Skip to main content

Typicality and the Role of the Lebesgue Measure in Statistical Mechanics

  • Chapter
  • First Online:
Probability in Physics

Part of the book series: The Frontiers Collection ((FRONTCOLL))

Abstract

Consider a finite collection of marbles. The statement “half the marbles are white” is about counting and not about the probability of drawing a white marble from the collection. The question is whether non-probabilistic counting notions such as half, or vast majority can make sense, and preserve their meaning when extended to the realm of the continuum. In this paper we argue that the Lebesgue measure provides the proper non-probabilistic extension, which is in a sense uniquely forced, and is as natural as the extension of the concept of cardinal number to infinite sets by Cantor. To accomplish this a different way of constructing the Lebesgue measure is applied. One important example of a non-probabilistic counting concept is typicality, introduced into statistical physics to explain the approach to equilibrium. A typical property is shared by a vast majority of cases. Typicality is not probabilistic, at least in the sense that it is robust and not dependent on any precise assumptions about the probability distribution. A few dynamical assumptions together with the extended counting concepts do explain the approach to equilibrium. The explanation though is a weak one, and in itself allows for no specific predictions about the behavior of a system within a reasonably bounded time interval. It is also argued that typicality is too weak a concept and one should stick with the fully fledged Lebesgue measure. We show that typicality is not a logically closed concept. For example, knowing that two ideally infinite data sequences are typical does not guarantee that they make a typical pair of sequences whose correlation is well defined. Thus, to explain basic statistical regularities we need an independent concept of typical pair, which cannot be defined without going back to a construction of the Lebesgue measure on the set of pairs. To prevent this and other problems we should hold on to the Lebesgue measure itself as the basic construction.

This chapter was written by Itamar Pitowsky for this volume shortly before his death. As we did not have Itamar‘s LaTex file, it had to be retyped, and proofread by us. Remaining typos are, therefore, our fault. We thank William Demopoulos for reading the chapter and streamlining some of its formulations.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 54.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Note that the number of white pixels in a picture may be considered a “macroscopic” observable, whose measurement requires no detailed knowledge of the pixel distribution. If we assume that white pixels emit light, and black pixels do not, we just measure the light emitted from a picture, and compare it with the all white picture.

  2. 2.

    This is not a very smart prior, though. It assumes independence, and therefore blocks the possibility of learning from experience.

  3. 3.

    This general result is due to Caratheodory, see [12, page 16].

  4. 4.

    The rate of convergence on the right hand side of (3.3) is better than the historical one derived by Bernoulli. We need the stronger result (essentially due to Borel) for later purposes. See [13], page 40.

  5. 5.

    The construction of \( \mathcal{B} \) is achieved by transfinite induction over the two operations, countable union and then countable intersection, all the way to the first uncountable ordinal.

  6. 6.

    Further extensions of the Lebesgue measure are possible. The validity of the strong version of the axiom of choice entails the existence of non-measurable sets, that is, \( C\, \subset \,{\left\{ {0,1} \right\}^\omega }\, \) such that \( {\hbox{C}}\, \notin \,\mathcal{L} \). We can add some of those to \( \mathcal{L} \) and extend the measure to them [14]. With this the measure is no longer regular (see below). Moreover, there are models of set theory, with weaker principles of choice, in which every subset of \( {\left\{ {0,1} \right\}^\omega } \) is Lebesgue measurable [15].

  7. 7.

    Actually \( {F_\varepsilon } = A\, \times \,\left\{ {0,1} \right\}\, \times \,\left\{ {0,1} \right\}\, \times \, \ldots \, \subseteq \,{\left\{ {0,1} \right\}^\omega }\,{\hbox{with }}\,A{ }\, \subseteq \,{\left\{ {{0,1}} \right\}^n} \) for some integer n, and we are counting the elements of A.

  8. 8.

    See [17]. In a later important dynamical extension of this result the authors adopt the typicality point of view [19].

  9. 9.

    Dimension is a topological invariant, as proved by Brouwer in 1911. Partial results concerning the non-existence of a homeomorphism between the real line and higher dimensional real spaces existed in Boltzmann’s time. For example, Lüroth in 1878.

  10. 10.

    The reason why the double sided sequence space is used is to make the Bernoulli shift well defined and invertible.

  11. 11.

    A similar point about the role of induction in statistical mechanics is made in [21].

