ∮ < 0.
𝑇1 𝑇2 𝑇𝑖 𝑇

Suppose that the variation of entropy of the engine at each equilibrium contact with a reservoir
is 𝛿𝑆𝑖 , we have, after a cycle, ∑𝑁
𝑖=1 𝛿𝑆𝑖 = 0, or ∮ 𝑑𝑆 = 0 for large N. Compare ∮ 𝑑𝑆 = 0 to
𝛿𝑄 𝛿𝑄 𝜹𝑸
∮ < 0, it is straightforward to write 𝑑𝑆 > for irreversible engine, or 𝒅𝑺 ≥ for any
𝑇 𝑇 𝑻

engine, reversible or irreversible. This inequality is the mathematical expression of the second
law of thermodynamics. For an isolated system without exchange of heat with the exterior,
𝛿𝑄 = 0, and 𝑑𝑆 ≥ 0. This is why the second law is often referred to as the law of increasing
entropy. Notice that the entropy of an open system exchanging heat with environment has 𝑑𝑆 ≥
𝛿𝑄
and can decrease (𝑑𝑆 < 0) if the system loses heat 𝛿𝑄 < 0 (like our body); this decrease
𝑇

does not violate the second law. Again, the violation of the second law implies the decreases of
entropy in an isolated system.

IV) Perpetual motion machine of the second kind

As shown in Figure A4, a perpetual motion machine of the second kind (PMM2) is capable
of extracting a heat 𝑄3 from a cold reservoir R2 at 𝑇2 and transform it entirely into a mechanical
work 𝑊3 = 𝑄3 . This cycle of PMM2 does not violate the first law of energy conservation.

However, the work 𝑊3 can be used by a Carnot engine in an inverse cycle, as shown in
Figure A4, to extract a heat 𝑄2 from the colder reservoir R2 and to transfer a heat 𝑄1 to a hotter
reservoir R1 at 𝑇1 , with 𝑄1 = 𝑄2 + 𝑊3 = 𝑄2 + 𝑄3 . During this cycle, the reservoir R2 loses an
𝑄2 +𝑄3 𝑄1 𝑄2 +𝑄3
entropy ∆𝑆2 = − , and the reservoir R1 receives an entropy ∆𝑆1 = = , giving a
𝑇2 𝑇1 𝑇1
1 1
total entropy change ∆𝑆 = ∆𝑆1 + ∆𝑆2 = 𝑄1 (𝑇 − 𝑇 ) < 0 , which violates the Clausius
1 2

expression of the second law Eq.(4) ∆𝑆 ≥ 0 for the isolated system composed of the two
reservoirs and the engines in Figure A4.

25
Figure A4: Illustration of how a perpetual machine of the second kind (PMM2)
allows a flow of heat 𝑄1 = 𝑄2 + 𝑄3 from a colder body to a hotter body without any
other effects upon the environment.

In order to show how this violation of the second law by PMM2 enables PMM1, we
introduce a second Carnot engine in Figure A4 (see Figure A5), which absorbs 𝑄1′ during a
cycle from the reservoir R1, does a work 𝑊 to the environment, and emits a heat 𝑄2′ to the
𝑇 𝑄′ 𝑄1′ 𝑄2′
reservoir R2. Its efficiency is  = 1 − 𝑇2 = 1 − 𝑄2′ , leading to = . During the same cycle,
1 1 𝑇1 𝑇2

the system composed of the two engines in Figure A4 transfers a heat 𝑄1 from R2 to R1. The
ensemble of the three engines in Figure A5 (surrounded by red line) is equivalent to an global
engine which in one cycle absorbs a heat 𝑄1′ − 𝑄1 from R1, does a work W, and emits a heat
𝑄 ′ −𝑄 𝑄2′ −𝑄1 𝑄′
𝑄2′ − 𝑄1 to R2, leading to an efficiency ′ = 1 − 𝑄2′ −𝑄1. It is easy to prove that < 𝑄2′ ,
1 1 𝑄1′ −𝑄1 1

making ′ >  for any 𝑄1 > 0. It turns out that the global engine in Figure A5 is more efficient
than the Carnot engine (violating the first part of Carnot theorem). We can of course use it to
construct a PMM1 with the same tricks of Figure A1.

In similar way, we can also create a PMM1 using a PPM2 associated with an irreversible
engine replacing the second Carnot engine in Figure A5. In this case, the entropy increase

26
1 1
created with the system of two engines of Figure A4, ∆𝑆 = 𝑄1 (𝑇 − 𝑇 ), must be larger than
1 2

𝑄′ 𝑄′
( 2− 1)
𝑄2′ 𝑄′ 𝑇2 𝑇1
− 𝑇1, i.e., 𝑄1 > 1 1 .
𝑇2 1 ( − )
𝑇1 𝑇2

𝑄 ′ −𝑄
Figure A5: Illustration of an engine (red line) having an efficiency ′ = 1 − 𝑄2′ −𝑄1
1 1
𝑇2 𝑄2′
larger than the efficiency of the Carnot engine  = 1 − 𝑇 = 1 − 𝑄′ .
1 1

27

You might also like