Equilibrium Statistical Mechanics

  • Chapter
  • First Online:
Making Sense of Statistical Mechanics

Part of the book series: Undergraduate Lecture Notes in Physics ((ULNP))

  • 1783 Accesses

Abstract

We present here Boltzmann’s ideas concerning the microscopic basis of equilibrium statistical mechanics. It is based on the distinction between the microstates and the macrostates of a system and the fact that the number of microstates corresponding to a given value of a macrostate depends very much on that value. We also give the formula for the microscopic or Boltzmann’s entropy and then give the corresponding microscopic formulas for the free energy and the grand potential. We define the various ensembles or probability distributions on sets of microstates and show that they are equivalent in the limit of large systems.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
EUR 29.95
Price includes VAT (Germany)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
EUR 50.28
Price includes VAT (Germany)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
EUR 64.19
Price includes VAT (Germany)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    This follows from the formula (3.3.1) for the energy, assuming the potential function to be bounded from below.

  2. 2.

    \(\Omega _{eq}\) coincides with what we called \(G_N(\epsilon )\) in (2.3.2).

  3. 3.

    In fact, one gets, from Stirling’s formula (6.A.3), that

    $$ \frac{|\{\omega \in \Omega \ | n_0= \frac{1}{2} \}|}{|\Omega |}= \mathcal {O} \left( \frac{1}{\sqrt{N}}\right) ; \nonumber $$

    Indeed, use (6.2.1), with \(N_0=\frac{N}{2}\), assuming N to be even. Then, since \(|\Omega |=2^N\), (6.A.3) gives:

    $$\begin{aligned} \frac{|\{\omega \in \Omega \ | n_0= \frac{1}{2} \}|}{|\Omega |}= \frac{N^N\exp (-N)\sqrt{2\pi N } }{2^N \left( \frac{N}{2 }\right) ^N \exp (-N)(\sqrt{\pi N})^2} = \mathcal {O} \left( \frac{1}{\sqrt{N}}\right) . \end{aligned}$$
    (6.2.7)
  4. 4.

    In [211] Lavis criticizes the approach outlined here, in particular the way I presented it in [53], by emphasizing the necessity of grou** macrostates close to the equilibrium value, but nobody ever denied that and the law of large numbers always speaks of values close to the average one, not exactly equal to it, see (2.3.2), (2.3.7), (2.3.11). See Lazarovici and Reichert [212, Sect. 5.3] for more discussion of Lavis’ views.

  5. 5.

    To get this formula, list all the permutations of the N particles; put the \(N_1\) first particles in state 1, the next \(N_2\) particles in state 2 etc. Thus we get N! permutations the N particles to be divided by the number \(\prod _{i=1}^L N_i!\) of permutations of the \(N_i\) particles in each group \(i=1, \ldots , L\), since permutations in those groups do not change the final distribution of the particles among states.

  6. 6.

    This can be shown by using Lagrange’s multipliers, as one does below, but with only the constraint (6.2.14) and not the constraint (6.2.15).

  7. 7.

    We denote \(\beta \) one of the multipliers because it will be shown later to coincide with the \(\beta =\frac{1}{kT}\) of thermodynamics.

  8. 8.

    That formula is inscribed on Boltzmann’s tomb in Vienna, with \(\Omega \) replaced by W, W for Wahrscheinlichkeit (probability in German), but it was first written in that form by Max Planck, who however credited the idea to Boltzmann.

  9. 9.

    This comes from his course notes on Statistical Physics, available on his webpage http://www.ge.infn.it/~zanghi/. It is quoted by Tumulka [314, Sect. 5.4], who adds: the clarification of the status of identical particles, both classical and quantum, is mainly due to Leinaass and Myrheim [228].

  10. 10.

    The gamma function is a generalization of the factorial: for \(z\in \mathbb C\), \(\Re z>0\),

    $$\begin{aligned} \Gamma (z)=\int _0^\infty x^{z-1} e^{-x} dx \end{aligned}$$
    (6.4.4)

    and for \(z=n\), \(n \in \mathbb N\), we have \(\Gamma (n)=(n-1)!\), see Appendix 6.A.2.

  11. 11.

    See Appendix 6.A.1 for a discussion of that method and the justification of (6.5.4) below.

  12. 12.

    We shall see in Chap. 7 that this definition coincides with the one of the Shannon entropy.

  13. 13.

    To see the connection with (2.A.20), take \(E(\mathbf{x})\) as the random variable F, \(\mu =\nu _m\), the micro-canonical probability distribution, and \(\nu \) the canonical distribution \(\nu _c\) given by (6.6.16).

