You are on page 1of 31

Boltzmann's Work in Statistical Physics

First published Wed 17 Nov, 2004

Ludwig Boltzmann (1844-1906) is generally acknowledged as one of the most important


physicists of the nineteenth century. Particularly famous is his statistical explanation of
the second law of thermodynamics. The celebrated formula S = k logW, expressing a
relation between entropy S and probability W has been engraved on his tombstone (even
though he never actually wrote this formula down). Boltzmann's views on statistical
physics continue to play an important role in contemporary debates on the foundations of
that theory.

However, Boltzmann's ideas on the precise relationship between the thermodynamical


properties of macroscopic bodies and their microscopic constitution, and the role of
probability in this relationship are involved and differed quite remarkably in different
periods of his life. Indeed, in his first paper in statistical physics of 1866, he claimed to
obtain a completely general theorem from mechanics that would prove the second law.
However, thirty years later he stated that the second law could never be proved by
mechanical means alone, but depended essentially on probability theory. In his lifelong
struggle with the problem he employed a varying arsenal of tools and assumptions. (To
mention a few: the so-called Stoßzahlansatz, the ergodic hypothesis, ensembles, the
permutational argument, the hypothesis of molecular disorder.) However, the exact role
of these assumptions, and the results he obtained from them, also shifted in the course of
time. Particularly notorious are the role of the ergodic hypothesis and the status of the so-
called H-theorem. Moreover, he used ‘probability’ in four different technical meanings. It
is, therefore, not easy to speak of a consistent, single "Boltzmannian approach" to
statistical physics. It is the purpose of this essay to describe the evolution of a selection of
these approaches and their conceptual problems.

• 1. Introduction
o 1.1 Popular perceptions of Boltzmann
o 1.2 Debates and controversies
o 1.3 Boltzmann's relevance for the foundations of statistical physics
o 1.4 A concise chronography of Boltzmann's writings
• 2. The Stoßzahlansatz and the ergodic hypothesis
o 2.1 Doubts about the ergodic hypothesis
• 3. The H-theorem and the reversibility objection
o 3.1 1872: The Boltzmann equation and H-theorem
o 3.2 Remarks and problems
o 3.3 1877: The reversibility objection
• 4. 1877b: The permutational argument
o 4.1 Remarks and problems
• 5. Some later work
o 5.1 Return of the ergodic hypothesis
o 5.2 Return of the reversibility objection
o 5.3 The debate in Nature
• Bibliography
• Other Internet Resources
• Related Entries

1. Introduction
1.1 Popular perceptions of Boltzmann

Boltzmann's work met with mixed reactions during his lifetime, and continue to do so
even today. It may be worthwhile, therefore, to devote a few remarks to the perception
and reception of his work.

Boltzmann is often portrayed as a staunch defender of the atomic view of matter, at a


time when the dominant opinion in the German-speaking physics community, led by
influential authors like Mach and Ostwald, disapproved of this view. Indeed, the story
goes, in the late nineteenth century any attempt at all to search for a hypothetical,
microphysical underpinning of macroscopic phenomena was regarded as suspicious.
Further, serious criticism on his work was raised by Loschmidt and Zermelo. Various
passages in Boltzmann's writing, especially in the late 1890s, complain that his work was
hardly noticed (entitling one article "On some of my lesser-known papers on gas theory
and their relation to the same" (1879b) or even about a "hostile attitude" (1898a, v)
towards gas theory, and of his awareness of "being a powerless individual struggling
against the currents of the time" (ibid.).

Thus, the myth has arisen that Boltzmann was ignored or resisted by his contemporaries.
[1]
Sometimes, his suicide in 1906 is attributed to the injustice he thus suffered, The fact
that his death occurred just at the dawn of the definitive victory of the atomic view in the
works of Einstein, Smoluchowski, Perrin et al. adds a further touch of drama to this
picture.

As a matter of fact, Boltzmann's reputation as a theoretical physicist was actually widely


known and well-respected. In 1888 he was offered (but declined, after a curious sequence
of negotiations) a most prestigious chair in Berlin. Later, several universities (Vienna,
Munich, Leipzig) competed to get him appointed, sometimes putting the salaries of
several professorships together in their effort (Lindley 2001). He was elected to
membership or honorary membership in many academies (cf. Höflechner 1994, 192),
received honorary doctorates, and was also awarded various medals. In short, there is no
factual evidence for the claim that Boltzmann was ignored or suffered any unusual lack
of recognition from his contemporaries. His suicide seems to have been due to factors in
his personal life (depressions and decline of health) rather than to any academic matters.

1.2 Debates and controversies


Boltzmann was involved in various disputes. But this is not to say that he was the
innocent victim of hostilities. In many debates he took the initiative by launching a
polemic attack on his colleagues. The most important debates were those with Mach and
Ostwald on the reality of atoms, and with colleagues who criticized Boltzmann's own
work in the form of the famous reversibility objection (Loschmidt) and the recurrence
objection (Zermelo).

Ostwald and Mach clearly resisted the atomic view of matter (although for different
reasons). Boltzmann certainly defended and promoted this view. But he was not the naive
realist or unabashed believer in the existence of atoms that the more popular literature has
made of him. Instead, he stressed from the 1880s onwards that the atomic view yielded at
best an analogy, or a picture or model of reality (cf. de Regt 1999). In his debate with
Mach he advocated (1897c, 1897d) this approach as a useful or economical way to
understand the thermal behavior of gases. This means that his views were quite
compatible with Mach's views on the goal of science.[2] What divided them was more a
strategic issue. Boltzmann claimed that no approach in natural science that avoids
hypotheses completely could ever succeed. He argued that those who reject the atomic
hypothesis in favor of a continuum view of matter were guilty of adopting hypotheses
too. Ultimately, the choice between such views should depend on their fruitfulness, and
here Boltzmann had no doubt that the atomic hypothesis would be more successful.[3]

In the case of Ostwald, and his ‘energetics’, Boltzmann did become involved in a more
heated dispute at a meeting in Lübeck in 1895. Roughly speaking, energetics presented a
conception of nature that took energy as the most fundamental physical entity, and thus
represented physical processes as transformations of various forms of energy. It resisted
attempts to comprehend energy, or these transformations in terms of mechanical pictures.

It has been suggested that in the 1890s "the adherents of energetics reigned supreme in
the German school and even throughout Europe" (Dugas 1959, 82). But this is surely a
great exaggeration. It seems closer to the truth to say that energetics represented a rather
small but vocal minority in the physics community, that claimed to put forward a
seemingly attractive conception of natural science, and being promoted in the mid-90s by
reputed scientists, could no longer be dismissed as the work of amateurs (cf. Deltete
1999).

The 1895 gathering of the Naturforscherversammlung in Lübeck (the annual meeting of


physicists, chemists, biologists and physicians) was programmed to devote special
sessions to the state of the art of energetics. Boltzmann, who was member of the
programme committee, had already shown interest in the development of energetics in
private correspondence with Ostwald. Georg Helm was asked to prepare a report, and at
Boltzmann's own suggestion, Ostwald also contributed a lecture. All agreed that the
meeting should follow the "British style", i.e., manuscripts would be circulated
beforehand and there would be ample room for discussion, following the example of the
British Association for the Advancement of Science meeting that Boltzmann had
attended the previous year.
Both Helm and Ostwald, apparently, anticipated that they would have the opportunity to
discuss their views on energetics in an open-minded atmosphere. But at the meeting
Boltzmann surprised them with devastating criticism. According to those who were
present Boltzmann was the clear winner of the debate.[4] Yet the energeticists experienced
the confrontation as an ambush (Höflechner 1994, I, 169), for which he had not been
prepared. Nevertheless, Boltzmann and Ostwald remained friends, and in 1902 Ostwald
made a great effort to persuade his home university in Leipzig to appoint Boltzmann (cf.
Blackmore 1995, 61–65).

Neither is there any hostile attitude in the famous ‘reversibility objection’ by Loschmidt
in 1875. Loschmidt was Boltzmann's former teacher and later colleague at the University
of Vienna, and a life-long friend. He had no philosophical reservations against the
existence of atoms at all. (Indeed, he is best known for his estimate of their size.) Rather,
his main objection was against the prediction by Maxwell and Boltzmann that a gas
column in thermal equilibrium in a gravitational field has the same temperature at all
heights. His now famous reversibility objection arose in his attempts to undermine this
prediction. Whether Boltzmann succeeded in refuting the objection or not is still a matter
of dispute, as we shall see below (section 4.1).

Zermelo's opposition had a quite different background. When he put forward the
recurrence objection in 1896, he was an assistant to Planck in Berlin. And like his
mentor, he did not favor the mechanical underpinning of thermal phenomena. Yet his
1896 paper (Zermelo 1896a) is by no means hostile. It presents a careful logical argument
that leads him to a dilemma: thermodynamics with its Second Law on the one hand and
gas theory (in the form as Zermelo understood it) on the other cannot both be literally
true. By contrast, it is Boltzmann's (1896b) reaction to Zermelo, drenched in sarcasm and
bitterness which (if anything) may have led to hostile feelings between these two authors.
In any case, the tone of Zermelo's (1896b) is considerably sharper. Still, Zermelo
maintained a keen, yet critical, interest in gas theory and statistical physics, and
subsequently played an important role in making Gibbs' work known in Germany.

In fact, I think that Boltzmann's rather aggressive reactions to Zermelo and Ostwald
should be compared to other polemical exchanges in which he was involved, and
sometimes initiated himself (e.g. against Clausius, Tait, Planck, and Bertrand — not to
mention his essay on Schopenhauer). It seems to me that Boltzmann enjoyed polemics,
and the use of sharp language for rhetorical effect.[5] Boltzmann's complaints in 1896–
1898 about an hostile environment are, I think, partly explained by his love of polemic
exaggerations, partly also by his mental depression in that period. (See Höflechner 1994,
198–202) for details.) Certainly, the debates with Ostwald and Zermelo might well have
contributed to this personal crisis. But it would be wrong to interpret Boltzmann's
plaintive moods as evidence that his critics were, in fact, hostile.