References

  1. Lebowitz, J.L.: Boltzmann’s entropy and time’s arrow. Phys. Today 46, 32–38 (1993)

    Article  Google Scholar 

  2. Lebowitz, J.L.: Macroscopic laws, microscopic dynamics, time’s arrow and Boltzmann’s entropy. Physica A 194, 1–27 (1993)

    Article  ADS  MathSciNet  Google Scholar 

  3. Dürr, D.: Über den Zufall in der Physik. http://www.mathematik.uni-muenchen.de/~duerr/Zufall/zufall.html (1998)

  4. Goldstein, S.: Boltzmann’s approach to statistical mechanics. In: Jean, B., Dürr, D., Galavotti, M.C., Ghirardi, G.C., Petruccione, F., Zangh, N. (eds.) Chance in Physics: Foundations and Perspectives, pp. 39–54. Springer, Berlin and New York (2001)

    Chapter  Google Scholar 

  5. Goldstein, S., Lebowitz, J.L.: On the (Boltzmann) entropy of non-equilibrium systems. Physica D 193, 53–66 (2004)

    Article  MATH  ADS  MathSciNet  Google Scholar 

  6. Lavis, D.: Boltzmann and Gibbs: an attempted reconciliation. Stud. Hist. Philos. Mod. Phys. 36, 245–273 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  7. Maudlin, T.: What could be objective about probabilities? Stud. Hist. Philos. Mod. Phys. 38, 275–291 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  8. Sheldon, G., Lebowitz, J.L.: Mastrodonato Christian, Tumulka Roderich, Zanghi Nino. Normal Typicality and von Neumann’s Quantum Ergodic Theorem. Proc. R. Soc. A 466(2123), 3203–3224 (2009)

    Google Scholar 

  9. Frigg, R.: Typicality and the approach to equilibrium in Boltzmannian statistical mechanics. In: Ernst, G., Huttemann, A. (eds.) Time Chance and Reduction. Philosophical Aspects of Statistical Mechanics, pp. 92–118. Cambridge University Press, Cambridge (2010)

    Google Scholar 

  10. Uffink, J.: Compendium of the foundations of classical statistical physics. In: Butterfield, J., Earman, J. (eds.) Handbook for Philsophy of Physics, pp. 923–1047. North Holland, Amsterdam (2006)

    Google Scholar 

  11. Buzaglo, M.: The Logic of Concept Expansion. Cambridge University Press, Cambridge (2002)

    MATH  Google Scholar 

  12. Patersen, K.: Ergodic Theory. Cambridge University Press, Cambridge (1983)

    Google Scholar 

  13. Chow, Y.S., Teicher, H.: Probability Theory Independence, Interchangeability, Martingales. Springer, Berlin and New York (1978)

    MATH  Google Scholar 

  14. Hewitt, E., Ross, K.A.: Abstract Harmonic Analysis. Springer, Berlin and New York (1994)

    Book  Google Scholar 

  15. Solovay, R.M.: A model of set-theory in which every set of reals is Lebesgue measurable. Ann. Math. 92, 1 (1970)

    Article  MATH  MathSciNet  Google Scholar 

  16. von Plato, J.: Creating Modern Probability. Cambridge University Press, Cambridge (1994)

    Book  Google Scholar 

  17. Sandu, P., Short, A.J.: Winter Andreas. The foundations of statistical mechanics from entanglement: individual states vs. averages. http://arxiv.org/abs/quantph/0511225 (2005)

  18. Sheldon, G., Lebowitz, J.L., Tumulka, R., Zanghi, N.: Canonical typicality. Phys. Rev. Lett. 96, 050403 (2006)

    Article  MathSciNet  Google Scholar 

  19. Noah, L., Sandu, P., Short, A.J.: Winter Andreas. Quantum mechanical evolution towards thermal equilibrium. http://arxiv.org/abs/0812.2385 (2008)

  20. Loewer, B.: Determinism and chance. Stud. Hist. Philos. Mod. Phys. 32, 609–629 (2001)

    Article  Google Scholar 

  21. Hemmo, M., Orly, S.: Measures over initial conditions, in Y. Ben-Menahem and M. Hemmo (eds.), Probability in Physics, pp. 87–98. The Frontiers Collection, Springer-Verlag Berlin Heidelberg (2011)

    Google Scholar 

  22. Vranas, P.B.M.: Epsilon-ergodicity and the success of equilibrium statistical mechanics. Philos. Sci. 65, 688–708 (1998)

    Article  MathSciNet  Google Scholar 

  23. Rudin, W.: Principles of Mathematical Analysis. McGraw Hill, New York (1964)

    MATH  Google Scholar 

  24. Bub, J., Pitowsky, I.: Critical notice on Popper’s postscript to the logic of scientific discovery. Can. J. Philos. 15, 539–552 (1986)

    Google Scholar 

Download references

Acknowledgement

I would like to thank Meir Hemmo and Orly Shenker for their valuable advice. This research is supported by the Israel Science Foundation, grant number 744/07.