  14. 14.

    Meaning the Boltzmannian.

  15. 15.

    My paraphrase.

  16. 16.

    We will also give an intuitive version of the entropy formula in Appendix 6.C.

  17. 17.

    For a detailed mechanical derivation of that formula, see e.g. Amit and Verbin [12, p. 231].

  18. 18.

    See Albert [9, Chap. 3] for an elaboration of this argument and criticisms of the other justifications given below.

  19. 19.

    This happens not only in the examples of Sect. 4.6.2 but also for more general “chaotic” dynamical systems for which \(\nu \) is a Sinai-Ruelle-Bowen measure and the set \(\mathcal A\) is a “strange attractor”, see e.g. Bowen [45], Bowen and Ruelle [46], Eckmann and Ruelle [120], Ruelle [281, 282], Sinai [295].

  20. 20.

    See Allori [11, Sect. 3.4.2] for a criticism of this observation.

  21. 21.

    Roderich Tumulka has pointed out to me (private communication) that non additive “measures” (that is, positive valued maps defined on families of sets that do not satisfy (2.2.1) and thus are not real measures but that may satisfy other properties than additivity), could be relevant for certain approaches to quantum mechanics and that a notion of typicality, distinct from the one of probability, may be relevant for statistical mechanics on unbounded phase spaces on which there is no natural probability measure.

  22. 22.

    The energy of exercise 6.6 is, with a suitable choice of constants, the energy of N quantum harmonic oscillators; the limit when \(e\rightarrow \infty \) for quantum harmonic oscillators coincides with the classical value obtained here.

  23. 23.

    This is the basis of Debye’s model of crystals, see footnote 7.9.

  24. 24.

    See Callen [66, Sect. 16-7] for the justification of this approximation.

  25. 25.

    Which the reader may derive by expanding \(\frac{1}{e^x -1}= \sum _{n=1}^\infty e^{-nx}\) and using \(\int _0^{\infty }{x^3}{e^{-x}} dx= 3 !\) and \(\sum _{n=1}^\infty \frac{1}{n^4}= \frac{\pi ^4}{90}\).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean Bricmont .

Appendices

Appendix

6.A Asymptotics

6.1.1 6.A.1 Laplace’s Method

Consider integrals of the form

$$\begin{aligned} \int _0^\infty \exp (\lambda f(x))dx \end{aligned}$$
(6.A.1)

for \(\lambda \) large and positive. Assume that \(f(x)\rightarrow -\infty \) as \(x\rightarrow \infty \), sufficiently last so that the integral converges for all \(\lambda >0\). Assume also that the function f has a unique maximum at \(x=x_0\), \(0<x_0<\infty \). Write that integral as:

$$\begin{aligned}&\exp (\lambda f(x_0)) \\&\Big ( \int _{|x-x_0|\le \epsilon } \exp (\lambda (f(x)-f(x_0))dx +\int _{|x-x_0|> \epsilon } \exp (\lambda (f(x)-f(x_0))dx \Big ), \end{aligned}$$

for some \(\epsilon >0\). Since, \(\forall x, |x-x_0|> \epsilon \), \(f(x)-f(x_0)< \delta \) for some \(\delta >0\) (since \(x_0\) is the unique maximum of f), and the integral in (6.A.1) converges, we have that the second integral in (6.A.1) is bounded by \(\exp (-c\lambda )\), for some \(c>0\).

When \(|x-x_0| \le \epsilon \), we approximate \(f(x)-f(x_0)\approx -\frac{a(x-x_0)^2}{2}\), with \(-a= f''(x_0)\); since \(x_0\) is a strict maximum of f, we have \(f'(x_0)=0\) and \(a>0\). Changing variables \(y=\sqrt{\lambda a}(x-x_0)\) in \(\int _{|x-x_0|\le \epsilon } \exp (-\lambda \frac{a(x-x_0)^2}{2})dx\), we approximate that latter integral by \( \frac{1}{\sqrt{\lambda a}}\int _\mathbb R\exp (-\frac{y^2}{2})dy= \sqrt{\frac{2\pi }{\lambda a}}\) (we replace \(\pm \lambda \epsilon \) by \(\pm \infty \) in the limits of integration).

Altogether, we get the approximation:

$$\begin{aligned} \int _0^\infty \exp (\lambda f(x))dx \approx \exp (\lambda f(x_0)) \sqrt{\frac{2\pi }{\lambda a}}= \exp (\lambda f(x_0)) \sqrt{\frac{2\pi }{\lambda |f''(x_0)| }}, \end{aligned}$$
(6.A.2)

as \(\lambda \rightarrow \infty \). With more work, one can obtain much more detailed estimates than (6.A.2), see e.g. Erdelyi [123].