Even today, commentators on Boltzmann's works are divided in their opinion. Some
praise them as brilliant and exceptionally clear. Often one finds passages suggesting he
possessed all the right answers all along the way — or at least in his later writings, while
his critics were simply prejudiced, confused or misguided (von Plato, Lebowitz, Kac,
Bricmont, Goldstein). Others (Ehrenfests, Klein, Truesdell) have emphasized that
Boltzmann's work is not always clear and that he often failed to indicate crucial
assumptions or important changes in his position, while friendly critics helped him in
clarifying and developing his views.

Fans and critics of Boltzmann's work alike agree that he pioneered much of the
approaches currently used in statistical physics, but also that he did not leave behind a
unified coherent theory. His scientific papers, collected in Wissenschaftliche Abhandlung,
contain more than 100 papers on statistical physics alone. Some of these papers are
forbiddingly long, full of tedious calculations and lack a clear coherent structure.
Sometimes, vital assumptions, or even a complete change of approach, are stated only
somewhere tucked away between the calculations, or at the very last page. Even
Maxwell, who might have been in the best position to appreciate Boltzmann's work,
expressed his difficulty with Boltzmann's longwindedness (in a letter to Tait, August
1873; see Garber, Brush, and Everett 1995, 123).[6] But not all of his prose is cumbersome
and heavy-going. Boltzmann at his best could be witty, passionate and a delight to read.
He excelled in such qualities in much of his popular work and some of his polemical
articles.

1.3 Boltzmann's relevance for the foundations of statistical physics

The foundations of statistical physics may today be characterized as a battlefield between


a dozen or so different schools, each firmly dug into their own trenches, e.g., ergodic
theory, coarse-graining (Markovianism), interventionism, BBKGY, Jaynes, Prigogine,
etc. Still, many of the protagonists of these schools, regardless of their disagreements,
frequently express their debt to ideas first formulated to Boltzmann. Even to those who
consider the concept of ensembles as the most important tool of statistical physics, and
claim Gibbs rather than Boltzmann as their champion, it has been pointed out that
Boltzmann introduced ensembles long before Gibbs. And those who advocate Boltzmann
while rejecting ergodic theory, may similarly be reminded that the latter theory too
originated with Boltzmann himself.

It appears, therefore, that Boltzmann is the father of many approaches, even if these
approaches are presently seen as conflicting with each other. This is due to the fact that
during his forty years of work on the subject, Boltzmann pursued many lines of thought.
Typically, he would follow a particular train of thought that seemed promising and
fruitful, only to discard it in the next paper for another one, and then pick it up again
years later. This meandering approach is of course not unusual among theoretical
physicists but it makes it hard to pin down Boltzmann on a particular set of rock-bottom
assumptions, that would reveal his true colors in the modern debate on the foundations of
statistical physics. The Ehrenfests (1912) in their famous Encyclopedia article, set
themselves the task of constructing a more or less coherent framework out of
Boltzmann's legacy. But their presentation of Boltzmann was, as is rather well known,
not historically adequate.
Without going into a more detailed description of the landscape of the battlefield of the
foundations of statistical physics, or a sketch of the various positions occupied, it might
be useful to mention only the roughest of distinctions. I use the term ‘statistical physics’
as a deliberately vague term that includes at least two more sharply distinguished
theories: the kinetic theory of gases and statistical mechanics proper.

The first theory aims to explain the properties of gases by assuming that they consist of a
very large number of molecules in rapid motion. (The term ‘kinetic’ is meant to underline
the vital importance of motion here, and to distinguish the approach from older static
molecular gas models.) During the 1860s probability considerations were imported into
this theory. The aim then became to characterize the properties of gases, in particular in
thermal equilibrium, in terms of probabilities of various molecular states. This is what the
Ehrenfests call "kineto-statistics of the molecule". Here, molecular states, in particular
their velocities, are regarded as stochastic variables, and probabilities are attached to such
molecular states of motion. These probabilities themselves are conceived of as
mechanical properties of the state of the total gas system. Either they represent the
relative number of molecules with a particular state, or the relative time during which a
molecule has that state.

In the course of time a transition was made to what the Ehrenfests called "kineto-statistics
of the gas model", or what is nowadays known as statistical mechanics. In this latter
approach, probabilities are not attached to the state of a molecule but of the entire gas
system. Thus, the state of the gas, instead of determining the probability distribution, now
itself becomes a stochastic variable.

A merit of this latter approach is that interactions between molecules can be taken into
account. Indeed, the approach is not restricted to gases, but also applicable to liquids or
solids. The price to be paid, however, is that the probabilities themselves become more
abstract. Since probabilities are attributed to the mechanical states of the total system,
they are no longer determined by such mechanical states. Instead, in statistical
mechanics, the probabilities are usually determined by means of an ‘ensemble’, i.e., a
fictitious collection of replicas of the system in question.

It is not easy to pinpoint this transition in the course of history, except to say that in
Maxwell's work in the 1860s definitely belong to the first category, and Gibbs' book of
1902 to the second. Boltzmann's own works fall somewhere in the middle. His earlier
contributions clearly belong to the kinetic theory of gases (although his 1868 paper
already applies probability to an entire gas system), while his work after 1877 is usually
seen as elements in the theory of statistical mechanics. However, Boltzmann himself
never indicated a clear distinction between these two different theories, and any attempt
to draw a demarcation at an exact location in his work seems somewhat arbitrary.

From a conceptual point of view, the transition from kinetic gas theory to statistical
mechanics poses two main foundational questions. On what grounds do we choose a
particular ensemble, or the probability distribution characterizing the system? Gibbs did
not enter into a systematic discussion of this problem, but only discussed special cases of
equilibrium ensembles (i.e. canonical, micro-canonical etc.). A second problem is to
relate the ensemble-based probabilities with the probabilities obtained in the earlier
kinetic approach for a single gas model.

The Ehrenfests (1912) paper was the first to recognize these questions, and to provide a
partial answer: Assuming a certain hypothesis of Boltzmann's, which they dubbed the
ergodic hypothesis, they pointed out that for an isolated system the micro-canonical
distribution is the unique stationary probability distribution. Hence, if one demands that
an ensemble of isolated systems describing thermal equilibrium must be represented by a
stationary distribution, the only choice for this purpose is the micro-canonical one.
Similarly, they pointed out that under the ergodic hypothesis infinite time averages and
ensemble averages were identical. This, then, would provide a desired link between the
probabilities of the older kinetic gas theory and those of statistical mechanics, at least in
equilibrium and in the infinite time limit. Yet the Ehrenfests simultaneously expressed
strong doubts about the validity of the ergodic hypothesis. These doubts were soon
substantiated when in 1913 Rozenthal and Plancherel proved that the hypothesis was
untenable for realistic gas models.

The Ehrenfests' reconstruction of Boltzmann's work thus gave a prominent role to the
ergodic hypothesis, suggesting that it played a fundamental and lasting role in his
thinking. Although this view indeed produces a more coherent view of his multifaceted
work, it is certainly not historically correct. Boltzmann himself also had grave doubts
about this hypothesis, and expressly avoided it whenever he could, in particular in his two
great papers of 1872 and 1877b. Since the Ehrenfests, many other authors have presented
accounts of Boltzmann's work. Particularly important are Klein (1973) and Brush (1976).
Still, much confusion remains about what exactly his approach to statistical physics was,
and how it developed.

1.4 A concise chronography of Boltzmann's writings

Roughly speaking, one may divide Boltzmann's work in four periods. The period 1866-
1871 is more or less his formative period. In his first paper (1866), Boltzmann set himself
the problem of deriving the full second law from mechanics. The notion of probability
does not appear in this paper. The following papers, from 1868 and 1871, were written
after Boltzmann had read Maxwell's work of 1860 and 1867. Following Maxwell's
example, they deal with the characterization of a gas in thermal equilibrium, in terms of a
probability distribution. Even then, he was set on obtaining more general results, and
extended the discussion to cases where the gas is subject to a static external force, and
might consist of poly-atomic molecules. He regularly switched between different
conceptions of probability: sometimes this referred to a time average, sometimes a
particle average or, in an exceptional paper (1871b), it referred to an ensemble average.
The main result of those papers is that from the Stoßzahlansatz (SZA) (or an analogous
assumption) the Maxwellian distribution function is stationary, and thus an appropriate
candidate for the equilibrium state. In some cases Boltzmann also argued it was the
unique such state.
However, in this period he also presented a completely different method, which did not
rely on the SZA but rather on the ergodic hypothesis. This approach led to a new form of
the distribution function that, in the limit N → ∞, reduces to the Maxwellian form. In the
same period, he also introduced the concept of ensembles, but this concept would not
play a prominent role in his thinking until the 1880's.

The next period is that of 1872-1878, in which he wrote his two most famous papers:
(1872) (Weitere Studien) and (1877b) (Über die Beziehung). The 1872 paper contained
the Boltzmann equation and the H-theorem. Boltzmann claimed that the H-theorem
provided the desired theorem from mechanics corresponding to the second law. However,
this claim came under a serious objection due to Loschmidt's criticism of 1876. The
objection was simply that no purely mechanical theorem could ever produce a time-
asymmetrical result. Boltzmann's response to this objection will be summarized later.