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Appendices

Appendix 1: Proof of Theorem 1

Theorem 1

Let \( E \in \mathcal{L} \) be any Lebesgue measurable set and let \( \varepsilon > 0 \); then there is \( {F_\varepsilon } \in \mathcal{F} \) such that \( \mu [(E\backslash {F_\varepsilon }) \cup ({F_\varepsilon }\backslash E)] < \varepsilon . \)

Proof

Consider first a Lebesgue measurable subset \( E \subseteq [0,1]. \) By the regularity of the Lebesgue measure (see [23], page 230) given any \( \varepsilon > 0 \) there is an open set U, with \( E \subseteq U \) and \( \mu (U\backslash E) < \frac{\varepsilon }{2}. \) The family of open intervals with dyadic endpoints forms a basis for the usual topology on [0, 1] (recall that a dyadic number is a rational whose denominator is a power of 2). Thus, we can represent U as a disjoint countable union \( U = \cup_{j = 1}^\infty ({c_j},{d_j}), \) where c j and d j are dyadic, and \( \mu (U) = \sum\nolimits_{j = 1}^\infty {({d_j} - {c_j}).} \) By choosing a sufficiently large natural number N we can make sure that \( U\prime = \cup_{j = 1}^N({c_j},{d_j}) \subseteq \,U \) satisfies \( \mu (U\prime) > \mu (U) - \frac{\varepsilon }{2}. \) Now define \( U\prime\prime \) to be the set obtained from \( U\prime \) by adding the endpoints of each interval: \( U\prime\prime = \cup_{j = 1}^N({c_j},{d_j}) \). Since we have added just finitely many points the measure of \( U\prime\prime \) is the same as that of \( U\prime \), and therefore, \( \mu [(E\backslash U\prime\prime) \cup (U\prime\prime\backslash E)] < \varepsilon . \)

Now apply the map\( \sum\nolimits_{j = 1}^\infty {{a_j}2^{ - j} \to ({a_1},\,} {a_2},{a_3}, \ldots ) \) which takes real numbers in [0, 1] to their sequence of binary coefficients in \( {\{ 0,1\}^\omega } \). Dyadic rationals have two binary developments, one ending with an infinite sequence of zeroes, and the other ending with an infinite sequence of ones. Adopt the convention that in case of a dyadic rational d, the map takes d to its two binary sequences. Since the set of dyadic numbers has measure zero the map is measure preserving. The set E is then mapped to a subset of \( \{ 0,1\}^\omega \)which we shall also denote by E. The set \( U\prime\prime \) is mapped to a finite subset of \( \{ 0,1\}^\omega \) which we will denote by \( {F_\varepsilon } \in \,\mathcal{F} \). The reason is that every closed interval with dyadic endpoints is mapped to a finite set, for example,\( \left[\frac{1}{4},\frac{5}{8}\right] \to \{ (0,1,0),(0,1,1),(1,0,0)\} \times \{ 0,1\} \times \{ 0,1\} \times \ldots \subseteq \{ 0,1\}^\omega,. \) and \( U\prime\prime \) is a finite union of such intervals. This completes the proof.

Appendix 2: Proof of Theorem 2

This theorem and proof appeared first in [24] as part of a criticism of the frequency interpretation of probability.

Theorem 2

Let A ⊂ {0, 1}ω be any measurable set with \( \mu (A) > \frac{1}{2} \), then there are a, bA such that ab has a divergent sequence of averages.

Proof

Denote by a ⊕ b the XOR of the elements a and b, in other words (a ⊕ b) i  = a i  + b i (mod 2). We first show that \( \mu (A) > \frac{1}{2} \) implies that A ⊕ A = {a ⊕ b; a, bA} = {0, 1}ω. Indeed if cA ⊕ A, then (c ⊕ A) ∩A = φ, where c ⊕ A = {c ⊕ a; aA}. Otherwise, if d ∈ (c ⊕ A) ∩ A then dA and d = (c ⊕ a) for some aA. Hence c = (d ⊕ a) ∈ A ⊕ A, contradiction. Therefore, (c ⊕ A) ∩ A = φ, but this also leads to a contradiction since \( \mu ({\mathbf{c}}\, \oplus \,A) = \mu (A) > \frac{1}{2} \), hence A ⊕ A = {0, 1}ω.

We can assume without loss of generality that all elements of A have a convergent sequence of averages. This is the case because the set of elements of {0, 1}ω whose averages diverge has measure zero. Let c ∈ {0, 1}ω be some sequence with a divergent sequence of averages. Then by the above argument there are a, bA such that c = (a ⊕ b), that is c i  = a i  + b i (mod 2) = a i  + b i − 2a i b i and therefore

$$ \frac{1}{n}\sum\limits_{i = 1}^n {{c_i} = \;\frac{1}{n}\sum\limits_{i = 1}^n {{a_i}} } + \frac{1}{n}\sum\limits_{i = 1}^n {{b_i}} - \frac{2}{n}\sum\limits_{i = 1}^n {{a_i}{b_i}} . $$

The sequence on the left diverges, and the first two sequences on the right converge. Hence, \( {n^{ - 1}}\sum\nolimits_{i = 1}^n {{a_i}{b_i}} \) diverges. This completes the proof.

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Pitowsky, I. (2012). Typicality and the Role of the Lebesgue Measure in Statistical Mechanics. In: Ben-Menahem, Y., Hemmo, M. (eds) Probability in Physics. The Frontiers Collection. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-21329-8_3

Download citation

  • DOI: https://doi.org/10.1007/978-3-642-21329-8_3

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-642-21328-1

  • Online ISBN: 978-3-642-21329-8

  • eBook Packages: Physics and AstronomyPhysics and Astronomy (R0)

Publish with us

Policies and ethics