6.1.2 6.A.2 Stirling’s Formula

Laplace’s method can be used to justify the approximation to the factorial function \(N!\approx N^N e^{-N}\) as \(N \rightarrow \infty \). First write:

$$ N!= \int _0^\infty x^N \exp (-x) dx \nonumber $$

which can be proven recursively from \(\int _0^\infty x^n \exp (-x) dx=n \int _0^\infty x^{n-1} \exp (-x) dx\) (which follows from integration by parts) and \(\int _0^\infty \exp (-x) dx=1\). Let \(x=Ny\); we get:

$$ \int _0^\infty x^N \exp (-x) dx= N^{N+1} \int _0^\infty y^N \exp (-Ny) dy= N^{N+1} \int _0^\infty \exp (N(\ln y-y)) dy \nonumber $$

The function \(f(y)=\ln y-y\) has a unique maximum at \(y=1\) with \(f''(1)=-1\), so that applying (6.A.2) gives:

$$\begin{aligned} N! \approx N^{N+1} \exp (-N) \sqrt{\frac{2\pi }{N }}= N^{N} \exp (-N) \sqrt{2\pi N } \end{aligned}$$
(6.A.3)

One can check that the same approximation holds for the Gamma function (6.4.4).

6.B “Derivation” of Formula (6.5.4)

In order to apply Laplace’s method to the integral (6.5.2), we need to verify that the function \(f(e)= -\beta e+ \frac{s(e, v)}{k}\) has a unique maximum, which follows from the concavity of the function s(ev), and check that the integral converges. That follows from the fact that \(f(e)\approx -\beta e\), as \(e\rightarrow \infty \), since \(s(e, v)\approx \ln e\) as \(e\rightarrow \infty \), see (6.4.7) (the latter computation is valid only for an ideal gas but the logarithmic growth of s(ev) as a function of e when \(e\rightarrow \infty \) holds more generally).

6.C An Intuitive Formula for the Entropy

An intuitive expression for the variation of entropy is most easily expressed in the situation of discrete variables.

Consider, as we did after (6.2.11), a system of N particles, each of which can be in L possible states, but with an “energy” variable \(e_i (\lambda )\), \(i=1, \ldots , L\), associated to each such state and depending on an external parameter \(\lambda \), with the sum of the energies being fixed: \(E(\mathbf{x}, \lambda )=\sum _{i= 1}^N e(x_{i}, \lambda )\).

The set of microstates is \(\Omega = \{1, \ldots , L\}^N\), and we can take as macrostates the fractions of particles in each state: \(\mathbf{n}= (n_1, \ldots , n_L)\), where \(n_i= \frac{N_i}{N}\) with \(N_i\) being the number of particles in state i. In terms of the macrostate, the total energy equals: \(E(\mathbf{x}, \lambda )= \sum _{i= 1}^L N_i e_i(\lambda )\) and \(\frac{E(\mathbf{x}, \lambda )}{N}= \sum _{i= 1}^L n_i e_i(\lambda )\).

In the canonical ensemble, we get:

$$ E =\sum _\mathbf{x} E(\mathbf{x}, \lambda ) \frac{\exp (-\beta E (\mathbf{x}, \lambda ))}{Z}=N\sum _{i =1}^L e_i (\lambda ) n_i (\beta ,\lambda ) \nonumber $$

with \(E= \sum _{i =1}^N e(x_i, \lambda )\) and \(n_i (\beta ,\lambda )\) the fraction of states with energy \(e(x_i, \lambda )=e_i (\lambda )\), given by \(n_i (\beta ,\lambda )= \frac{\exp (-\beta e_i (\lambda ))}{Z}\). We get:

$$\begin{aligned} dE =N \Big ( \sum _{i =1}^L d e_i (\lambda ) n_i (\beta ,\lambda ) + \sum _{i =1}^L e_i (\lambda ) d n_i (\beta ,\lambda )\Big ). \end{aligned}$$
(6.C.1)

In the sum, the first term equals, using (6.7.2), (6.7.3),

$$ N\sum _{i =1}^L \frac{\partial e_i}{\partial \lambda } n_i (\beta ,\lambda ) d \lambda = \sum _\mathbf{x} \frac{\partial E(\mathbf{x}, \lambda )}{\partial \lambda } \frac{\exp (-\beta E (\mathbf{x}, \lambda ))}{Z} d \lambda =-dW. \nonumber $$