The result was, however, that Boltzmann rethought the basis of his approach and in
1877b produced a conceptually very different analysis, which might be called the
permutational argument, of equilibrium and evolutions towards equilibrium, and the role
of probability theory. The distribution function, which formerly represented the
probability distribution, was now conceived of as a stochastic variable (nowadays called
a macrostate) subject to a probability distribution. That probability distribution was now
determined by the size of the volume in phase space corresponding to all the microstates
giving rise to the same macrostate, (essentially given by calculating all permutations of
the particles in a given macrostate). Equilibrium was now conceived of as the most
probable macrostate instead of a stationary macrostate. The evolution towards
equilibrium could then be reformulated as an evolution from less probable to more
probable states.

Even though all commentators agree on the importance of these two papers, there is still
disagreement about what Boltzmann's claims actually were, and whether he succeeded
(or indeed even attempted) in avoiding the reversibility objection in this new
permutational argument, whether he intended or succeeded to prove that most evolutions
go from less probable to more probable states and whether or not he (implicitly) relied on
the ergodic hypothesis in these works. I shall comment on these issues in due course.

The third period is taken up by the papers Boltzmann wrote during the 1880's have
attracted much less attention. During this period, he abandoned the permutational
argument, and went back to an approach that relied on a combination of the ergodic
hypothesis and the use of ensembles. For a while Boltzmann worked on an application of
this approach to Helmholtz's concept of monocyclic systems. However, after finding that
concept did not always provide the desired thermodynamical analogies, he abandoned
this topic again.

Next, in the 1890s the reversibility problem resurfaced again, this time in a debate in the
columns of Nature. This time Boltzmann chose an entirely different line of
counterargument than in his debate with Loschmidt. A few years later, Zermelo presented
another objection, now called the recurrence objection. The same period also saw the
publication of the two volumes of his Lectures on Gas Theory. In this book, he takes the
hypothesis of molecular disorder (a close relative of the SZA) as the basis of his
approach. The permutational argument is only discussed as an aside, and the ergodic
hypothesis is not mentioned at all. His last paper is an Encyclopedia article with Nabl
presenting a survey of kinetic theory.

2. The Stoßzahlansatz and the ergodic hypothesis


Boltzmann's first paper (1866) in statistical physics aimed to reduce the second law to
mechanics. Within the next two years he became acquainted with Maxwell's papers on
gas theory of 1860 and 1867, which introduced probability notions in the description of
the gas. Maxwell had studied specific mechanical models for a gas (as a system of hard
spheres (1860) or of point particles exerting a mutual force on each other inversely
proportional to the fifth power of their distance), and characterized the state of such a gas
by means of a probability distribution f over the various values of the molecular velocities
. For Maxwell, the probability f( )d3 denoted the relative number of particles in the
gas with a velocity between and + d3 . In particular, he had argued that the state of
equilibrium is characterized by the so-called Maxwell distribution function:
(1) f( ) = Ae−| |2/B

where A is a normalization constant and B is proportional to the absolute temperature.

The argument that Maxwell had given in 1860 to single out this distribution relied on the
fact that this is the only probability distribution that is both spherically symmetric and
factorizes into functions of the orthogonal components x, y, z separately. In 1867,
however he replaced these desiderata with the more natural requirement that the
equilibrium distribution should be stationary, i.e. it should not change shape as a result of
the continual collisions between the particles. This called for a more elaborate argument,
involving a detailed consideration of the collisions between particles. The crucial
assumption in this argument is what is now known as the SZA. Roughly speaking, it
states that the number of particle pairs, dN( 1, 2) with initial velocities between 1 and
3 3
1+d 1, and between 2 and 2+d 2, respectively, which are about to collide in a time
span dt is proportional to

(2) N( 1, 2) ∝ N2f( 1)f( 2) dt d3 1 d3 2

where the proportionality constant depends on the geometry of the collision and the
relative velocity. For Maxwell, and Boltzmann later, this assumption seemed almost self-
evident. One ought to note, however, that by choosing the initial, rather than the final
velocities of the collision, the assumption introduced an explicit time-asymmetric
element. This, however, was not noticed until 1895. Maxwell showed that, under the
SZA, the distribution (1) is indeed stationary. He also argued, but much less
convincingly, that it should be the only stationary distribution.

In his (1868), Boltzmann set out to apply this argument to a variety of other models
(including gases in a static external force field). However, Boltzmann started out with a
somewhat different interpretation of probability in mind than Maxwell. For him, f( )d3
is introduced firstly as the relative time during which a (given) particle has a velocity
between and +d3 (WA I, 50). But, in the same breath, he identifies this with the
relative number of particles with this velocity. This equivocation between different
meanings of probability returned again and again in Boltzmann's writing.[7] Either way, of
course, whether we average over time or particles, probabilities are defined here in
strictly mechanical terms, and therefore objective properties of the gas. Yet apart from
this striking difference in interpretation, the first section of the paper is a straightforward
continuation of the ideas Maxwell had developed in his 1867. In particular, the main
ingredient is always played by the SZA, or a version of that assumption suitably modified
for the case discussed.

But in the last section of the paper he suddenly shifts course. He now focuses on a
general Hamiltonian system, i.e., a system of N material points with an arbitrary
interaction potential. The state of this system may be represented as a phase point x = (
1,…, N, 1,…, N) in the mechanical phase space Γ. By the Hamiltonian equations of
motion, this point evolves in time, and thus describes a trajectory xt. This trajectory is
constrained to lie on a given energy hypersurface H(x) = E, where H(x) denotes the
Hamiltonian function.

Now consider an arbitrary probability density ρ(x) over this phase space. He shows, by
(what is now known as) Liouville's theorem, that ρ remains constant along a trajectory,
i.e., ρ(x0) = ρ(xt). Assuming now for simplicity that all points in a given energy
hypersurface lie on a single trajectory, the probability should be a constant over the
energy hypersurface. In other words, the only stationary probability with fixed total
energy is the microcanonical distribution.

(3) ρmc = δ(H(x) − E),

where δ is Dirac's delta function.

By integrating this expression over all momenta but one, and dividing this by the integral
of ρmc over all momenta, Boltzmann obtained the marginal probability density ρmc( 1 | 
1,…, N) for particle 1's momentum, conditionalized on the particle positions 1… N  .
He then showed that this marginal probability distribution tends to the Maxwell
distribution when the number of particles tends to infinity.

Some comments on this result.

First, the difference between the approach relying on the ergodic hypothesis and that
relying on the SZA is rather striking. Instead of concentrating on a specific gas model,
Boltzmann here assumes a much more general model with an arbitrary interaction
potential V( 1,…, N). Moreover, the probability density ρ is defined over phase space,
instead of the space of molecular velocities. This is the first occasion where probability
considerations are applied to the state of the mechanical system as whole, instead of its
individual particles. If the transition between kinetic gas theory and statistical mechanics
may be identified with this caesura, (as argued by the Ehrenfests and by Klein) it would
seem that the transition has already been made right here.

But of course, for Boltzmann the transition did not involve a major conceptual move,
thanks to his conception of probability as a relative time. Thus, the probability of a
particular state of the total system is still identified with the fraction of time in which that
state is occupied by the system. In other words, he had no need for ensembles or non-
mechanical probabilistic assumptions in this paper.

However, note that the equivocation between relative times and relative numbers, which
was relatively harmless in the first section of the 1868 paper, is no longer possible in the
interpretation of ρ. The probability ρmc( 1 |  1,…, n)d3 1 gives us the relative time that
the total system is in a state for which particle 1 has a momentum between 1 and 1 +
d3 1, for given values of all positions. There is no route back to infer that this has
anything to do with the relative number of particles with this momentum.

Second, and more importantly, these results open up a perspective of great generality. It
suggests that the probability of the molecular velocities for an isolated system in a
stationary state will always assume the Maxwellian form if the number of particles tends
to infinity. Notably, this argument completely dispenses with any particular assumption
about collisions, or other details of the mechanical model involved, apart from the
assumption that it is Hamiltonian. Indeed it need not even represent a gas.

Third, and finally, the main weakness of the present result is its assumption that the
trajectory actually visits all points on the energy hypersurface. This is what the Ehrenfests
called the ergodic hypothesis.[8] Boltzmann returned to this issue on the final page of the
paper (WA I, 96). He notes there that exceptions to his theorem might occur, if the
microscopic variables would not, in the course of time, take on all values compatible with
the conservation of energy. For example this would be the case when the trajectory is
periodic. However, Boltzmann observed, such cases would be immediately destroyed by
the slightest disturbance from outside, e.g., by the interaction of a single external atom.
He argued that these exceptions would thus only provide cases of unstable equilibrium.

Still, Boltzmann must have felt unsatisfied with his own argument. According to an
editorial footnote in the collection of his scientific papers (WA I, 96), Boltzmann's
personal copy of the paper contains a hand-written remark in the margin stating that the
point was still dubious and that it had not been proven that, even including interaction
with an external atom, the trajectory would traverse all points on the energy hypersurface.

2.1 Doubts about the ergodic hypothesis

However, his doubts were still not laid to rest. His next paper on gas theory (1871a)
returns to the study of a detailed mechanical gas model, this time consisting of
polyatomic molecules, and avoids any reliance on the ergodic hypothesis. And when he
did return to the ergodic hypothesis in (1871b), it was with much more caution. Indeed, it
is here that he actually first described the worrying assumption as an hypothesis,
formulated as follows:

The great irregularity of the thermal motion and the multitude of forces that act on a body
make it probable that its atoms, due to the motion we call heat, traverse all positions and
velocities which are compatible with the principle of [conservation of] energy. (WA I,
284)

Note that Boltzmann formulates this hypothesis for an arbitrary body, i.e., it is not
restricted to gases. He also remarks, at the end of the paper, that "the proof that this
hypothesis is fulfilled for thermal bodies, or even is fullfillable, has not been provided"
(WA I, 287).

There is a major confusion among modern commentators about the role and status of the
ergodic hypothesis in Boltzmann's thinking. Indeed, the question has often been raised
how Boltzmann could ever have believed that a trajectory traverses all points on the
energy hypersurface, since, as the Ehrenfests conjectured in 1911, and was shown almost
immediately in 1913 by Plancherel and Rozenthal, this is mathematically impossible
when the energy hypersurface has a dimension larger than 1.

It is a fact that both (1868) [WA I, 96] and (1871b) [WA I, 284] mention external
disturbances as an ingredient in the motivation for the ergodic hypothesis. This might be
taken as evidence for ‘interventionalism’, i.e., the viewpoint that such external influences
are crucial in the explanation of thermal phenomena (see Blatt 1959, Ridderbos &
Redhead 1998). Yet even though Boltzmann clearly expressed the thought that these
disturbances might help to motivate the ergodic hypothesis, he never took the idea very
seriously. The marginal note in the 1868 paper mentioned above indicated that, even if
the system is disturbed, there is still no easy proof of the ergodic hypothesis, and all his
further investigations concerning this hypothesis assume a system that is either
completely isolated from its environment or at most acted upon by a static external force.
Thus, interventionalism did not play a significant role in his thinking.[9]

It has also been suggested, in view of Boltzmann's later habit of discretising continuous
variables, that he somehow thought of the energy hypersurface as a discrete manifold
containing only finitely many discrete cells (Gallavotti 1994). In this reading, obviously,
the mathematical no-go theorems of Rozenthal and Plancherel no longer apply. Now it is
definitely true that Boltzmann developed a preference towards discretizing continuous
variables, and would later apply this procedure more and more (although usually adding
that this was fictitious and purely for purposes of illustration and more easy
understanding). However, there is no evidence in the (1868) and (1871b) papers that
Boltzmann implicitly assumed a discrete structure of mechanical phase space or the
energy hypersurface.
Instead, the context of his (1871b) makes clear enough how he intended the hypothesis,
as has already been argued by (Brush 1976). Immediately preceding the section in which
the hypothesis is introduced, Boltzmann discusses trajectories for a simple example: a
two-dimensional harmonic oscillator with potential V(x, y) = ax2 + by2. For this system,
the configuration point (x, y) moves through the surface of a rectangle. See Figure 1
below. (See also Cercignani 1998, 148.)

Figure 1: a/b is rational (= 4/7).

He then notes that if a/b is rational, (actually: if √(a/b) is rational) this motion is periodic.
However, if this value is irrational, the trajectory will, in the course of time, traverse
"almählich die ganze Fläche" (WA I, 271) of the rectangle. See Figure 2:

Figure 2: a/b is irrational (= 1/e).

He says in this case that x and y are independent, since for each values of x an infinity of
values for y in any interval in its range are possible. The very fact that Boltzmann
considers intervals for the values of x and y of arbitrary small sizes, and stressed the
distinction between rational and irrational values of the ratio a/b, indicates that he did not
silently presuppose that phase space was essentially discrete, where those distinctions
would make no sense.

Now clearly, in modern language, one should say in the second case that the trajectory
lies densely in the surface, but not that it traverses all points. Boltzmann did not possess
this language. In fact, he could not have been aware of Cantor's insight that the
continuum contains more than a countable infinity of points. Thus, the correct statement
that, in the case that √(a/b) is irrational, the trajectory will traverse, for each value of x, an
infinity of values of y within any interval however small, could easily have lead him to
believe (incorrectly) that all values of x and y are traversed in the course of time.

It thus seems eminently plausible, by the fact that this discussion immediately precedes
the formulation of the ergodic hypothesis, that the intended reading of the ergodic
hypothesis is really what the Ehrenfests dubbed the quasi-ergodic hypothesis, namely, the
assumption that the trajectory lies densely (i.e. passes arbitrarily close to every point) on
the energy hypersurface.[10] The quasi-ergodic hypothesis is not mathematically
impossible in higher-dimensional phase spaces. However, the quasi-ergodic hypothesis
does not entail the desired conclusion that the only stationary probability distribution over
the energy surface is micro-canonical. One might then still conjecture that if the system is
quasi-ergodic, the only continuous stationary distribution is microcanonical. But even this
is fails in general (Nemytskii and Stepanov 1960).

Nevertheless, Boltzmann remained skeptical about the validity of his hypothesis. For this
reason, he attempted to explore different routes to his goal of characterizing thermal
equilibrium in mechanics. Indeed, both the preceding (1871a) and his next paper (1871c)
present alternative arguments, with the explicit recommendation that they avoid
hypotheses. In fact, he did not return to this hypothesis until the 1880s (stimulated by
Maxwell's 1879 review of the last section of Boltzmann's 1868 paper). At that time,
perhaps feeling fortified by Maxwell's authority, he would express much more
confidence in the ergodic hypothesis (see Section 5).

So what role did the ergodic hypothesis play? It seems that Boltzmann regarded the
ergodic hypothesis as a special dynamical assumption that may or may not be true,
depending on the nature of the system, and perhaps also on its initial state. Its role was
simply to help derive a result of great generality: for any system for which the hypothesis
is true, its equilibrium state is characterized by (3), from which a form of the Maxwell
distribution may be recovered in the limit N → ∞, regardless of any details of the inter-
particle interactions, or indeed whether the system represented is a gas, fluid, solid or any
other thermal body.

The Ehrenfests have suggested that the ergodic hypothesis played a much more
fundamental role. In particular they have pointed out that if the hypothesis is true,
averaging over an (infinitely) long time would be identical to phase averaging with the
microcanonical distribution. Thus, they suggested that Boltzmann relied on the ergodic
hypothesis in order to equate time averages and phase averages, or in other words, to
equate two meaning of probability (relative time and relative volume in phase space.)
There is however no evidence that Boltzmann ever followed this line of reasoning. He
simply never gave any justification for equivocating time and particle averages, or phase
averages, at all. Presumably, he thought nothing much depended on this issue and that it
was a matter of taste.

3. The H-theorem and the reversibility objection


3.1 1872: The Boltzmann equation and H-theorem

In 1872 Boltzmann published one of his most important papers (Weitere Studien). It
contained two celebrated results nowadays known as the Boltzmann equation and the H-
theorem. The latter result was the basis of Boltzmann's renewed claim to have obtained a
general theorem corresponding to the second law. This paper has been studied and
commented upon by numerous authors. Indeed an integral translation of the text has been
provided by (Brush 1966). Thus, for the present purposes, a succinct summary of the
main points might have been sufficient. However, there is still dispute among modern
commentators about its actual content.

The issue at stake is the question whether the results obtained in this paper are presented
as necessary consequences of the mechanical equations of motion, or whether Boltzmann
explicitly acknowledged that they would allow for exceptions. Klein has written

I can find no indication in his 1872 memoir that Boltzmann conceived of possible
exceptions to the H-theorem, as he later called it. (Klein 1973, 73)

Klein argues that Boltzmann only came to acknowledge the existence of such exceptions
thanks to Loschmidt's critique in 1877. An opposite opinion is expressed by von Plato
(1994). He argues that, already in 1872, Boltzmann was well aware that his H-theorem
had exceptions, and thus "already had a full hand against his future critics". Indeed, von
Plato states that

… contrary to a widely held opinion, Boltzmann is not in 1872 claiming that the Second
Law and the Maxwellian distribution are necessary consequences of kinetic theory. (von
Plato 1994, 81)

It might be of some interest to try and settle this dispute.

The Weitere Studien starts with an appraisal of the role of probability theory in the
context of gas theory. The number of particles in a gas is so enormous, and their
movements are so swift that we can observe nothing but average values. The
determination of averages is the province of probability calculus. Therefore, "the
problems of the mechanical theory of heat are really problems in probability calculus"
(WA I, 317). But, Boltzmann says, it would be a mistake to believe that the theory of heat
would therefore contain uncertainties.

He emphasizes that one should not confuse incompletely proven assertions with
rigorously derived theorems of probability theory. The latter are necessary consequences
from their premisses, as in any other theory. They will be confirmed by experience as
soon as one has observed a sufficiently large number of cases. This last condition,
however, should be no significant problem in the theory of heat because of the enormous
number of molecules in macroscopic bodies. Yet, in this context, one has to make doubly
sure that we proceed with the utmost rigor.
Thus, the message expressed in the opening pages of this paper seems clear enough: the
results Boltzmann is about to derive are advertised as doubly checked and utterly
rigorous. Of course, their relationship with experience might be less secure, since any
probability statement is only reproduced in observations by sufficiently large numbers of
independent data. Thus, Boltzmann would have allowed for exceptions in the relationship
between theory and observation, but not in the relation between premisses and
conclusion.

He continues by saying what he means by probability, and repeats its equivocation as a


fraction of time and the relative number of particles that we have seen earlier in 1868a:

If one wants […] to build up an exact theory […] it is before all necessary to determine
the probabilities of the various states that one and the same molecule assumes in the
course of a very long time, and that occur simultaneously for different molecules. That is,
one must calculate how the number of those molecules whose states lie between certain
limits relates to the total number of molecules (WA I, 317).

This equivocation is not vicious however. For most of the paper the intended meaning of
probability is always the relative number of molecules with a particular molecular state.
Only at the final stages of his paper (WA I, 400) does the time-average interpretation of
probability (suddenly) recur.

Boltzmann says that both he and Maxwell had attempted the determination of these
probabilities for a gas system but without reaching a complete solution. Yet, on a closer
inspection, "it seems not so unlikely that these probabilities can be derived on the basis of
the equations of motion alone…" (WA I, 317). Indeed, he announces, he has solved this
problem for gases whose molecules consist of an arbitrary number of atoms. His aim is to
prove that whatever the initial state in such a system of gas molecules, it must inevitably
approach the state characterized by the Maxwell distribution (WA I, 320).