Therefore, we get from (6.7.1):

$$\begin{aligned} Tds = \frac{T}{N}dS = \sum _{i =1}^L e_i (\lambda ) d n_i (\beta ,\lambda ). \end{aligned}$$
(6.C.2)

So we see that the variation of entropy is given (up to the factor T) by the variation of the occupation numbers of the different energy levels, while the work is given by the variation of the individual energies. If no work is done (\(\lambda \) is constant) then we can still have a variation of the total energy through heat transfer and that translates microscopically into a variation of the occupation numbers \(n_i\)’s: the occupation numbers of higher energy levels can increase or decrease (depending on the direction of the heat transfer) and that leads to a change of entropy.

One can also obtain directly (6.C.2) from the formula (6.7.7)

$$Tds =\frac{1}{N}\left( kT\frac{\partial \ln Z}{\partial \lambda } -\frac{\partial ^2 \ln Z}{\partial \beta \partial \lambda }\right) d\lambda - \frac{\partial ^2 \ln Z}{\partial ^2 \beta } d\beta $$

and \(n_i (\beta ,\lambda )= \frac{\exp (-\beta e_i (\lambda ))}{Z}\):

We have:

$$\frac{1}{N}kT\frac{\partial \ln Z}{\partial \lambda }= -<\frac{\partial e (\lambda )}{\partial \lambda }> $$

and

$$ \frac{1}{N}\frac{\partial ^2 \ln Z}{\partial \beta \partial \lambda }= -<\frac{\partial e (\lambda )}{\partial \lambda }> + \beta \left(<\frac{\partial e (\lambda )}{\partial \lambda } e (\lambda )> -<\frac{\partial e (\lambda )}{\partial \lambda }> <e (\lambda )>\right) . \nonumber $$

Also,

$$ \frac{1}{N}\frac{\partial ^2 \ln Z}{\partial ^2 \beta } =<e^2(\lambda )>- <e(\lambda )>^2. $$

So,

$$\begin{aligned} \begin{aligned}&Tds = -<\frac{\partial e (\lambda )}{\partial \lambda }> d\lambda +<\frac{\partial e (\lambda )}{\partial \lambda }> d\lambda \\&- \beta \left(<\frac{\partial e (\lambda )}{\partial \lambda } e (\lambda )> -<\frac{\partial e (\lambda )}{\partial \lambda }> <e (\lambda )>\right) d\lambda \\&- (<e(\lambda )^2>-<e(\lambda )>^2) d\beta \\&=-\beta \left(<\frac{\partial e (\lambda )}{\partial \lambda } e (\lambda )> -<\frac{\partial e (\lambda )}{\partial \lambda }> <e (\lambda )>\right) d\lambda \\&- (<e(\lambda )^2>- <e(\lambda )>^2) d\beta \\ \end{aligned}\end{aligned}$$
(6.C.3)

Let us compute the right hand side of (6.C.2)

$$\begin{aligned} \sum _{i =1}^L e_i (\lambda ) d n_i (\beta ,\lambda )= \sum _{i =1}^L e_i (\lambda ) \frac{\partial n_i (\beta ,\lambda )}{\partial \beta } d\beta + \sum _{i =1}^L e_i (\lambda ) \frac{\partial n_i (\beta ,\lambda )}{\partial \lambda } d\lambda \end{aligned}$$
(6.C.4)

Computing each term in (6.C.4), we have:

$$ \sum _{i =1}^L e_i (\lambda ) \frac{\partial n_i (\beta ,\lambda )}{\partial \beta }= - (<e(\lambda )^2>- <e(\lambda )>^2),$$

and

$$ \sum _{i =1}^L e_i (\lambda ) \frac{\partial n_i (\beta ,\lambda )}{\partial \lambda }= - \beta \left(<\frac{\partial e (\lambda )}{\partial \lambda } e (\lambda )> -<\frac{\partial e (\lambda )}{\partial \lambda }> <e (\lambda )>\right) . $$

Inserting this in (6.C.4) and comparing with (6.C.3) proves (6.C.2).

Rights and permissions

Reprints and permissions

Copyright information

© 2022 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Bricmont, J. (2022). Equilibrium Statistical Mechanics. In: Making Sense of Statistical Mechanics. Undergraduate Lecture Notes in Physics. Springer, Cham. https://doi.org/10.1007/978-3-030-91794-4_6

Download citation

Publish with us

Policies and ethics

Navigation