The next section specializes to the simplest case of monatomic gases and also provides a
more complete specification of the problem he aims to solve. The gas molecules are
modelled as hard spheres, contained in a fixed vessel with perfectly elastic walls (WA I,
320). Boltzmann represents the state of the gas by a time-dependent distribution function
ft( ) which gives us, at each time t, the relative number of molecules with velocity .[11]
He also states three more special assumptions:

a. Already in the initial state of the gas, each direction of velocity is equally
probable, i.e.,

f0( ) = f0( ).

b. The gas is spatially uniform. That is, the relative number of molecules with their
velocities in any given interval does not depend on the location within the vessel.
c. The Stoßzahlansatz (2).
After a few well-known manipulations, the result from these assumptions is a differentio-
integral equation (the Boltzmann equation) that determines the evolution of the
distribution function ft( ) from any given initial form.

There are also a few unstated assumptions that go into the derivation of this equation.
First, the number of molecules must be large enough so that the (discrete) distribution of
their velocities can be well approximated by a continuous and differentiable function f.
Secondly, f changes under the effect of binary collisions only. This means that the density
of the gas should be low (so that three-particle collisions can be ignored) but not too low
(so that collisions would be too infrequent to change f at all. (The modern procedure to
put these requirements in a mathematically precise form is that of taking the so-called
Boltzmann-Grad limit.) A final ingredient is that all the above assumptions are not only
valid at an instant but remain true in the course of time.

The H-theorem. Assuming that the Boltzmann equation is valid for all times, one can
prove without difficulty the "H-theorem": the quantity

(4) H[ft ] := ∫ft( ) ln ft( ) d3


decreases monotonically in time, i.e.,
dH [ft ]
(5) ----- ≤ 0
dt
as well as its stationarity for the Maxwell distribution, i.e.,
dH [ft ]/dt = 0, if ft( ) = Ae−B 2

Boltzmann concludes this section of the paper as follows:

It has thus been rigorously proved that whatever may have been the initial distribution of
kinetic energy, in the course of time it must necessarily approach the form found by
Maxwell. […] This [proof] actually gains much in significance because of its
applicability on the theory of multi-atomic gas molecules. There too, one can prove for a
certain quantity E that, because of the molecular motion, this quantity can only decrease
or in the limiting case remain constant. Thus, one may prove that, because of the atomic
movement in systems consisting of arbitrarily many material points, there always exists a
quantity which, due to these atomic movements, cannot increase, and this quantity agrees,
up to a constant factor, exactly with the value that I found in [Boltzmann 1871c] for the
well-known integral ∫dQ/T.

This provides an analytical proof of the Second Law in a way completely different from
those attempted so far. Up till now, one has attempted to proof that ∫dQ/T = 0 for
reversible (umkehrbaren) cyclic[12] processes, which however does not prove that for an
irreversible cyclic process, which is the only one note-that occurs in nature, it is always
negative; the reversible process being merely an idealization, which can be approached
more or less but never perfectly. Here, however, we immediately reach the result that
∫dQ/T is in general negative and zero only in a limit case…" (WA I, 345)
Thus, as in his 1866 paper, Boltzmann claims to have a rigorous, analytical and general
proof of the Second Law.

3.2 Remarks and problems

1. As we have seen, The H-theorem formed the basis of a renewed claim by Boltzmann to
have obtained a theorem corresponding to the second law, at least for gases. A main
difference with his previous (1866) claim, is that he now strongly emphasized the role of
probability calculus in his derivation. Even so, it will be noted that his conception of
probability is still a fully mechanical one. Thus, there is no conflict between his claims
that on the one hand, "the problems of the mechanical theory of heat are really problems
in probability calculus" and that the probabilities themselves are "derived on the basis of
the equations of motion alone", on the other hand. Indeed, it seems to me that
Boltzmann's emphasis on the crucial role of probability is only intended to convey that
probability theory provides a particularly useful and appropriate language for discussing
mechanical problems in gas theory. There is no indication in this paper yet that
probability theory could play a role by furnishing assumptions of a non-mechanical
nature, i.e., independent of the equations of motion.

2. Note that Boltzmann stresses the generality, rigor and "analyticity" of his proof. He put
no emphasis on the special assumptions that go into the argument. Indeed, the
Stoßzahlansatz, later identified as the key assumption that is responsible for the time-
asymmetry of the H-theorem, is announced as follows:

The determination [of the number of collisions] can only be obtained in a truly tedious
manner, by consideration of the relative velocities of both particles. But since this
consideration has, apart from its tediousness, not the slightest difficulty, nor any special
interest, and because the result is so simple that one might almost say it is self-evident I
will only state this result. (WA I, 323)

It thus seems natural that Boltzmann's contemporaries must have understood him as
claiming that the H-theorem followed necessarily from the dynamics of the mechanical
gas model. Indeed this is exactly how Boltzmann's claims were understood. For example,
the recommendation written in 1888 for his membership of the Prussian Academy of
Sciences mentions as Boltzmann's main feat that had proven that, whatever its initial
state, a gas must necessarily approach the Maxwellian distribution (Kirsten and Körber
1975, 109).

Is there then no evidence at all for von Plato's reading of the paper? Von Plato quotes a
passage from Section II, where Boltzmann repeats the previous analysis by assuming that
energy can take on only discrete values, and replacing all integrals by sums. He recovers,
of course, the same conclusion, but now adds a side remark, which touches upon the case
of non-uniform gases:

Whatever may have been the initial distribution of states, there is one and only one
distribution which will be approached in the course of time. […] This statement has been
proved for the case that the distribution of states was already initially uniform. It must
also be valid when this is not the case, i.e. when the molecules are initially distributed in
such a way that in the course of time they mix among themselves more and more, so that
after a very long time the distribution of states becomes uniform. This will always be the
case, with the exception of very special cases, e.g., when all molecules were initially
situated along a straight line, and were reflected by the walls onto this line. (WA I, 358)

True enough, Boltzmann in the above quote indicates that there are exceptions. But he
mentions them only in connection with an extension of his results to the case when the
gas is not initially uniform, i.e., when condition (b) above is dropped. There can be no
doubt that under the assumption of the conditions (a. – c.), Boltzmann claimed rigorous
validity of the H-theorem.

3. Note that Boltzmann misconstrues, or perhaps understates, the significance of his


results. Both the Boltzmann equation and the H theorem refer to a body of gas in a fixed
container that evolves in complete isolation from its environment. There is no question of
heat being exchanged by the gas during a process, let alone in an irreversible cyclic
process. His comparison with Clausius' integral ∫dQ/T (i.e., Q/T in modern notation)
is therefore really completely out of place.

The true import of Boltzmann's results is rather that they provide a generalization of the
entropy concept to non-equilibrium states, and a claim that this non-equilibrium entropy
−kH increases monotonically as the isolated gas evolves from non-equilibrium towards
an equilibrium state. The relationship with the second law is, therefore, indirect. On the
one hand, Boltzmann proves much more than was required, since the second law does not
speak of non-equilibrium entropy, nor of monotonic increase; on the other hand it proves
also less, since Boltzmann does not consider more general adiabatic processes.

3.3 1877: The reversibility objection

According to Klein (1973) Boltzmann seemed to have been satisfied with his treatments
of 1871 and 1872 and turned his attention to other matters for a couple of years. He did
come back to gas theory in 1875 to discuss an extension of the Boltzmann equation to
gases subjected to external forces. But this paper does not present any fundamental
changes of thought. However, the 1875 paper did contain a result which, two years later,
led to a debate with Loschmidt. It showed that a gas in equilibrium in an external force
field (such as the earth's gravity) should have a uniform temperature, and therefore, the
same average kinetic energy at all heights. This conclusion conflicted with the intuition
that rising molecules must do work against the gravitational field, and pay for this by
having a lower kinetic energy at greater heights.

Now Boltzmann (1875) was not the first to reach the contrary result, and Loschmidt was
not the first to challenge it. Maxwell and Guthrie entered into a debate on the very same
topic in 1873. But actually their main point of contention need not concern us very much.
The discussion between Loschmidt and Boltzmann is important for quite another issue
which Loschmidt only introduced as a side remark:
By the way, one should be careful about the claim that in a system in which the so-called
stationary state has been achieved, starting from an arbitrary initial state, this average
state can remain intact for all times. […]

Indeed, if in the above case [i.e. starting in a state where one particle is moving, and all
the others lie still on the bottom], after a time τ which is long enough to obtain the
stationary state, one suddenly assumes that the velocities of all atoms are reversed, we
would obtain an initial state that would appear to have the same character as the
stationary state. For a fairly long time this would be appropriate, but gradually the
stationary state would deteriorate, and after passage of the time τ we would inevitable
return to our original state: only one atom has absorbed all kinetic energy of the system
[…], while all other molecules lie still on the bottom of the container.

Obviously, in every arbitrary system the course of events must be become retrograde
when the velocities of all its elements are reversed. (Loschmidt 1876, 139)

Putting the point in more modern terms, the laws of (Hamiltonian) mechanics are such
that for every solution one can construct another solution by reversing all velocities and
replacing t by −t. Since H[f] is invariant under the velocity reversal, it follows that if H[f]
decreases for the first solution, it will increase for the second. Accordingly, the
reversibility objection is that the H-theorem cannot be a general theorem for all
mechanical evolutions of the gas.

Boltzmann's response (1877a). Boltzmann's responses to the reversibility objection are


not easy to make sense of, and varied in the course of time. In his immediate response to
Loschmidt he acknowledges that certain initial states of the gas would lead to an increase
of the H function, and hence a violation of the H theorem. The crux of his rebuttal was
that such initial states were extremely improbable, and could hence safely be ignored.

This argument shows that Boltzmann was already implicitly embarking on an approach
that differed from the context of the 1872 paper. Recall that this paper used the concept
of probability only in the guise of a distribution function, giving the probability of
molecular velocities. There was no such thing in that paper as the probability of a state of
the gas as whole. This conceptual shift would become more explicit in Boltzmann's next
paper (1877b).

This rebuttal of Loschmidt is far from satisfactory. Any reasonable probability


assignment to gas states is presumably invariant under the velocity reversal of the
molecules. If an initial state leading to an increase of H is to be ignored on account of its
small probability, one ought to assume the same for the state from which it was
constructed by velocity reversal. In other words, any non-equilibrium state would have to
be ignored. But that in effect saves the H-theorem by restricting it to those cases where it
is trivially true, i.e., where H is constant.

The true source of the reversibility problem was only identified by Burbury (1894a) and
Bryan1 (1894), by pointing out that already the Stoßzahlansatz contained a time-
asymmetric assumption. Indeed, if we replace the SZA by the assumption that the
number of collisions is proportional to the product f( 1′)f( 2′) for the velocities 1′, 2′
after the collision, we would obtain, by a similar reasoning, dH/dt ≤ 0. The question is
now, of course, we would prefer one assumption above the other, without falling into
some kind of double standards. One thing is certain, and that is that any such preference
cannot be obtained from mechanics and probability theory alone.

4. 1877b: The permutational argument


Succinctly, and rephrased in modern terms, the argument is as follows. Apart from Γ, the
mechanical phase space containing the possible states x for the total gas system, we
consider the so-called μ-space, i.e., the state space of a single molecule. For monatomic
gases, this space is just a six-dimensional space with ( , ) as coordinates. With each
state x is associated a collection of N points in μ-space.

We now partition μ into m disjoint cells: μ = ω1∪…∪ωm. These cells are taken to be
rectangular in the position and momentum coordinates and of equal size. Further, it is
assumed we can characterize each cell in μ with a molecular energy εi.

For each x, henceforth also called the microstate, we define the macrostate (Boltzmann's
term was Komplexion) as Ζ := (n1,…,nm), where ni is the number of particles that have
their molecular state in cell ωi. The relation between macro- and microstate is obviously
non-unique since many different microstates, e.g., obtained by permuting the molecules,
lead to the same macrostate. One may associate with every given macrostate Ζ0 the
corresponding set of microstates:

AΖ0 = {x∈Γ : Ζ(x) = Ζ0}

The volume |AΖ0| of this set is proportional to the number of permutations that lead to this
macrostate. Boltzmann proposes the problem to determine for which macrostate Ζ the
volume |AΖ| is maximal, under the constraints of a given total number of particles, and a
given total energy:

m m
(6) N = ∑ ni , E = ∑ niεi .
i=1 i=1

This problem can easily be solved with the Lagrange multiplier technique. Under the
Stirling approximation for ni >> 1 we find

ni ∝ eλεi ,

which is a discrete version of the Maxwell distribution.


Moreover, the volume of the corresponding set in Γ is related to a discrete approximation
of the H-function. Indeed, one finds

(7) −NH ≈ ln |AΖ|

In other words, if we take −kNH as the entropy of a macrostate, it is also proportional to


the logarithm of the volume of the corresponding region in phase space.

Boltzmann also refers to these volumes as the "probability" of the macrostate. He


therefore now expresses the second law as a tendency to evolve towards ever more
probable macrostates.

4.1 Remarks and problems

1. No dynamical assumption is made; i.e., it is not relevant to the argument whether or


how the particles collide. It might seem that this makes the present argument more
general than the previous one. Indeed, Boltzmann suggests at the end of the paper that the
same argument might be applicable also to dense gases and even to solids.

However, it should be noticed that the assumption that the total energy can be expressed
in the form E = ∑i niεi means that the energy of each particle depends only on the cell in
which it is located, and not the state of other particles. This can only be maintained,
independently of the number N, if there is no interaction at all between the particles. The
validity of the argument is thus really restricted to ideal gases.

2. The procedure of dividing μ space into cells is essential here. Indeed, the whole
prospect of using combinatorics would disappear if we did not adopt a partition. But the
choice to take all cells equal in size in position and momentum variables is not quite self-
evident, as Boltzmann himself shows. In fact, before he develops the argument above, his
paper first discusses an analysis in which the particles are characterized by their energy
instead of position and momentum. This leads him to carve up μ-space into cells of equal
size in energy. He then shows that this analysis fails to reproduce the desired Maxwell
distribution as the most probable state. This failure is remedied by taking equally sized
cells in position and momentum variables. The latter choice is apparently ‘right’, in the
sense that leads to the desired result. However, since the choice clearly cannot be
relegated to a matter of convention, it leaves the question for a justification.

3. A crucial new ingredient in the argument is the distinction between micro- and
macrostates. Note in particular that where in the previous work the distribution function f
was identified with a probability (namely of a molecular state), in the present paper it, or
its discrete analogy Ζ is a description of the macrostate of the gas. Probabilities are not
assigned to the particles, but to the macrostate of the gas as a whole. According to Klein
(1973, 84), this conceptual transition in 1877b marks the birth of statistical mechanics.
While this view is not completely correct (as we have seen, Boltzmann 1868 already
applied probability to the total gas), it is true that (1877b) is the first occasion where
Boltzmann identifies probability of a gas state with relative volume in phase space, rather
than its relative time of duration.

Another novelty is that Boltzmann has changed his concept of equilibrium. Whereas
previously the essential characteristic of an equilibrium state was always that it is
stationary, in Boltzmann's new view it is conceived as the macrostate (i.e., a region in
phase space) that can be realized in the largest number of ways. As a result, an
equilibrium state need not be stationary: in the course of time, the system may fluctuate
in and out of equilibrium.

5. Some later work


5.1 Return of the ergodic hypothesis

As we have seen, the 1877 papers introduced some conceptual shifts in Boltzmann’
approach. Accordingly, this year is frequently seen as a watershed in Boltzmann's
thinking. Concurrent with that view, one would expect his subsequent work to build on
his new insights and turn away from the themes and assumptions of his earlier papers.
Actually, Boltzmann's subsequent work in gas theory in the next decade and a half was
predominantly concerned with technical applications of his 1872 Boltzmann equation, in
particular to gas diffusion and gas friction. And when he did touch on fundamental
aspects of the theory, he returned to the issues and themes raised in his 1868–1871
papers, in particular the ergodic hypothesis and the use of ensembles.

This step was again triggered by a paper of Maxwell, this time one that must have
pleased Boltzmann very much, since it was called "On Boltzmann's theorem" (Maxwell
1879) and dealt with the theorem discussed in the last section of his (1868). He pointed
out that this theorem does not rely on any collision assumption. But Maxwell also made
some pertinent observations along the way. He is critical about Boltzmann's ergodic
hypothesis, pointing out that "it is manifest that there are cases in which this does not take
place" (Maxwell 1879, 694). Apparently, Maxwell had not noticed that Boltzmann's later
papers had also expressed similar doubts. He rejected Boltzmann'a time-average view of
probability and instead preferred to interpret ρ as an ensemble density. Further, he states
that any claim that the distribution function obtained was the unique stationary
distribution "remained to be investigated" (Maxwell 1879, 722). Maxwell's paper seems
to have revived Boltzmann's interest in the ergodic hypothesis, which he had been
avoiding for a decade. This renewed confidence is expressed, for example in Boltzmann
(1887):

Under all purely mechanical systems, for which equations exist that are analogous to the
so-called second law of the mechanical theory of heat, those which I and Maxwell have
investigated … seem to me to be by far the most important. … It is likely that thermal
bodies in general are of this kind [i.e., they obey the ergodic hypothesis]

However, he does not return to this conviction in later work. His Lectures on Gas Theory
(1896,1898), for example, does not even mention the ergodic hypothesis.
5.2 Return of the reversibility objection

The first occasion on which Boltzmann returned to the reversibility objection is in


(1887b). This paper delves into a discussion between Tait and Burbury about the
approach to equilibrium for a system consisting of gas particles of two different kinds.
The details of the debate need not concern us, except to note that Tait raised the
reversibility objection to show that taking any evolution approaching equilibrium one
may construct, by reversal of the velocities, another evolution moving away from
equilibrium. At this point Boltzmann entered the discussion:
I remark only that the objection of Mr. Tait regarding the reversal of the direction of all
velocities, after the special state [i.e., equilibrium] has been reached, […] has already
been refuted in my [(1877a)]. If one starts with an arbitrary non-special state, one will get
[…] the to special state (of course, perhaps after a very long time). When one reverses the
directions of all velocities in this initial state, then, going backwards, one will not (or
perhaps only during some time) reach states that are even further removed from the
special state; instead, in this case too, one will eventually again reach the special state.
(WA III, 304)

This reply to the reversibility objection uses an entirely different strategy from his
(1877a). Here, Boltzmann does not exclude the reversed motions on account of their
vanishing probability, but rather argues that, sooner or later, they too will reach the
equilibrium state.

Note how much Boltzmann's strategy has shifted: whereas previously the idea was that a
gas system should approach equilibrium because of the H-theorem; Boltzmann's idea is
now, apparently, that regardless of the behavior of H as a function of time, there are
independent reasons for assuming that the system approaches equilibrium. Boltzmann's
contentions may of course very well be true. But they do not follow from the H-theorem,
or by ignoring its exceptions, and would have to be proven otherwise.

5.3 The debate in Nature

The 1890s brought three major events in Boltzmann's work on statistical physics. The
foremost of these was his participation in the 1894 meeting of the British Association for
the Advancement of Science (BAAS) in Oxford, where issues in gas theory were lively
debated. Another was his debate with Zermelo, in 1896–1897. The third was the
appearance of his two-volume book Lectures in Gas Theory in 1896 and 1898.

The BAAS meeting had a lively aftermath: A discussion between half a dozen authors in
the columns of Nature in the years 1894-1895. The central topic of this debate was the
paradoxical relation between the H-theorem and the reversibility objection.

The kick-off was an letter by Culverwell (1894) with the innocently-sounding question,
"Will anyone say exactly what the H-theorem proves?" The question triggered several
responses. Culverwell noted, with some amusement, in a subsequent letter (1895) that
they all seemed to argue quite differently. Perhaps the most important responses were
Burbury (1894a) and (1894b) and Bryan (1894). Burbury argued that the H-theorem rests
on an additional assumption, which was independent of mechanical theory, which he
coined "Condition A". His actual phrasing of the condition was somewhat involved. In
any case, Burbury, argued out that the reversed motion would not satisfy Condition A,
and that the H-theorem would thus fail to be applicable. He thus succeeded in finally
illuminating the logical situation between the reversibility objection and the H-theorem.
The theorem assumed the validity of a condition, which would be violated in the reversed
motion; i.e., Condition A itself already contained a time-asymmetrical element. Further,
given the time-reversal invariance of the mechanical laws governing the system, any
motivation for the condition would have to come from beyond these laws and must thus
be non-mechanical.

Although Burbury had thus succeeded in clarifying the logic of the H-theorem, his actual
formulation of Condition A was not particularly transparent. By contrast, Bryan's
contribution to the debate may be said to possess the opposite qualities. He was the first
to clearly point out that the SZA itself is not invariant under time-reversal. Less
convincing is his argument that the time-reversed assumption would suppose the
molecules possessing some kind of foresight.

Finally Boltzmann himself intervened in the debate (Boltzmann 1895). He recognized, of


course, that the same issues that he discussed with Loschmidt 20 years earlier were again
at stake, and elucidated:

My minimum theorem as well as the so-called Second Law of Thermodynamics are only
theorems of probability (WA III, 539).

It can never be proved from the equations of motion alone, that the minimum function H
must always decrease. It can only be deduced from the laws of probability, that if the
initial state is not specially arranged for a certain purpose, but haphazard governs freely,
the probability that H decreases is always greater than that it increases. (WA III, 540)

In more detail, his argument is as follows. Consider a gas in a vessel with perfectly
smooth and elastic walls, in an arbitrary initial state and let it evolve in the course of
time. At each time t we can calculate H(t). Now draw a graph of this function: the H-
curve. (In a later discussion, Boltzmann 1897, he actually produced a diagram.)

Barring all cases in which the motion is ‘regular’, e.g., when all the molecules move in
one plane, Boltzmann claims the following properties of the curve:

i. For most of the time, H(t) will be very close to its minimum value, say Hmin.
ii. Because greater values of H are improbable but not impossible, the curve will
occasionally, but very rarely, rise to a peak or summit, that may be well above
Hmin.
iii. Higher summits are extremely less probable then lower summits.
Suppose that, at some time t = 0, the function attain a certain value H0, well above the
minimum value. Now, Boltzmann says, two[13] cases are possible for the evolution of H in
time. (a) H0 lies at or near the top of a peak. Then H(t) will decrease, whether we move
away in either the positive are negative time direction. (b) H0 lies on an ascending or
descending part of the curve, so that H(t) decreases or increases. But, because of (iii),
case (a) is much more probable than case (b). Hence, Boltzmann says,

… if we choose an ordinate of given magnitude H0 guided by haphazard in the curve, it


will not be certain but very probable that the ordinate decreases if we go in either
direction. (WA III, 540)

And,

What I have proved in my papers is as follows: It is extremely probable that H is very


near to its minimum value; if it is greater, it may increase or decrease, but the probability
that it decreases is always greater. (WA III, 541)

Together, the claims (i)–(iii) constitute a statement of what the Ehrenfests called a
‘statistical H-theorem’.

After having thus elucidated the content of his theorem, Boltzmann addressed the
reversibility objection. Suppose the gas is initially in a non-equilibrium state, with a large
value of H. Then it is probable, but certain, that it will decrease, and eventually reach its
minimum

If at an intermediate stage we reverse all velocities we get an exceptional state where H


increases for a certain time and decreases again. But the existence of such cases does not
disprove our theorem. On the contrary the theory of probability itself shows that the
probability of such cases is not mathematically zero, only extremely small. (WA III, 541)

It is not immediately clear how this refutes the reversibility objection. If we focus on the
fact that the reversed state is "exceptional", and take it to belong to those cases ("regular
motions") that were explicitly barred in the statement of the theorem, then his reply is
like that to Loschmidt. But perhaps his central idea here is rather that in the reversed
motion, H increases for a short while, but will eventually decrease again, and continue to
do so. In that case his defence is more like the reply to Tait. In actual fact the two are
hard to combine. If we admit that the reversed state is exceptional in the sense that the
statistical H-theorem does not hold for it, we have no grounds for claiming that
eventually H will decrease again.

This particular article is often cited as one of the clearest expositions that Boltzmann ever
wrote, and one in which "he is right on the money" (Lebowitz 1999). Still there are a few
thing that deserve attention.

1. It seems to me that Boltzmann here adopts an (implicit) identification of probability


and relative duration, At the very least, he does not indicate any difference in meaning
between phrases like "During the greater part of that time, H will be very near to its
minimum" and "It is extremely probable that H lies near its minimum". Also, it seems
that to me that his claim (ii) that peaks in the curve are "very rare", and his claim (iii) that
they are "very improbable" are intended as synonymous. This way of reading Boltzmann
would indeed not fall out of tune with his previous interpretations of probability.

2. But a larger problem looms. Boltzmann says that the statistical H-theorem is
something he had "proved in his papers". But that is surely a gross overstatement. No
proof of the claims (i)–(iii) for a gas system is to be found anywhere.

The most well-known view to the problem of how to explain this lacuna is, again,
supplied by the Ehrenfests. In essence, they suggest that Boltzmann somehow silently
relied on the ergodic hypothesis.

It is indeed evident that if the ergodic hypothesis holds, a state will spend time in the
various regions of phase space in proportion to their volume. That is to say, during the
evolution of the system along its trajectory, regions with a small volume, are visited only
sporadically, and regions with larger volume more often.

This would make it plausible how Boltzmann could identify probabilities with relative
times. But also, it would make it plausible that if a system starts out from a very small
region (an improbable state) it will display a ‘tendency’ to evolve towards the
overwhelmingly larger equilibrium state. Of course, this ‘tendency’ would have to be
interpreted in a qualified sense: the same ergodic hypothesis would imply that the system
cannot stay inside the equilibrium state forever and thus there would indeed be incessant
fluctuations in and out of equilibrium. Indeed, one would have to state that the tendency
to evolve from improbable to probable states is itself a probabilistic affair: as something
that holds true for most of the initial states, or for most of the time, or as some type of
expectation value or another. In short, we would then hopefully obtain some statistical
version of the H-theorem. Exactly how this statistical H-theorem should be formulated
remains an open problem in the Ehrenfests' point of view. Indeed they distinguish
between several possible statistical interpretations of the theorem.

The Ehrenfests' reading of Boltzmann's intentions has thus some undeniable advantages.
However, there is no evidence that Boltzmann really had the ergodic hypothesis in mind.
It seems m more likely that he relied on a naive identification of the various meanings of
probability. Further, nobody has ever succeeded in proving a statistical H-theorem on the
basis of the ergodic hypothesis, or on the basis of a modern relative such as the
hypothesis of ‘metrical transitivity’. Much stronger dynamical and probabilistic
conditions seem to be needed for such a proof, and even then it remains a difficult
problem whether these are satisfied in a realistic gas model.

Bibliography
In the foregoing and below, "WA" refers to:
Boltzmann, L. (1909), Wissenschaftliche Abhandlungen, Vol. I, II, and III, F. Hasenöhrl
(ed.), Leipzig: Barth; reissued New York: Chelsea, 1969.

Primary Sources

• 1866, Über die Mechanische Bedeutung des Zweiten Hauptsatzes der


Wärmetheorie, Wiener Berichte, 53: 195–220; in WA I, paper 2.
• 1868, Studien über das Gleichgewicht der lebendigen Kraft zwischen bewegten
materiellen Punkten, Wiener Berichte, 58: 517–560, 1868; page reference is to
WA I, paper 5.
• 1871a, Über das Wärmegleichgewicht zwischen mehratomigen Gasmolekülen,
Wiener Berichte, 63: 397–418; in WA I, paper 18.
• 1871b, Einige allgemeine Sätze über Wärmegleichgewicht, Wiener Berichte, 63:
679–711; in WA I, paper 19.
• 1871c, Analytischer Beweis des zweiten Haubtsatzes der mechanischen
Wärmetheorie aus den Sätzen über das Gleichgewicht der lebendigen Kraft,
Wiener Berichte, 63: 712–732; in WA I, paper 20.
• 1872, Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen,
Wiener Berichte, 66: 275–370; in WA I, paper 23.
• 1877a, Bermerkungen über einige Probleme der mechanische Wärmetheorie,
Wiener Berichte, 75: 62–100; in WA II, paper 39.
• 1877b, Über die beziehung dem zweiten Haubtsatze der mechanischen
Wärmetheorie und der Wahrscheinlichkeitsrechnung respektive den Sätzen über
das Wärmegleichgewicht, Wiener Berichte, 76: 373–435; in WA II, paper 42.
• 1881, Referat über die Abhandlung von J.C. Maxwell: "Über Boltzmann's
Theorem betreffend die mittlere verteilung der lebendige Kraft in einem System
materieller Punkte", Wied. Ann. Beiblätter, 5: 403–417; in WA II, paper 63.
• 1884, Über die Eigenschaften Monocyklischer und andere damit verwandter
Systeme, Crelles Journal, 98: 68–94; in WA III, paper 73.
• 1887, Neuer Beweis zweier Sätze über das Wärmegleichgewicht unter
mehratomigen Gasmolekülen, Wiener Berichte, 95: 153–164; in WA III, paper
83.
• 1887b, Über einige Fragen der Kinetische Gastheorie, Wiener Berichte, 96: 891–
918; in (WA III, paper 86.
• 1892, III. Teil der Studien über Gleichgewicht der lebendigen Kraft, Münch. Ber.,
22: 329–358; in WA III, paper 97.
• 1895, On certain questions in the theory of gases, Nature, 51: 413–415; in WA
III, paper 112.
• 1895b, On the minimum theorem in the theory of gases, Nature, 52: 221; in WA
III, paper 114.
• 1896, Vorlesungen über Gastheorie: Vol I, Leipzig, J.A. Barth; translated together
with Volume II, by S.G. Brush, Lectures on Gas Theory, Berkeley: University of
California Press, 1964.
• 1896b, Entgegnung an die wärmetheoretischen Betrachtungen des Hrn. E.
Zermelo, Wied. Ann., 57: 772–784; in WA III, paper 119.
• 1897a, Zu Hrn Zermelos Abhandlung "Über die mechanische Erklärung
irreversibler Vorgänge", Wied. Ann., 60: 392–398, in WA III, paper 120.
• 1897b, Über einige meiner weniger bekannte Abhandlungen über Gastheorie und
deren Verhältnis zu derselben, Verh. des 69. Vers. Deutschen Naturf. und Ärzte,
Braunschweig, 19–26; in WA III, paper 123
• 1897c, Über die Unentbehrlichkeit der Atomistik in der Naturwissensschaft,
Annalen der Physik und Chemie, 60: 231; in Boltzmann 1905, pp. 78–93.
• 1897d, Über die Frage der objektiven Existenz der Vörgange in der unbelebten
Natur, Wiener Berichte, 106: 83; in Boltzmann 1905, pp. 94–119.
• 1898a, Vorlesungen über Gastheorie: Vol II, Leipzig, J.A. Barth; translated
together with Volume I, by S.G. Brush, Lectures on Gas Theory, Berkeley:
University of California Press, 1964.
• 1898b, Über die sogenannte H-Kurve, Math. Ann., 50: 325–332; in WA III, paper
127.
• 1904, (with J. Nabl), Kinetische theorie der Materie, Encyclopädie der
Mathematischen Wissenschaften, Vol V/1, pp. 493–557.
• 1905, Populäre Schriften, Leipzig: J.A. Barth; re-issued Braunschweig: F.
Vieweg, 1979.

Secondary Sources

• Blackmore, J. (1982), "Boltzmann's concessions to Mach's philosophy of


science", in Ludwig Boltzmann Gesamtausgabe: Ausgewählte Abhandlungen,
Band 8, R. Sexl and J. Blackmore (eds), Graz: Akademische Druck- und
Verlagsanstalt.
• Blackmore, J. (1995), Ludwig Boltzmann: His Later Life and Philosophy, 1900–
1906 (Dordrecht: Kluwer).
• Blatt, J.M (1959), "An alternative approach to the ergodic problem", Progress of
Theoretical Physics, 22: 745-756.
• Bricmont, J. (1996), "Science of Chaos and Chaos in Science", in The Flight from
Science and Reason, P.R. Gross, N.Levitt and M.W.Lewis (eds), New York
Academy of Science.
• Brush, S.G., (1966), Kinetic Theory, Vol. 1 and 2., Oxford, Pergamon.
• Brush, S.G., (1976), The Kind of Motion We Call Heat, Amsterdam, North
Holland.
• Bryan, G.H. (1894), "Letter to the editor", Nature, 51: 175.
• Burbury, S.H. (1894a), "Boltzmann's minimum theorem", Nature, 51: 78.
• Burbury, S.H. (1894b), "The kinetic theory of gases", Nature, 51: 175.
• Cercignani, C. (1998), Ludwig Boltzmann, the man who trusted atoms, Oxford:
Oxford University Press.
• Cohen, E.G.D., (1996), Boltzmann and Statistical Mechanics, in Boltzmann's
Legacy 150 Years After His Birth, Atti dei Convegni Lincei, 131: 9–23, Rome:
Accademia Nazionale dei Lincei. [Preprint available online]
• Culverwell, E.P., (1894), "Dr. Watson's proof of Boltzmann's theorem on
permanence of distributions", Nature, 50: 617.
• Culverwell, E.P. (1895), "Boltzmann's minimum theorem", Nature, 51: 246.
• Deltete, R. (1999), "Helm and Boltzmann. Energetics at the Lübeck
Naturforscherversammlung", Synthese, 119: 45–68.
• Dugas, R. (1959), La théorie physique au sens de Boltzmann et ses
prolongements modernes, Neuchâtel: Griffon.
• Ehrenfest, P. and T. Ehrenfest-Afanassjewa (1912), The Conceptual Foundations
of the Statistical Approach in Mechanics, New York: Cornell University Press,
1959; translation of “Begriffliche Grundlagen der statistischen Auffassung in der
Mechanik”, in Encyklopädie der mathematischen Wissenschaften, Volume
IV/2/II/6, Leipzig: B. G. Teubner, 1912.
• Gallavotti, G. (1994), "Ergodicity, ensembles, irreversibility in Boltzmann and
beyond", Journal of Statistical Physics, 78: 1571–1589. [Preprint available
online.]
• Garber, E. S.G. Brush and C.W.F. Everitt (eds.) (1986), Maxwell on Molecules
and Gases, MIT Press, Cambridge Mass.
• Garber, E. S.G. Brush and C.W.F. Everitt (eds.) (1995), Maxwell on Heat and
Statistical Mechanics, Bethlehem: Lehigh University Press.
• Gibbs, J.W., (1902), Elementary Principles in Statistical Mechanics, New York:
Scribner.
• Goldstein, S. (2001), "Boltzmann's approach to statistical mechanics," in Chance
in Physics: Foundations and Perspectives, in J. Bricmont et al. (eds), Lecture
Notes in Physics, 574, Springer-Verlag. [Preprint available online.]
• Höflechner, W. (ed) (1994), Ludwig Boltzmann Leben und Briefe, (Graz:
Akademische Druck- und Verlagsanstalt.
• Kirsten, C. and H.-G. Körber, (1975), Physiker über Physiker, Berlin: Akademie-
Verlag.
• Klein, M.J., (1973), "The Development of Boltzmann's Statistical Ideas", in
E.G.D. Cohen and W. Thirring (eds.), The Boltzmann Equation, Wien: Springer,
pp. 53–106.
• Lebowitz, J.L., (1999), "Statistical mechanics: A selective review of two central
issues", Reviews of Modern Physics, 71: S346-S357. [Preprint available online.]
• Lindley, D. (2001), Boltzmann's Atom, New York: Free Press.
• Loschmidt, J., (1876/1877), "Über die Zustand des Wärmegleichgewichtes eines
Systems von Körpern mit Rücksicht auf die Schwerkraft", Wiener Berichte, 73:
128, 366 (1876); 75: 287; 76: 209 (1877).
• Maxwell, J.C. (1860), "Illustrations of the Dynamical Theory of Gases",
Philosophical Magazine, 19: 19–32; 20: 21–37; in Garber, Brush, & Everett 1995,
285–318.
• Maxwell, J.C. (1867), "On the dynamical theory of gases", Philosophical
Transactions of the Royal Society of London, 157: 49–88; in Brush 1966 and
Garber, Brush, & Everett 1986, 419–472.
• Maxwell, J.C., (1879), "On Boltzmann's theorem on the average distribution of
energy in a system of material points", Trans. Cambridge Phil. Soc., 12: 547–570;
in Garber, Brush, & Everett 1995, 357–386.
• Nemytskii, V.V., and V.V. Stepanov (1960), Qualitative Theory of Differential
Equations, Princeton: Princeton University Press.
• Plato, J. von (1992), "Boltzmann's ergodic hypothesis", Arch. Hist. Exact Sci. 44:
71.
• Plato, J. von, (1994), Creating Modern Probability, Cambridge: Cambridge
University Press.
• Regt, H. de (1999), "Ludwig Boltzmann's Bildtheorie and scientific
understanding", Synthese, 119: 113–134.
• Ridderbos, T.M and Redhead, M.L.G. (1998), "The spin-echo experiment and the
second law of thermodynamics", Foundations of Physics, 28: 1237–1270.
• Sklar, L. (1993) Physics and Chance. Philosophical Issues in the Foundations of
Statistical Mechanics, Cambridge: Cambridge University Press.
• Truesdell, C. (1961), "Ergodic theory in classical statistical mechanics", in P.
Caldirola (ed.), Ergodic Theories, New York: Academic Press, pp. 21–56.
• Zermelo, E. (1896a), "Über einen Satz der Dynamik und die mechanische
Wärmetheorie", Annalen der Physik, 57: 485–494.
• Zermelo, E. (1896b), "Über mechanische Erklärungen irreversibler Vorgänge",
Annalen der Physik, 59: 793–801.

Other Internet sources


• Ludwig Boltzmann: Bibliographie, an extensive bibliography of Boltzmann's
articles maintained by the Universitätsbibliotek at the Universität Heidelberg
• Electronics Research Archive for Mathematics (in German), Search on
"Boltzmann" to get access to the abstracts of many of Boltzmann's papers
• Biographical information about Joseph Loschmidt
• Biographical information about Ostwald

You might also like