Skip to main content
Log in

Lanford’s Theorem and the Emergence of Irreversibility

  • Original Paper
  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

It has been a longstanding problem to show how the irreversible behaviour of macroscopic systems can be reconciled with the time-reversal invariance of these same systems when considered from a microscopic point of view. A result by Lanford (Dynamical systems, theory and applications, 1975, Asterisque 40:117–137, 1976, Physica 106A:70–76, 1981) shows that, under certain conditions, the famous Boltzmann equation, describing the irreversible behaviour of a dilute gas, can be obtained from the time-reversal invariant Hamiltonian equations of motion for the hard spheres model. Here, we examine how and in what sense Lanford’s theorem succeeds in deriving this remarkable result. Many authors have expressed different views on the question which of the ingredients in Lanford’s theorem is responsible for the emergence of irreversibility. We claim that these interpretations miss the target. In fact, we argue that there is no time-asymmetric ingredient at all.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. The similarity is not complete, of course. For example, the BBGKY hierarchy is finite while the Boltzmann hierarchy is not.

  2. See ([23], p. 45; [9], p. 85) for a more extensive discussion of this property.

  3. It is true that Illner and Pulvirenti [14], have derived a longer validity but only under much more stringent conditions, i.e. for a gas cloud expanding into a vacuum. As a matter of fact, this repeated attention to time scale has deluded attention from more serious problems. Indeed, Lanford already pointed out that there is a simple, if merely technical, “fix” to the above problem: one would only need to require that assumption (i) of the theorem holds for arbitrary times, and not just at \(t = 0\), and Lanford’s result may be extended to all times (see [27] for a discussion of this issue).

  4. This can actually be seen as a counterexample to [20], who claimed that the irreversible approach to equilibrium would follow from taking the limit for \(N \longrightarrow \infty \).

  5. This argument is really an adaptation to the present context of a discussion of the notion of pre-collisional chaos contained in [28], p.35.

  6. In fact, in this case, that is when \(s = 0\), the condition coincides with its time-reversal.

  7. The proof of this result proceeds in the same way as that for the theorem for positive times, except that here one explicitly appeals to the outgoing configuration instead of the incoming one. However, in light of Proposition 3 establishing the equivalence of the BBGKY hierarchy written in terms of each configuration, the latter step is not an independent ingredient.

References

  1. Alexander, R.K.: The infinite hard-sphere system, Ph.D. Thesis, University of California at Berkeley (1975)

  2. Boltzmann, L.: Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen, Wien. Ber. 66, 275–370 (1872); in Boltzmann (1909) vol.I, pp. 316–402

  3. Boltzmann, L.: Über das Wärmegleichgewicht Von Gasen, auf welche äußere Kräfte wirken, Wien. Ber. 72, 427–457 (1872); in Boltzmann (1909) vol.II, pp. 1–30

  4. Boltzmann, L.: Bemerkungen über einige probleme der mechanische Wärmetheorie, Wien. Ber. 66, 275–370 (1877); in Boltzmann (1909) vol.II, pp. 112–148

  5. Boltzmann, L.: On certain questions of the theory of gases. Nature, 51, 413–415 (1895); in Boltzmann (1909) vol.III, pp. 535–544

  6. Boltzmann L.: Zu Hrn Zermelos Abhandlung Über die mechanische Erklärung irreversibler Vorgänge, Wied. Ann. 60, 392–398 (1897); in Boltzmann (1909), Vol III, paper 119

  7. Boltzmann, L.: Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen. In: Hasenöhrl, F. (ed.) Wissenschaftliche Abhandlungen. Chelsea, Leipzig (1909)

    Google Scholar 

  8. Cercignani, C.: 134 years of Boltzmann equation. In: Gallavotti, G., Gallavotti, W.L., Yngvason, J. (eds.) Boltzmann’s Legacy, pp. 107–128. Zürich: European Mathematical Society, Zürich (2008)

    Chapter  Google Scholar 

  9. Cercignani, C., Illner, R., Pulvirenti, M.: The Mathematical Theory of Diluted Gases. Springer, New York (1994)

    Book  Google Scholar 

  10. Culverwell, E.P.: Dr. Watson’s proof of Boltzmann’s theorem on permanence of distributions. Nature 50, 617 (1894)

    Article  ADS  MATH  Google Scholar 

  11. Drory, A.: Is there a reversibility paradox? Recentering the debate on the thermodynamic time arrow. Stud. Hist. Philos. Mod. Phys. 39, 889–913 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  12. Ehrenfest, P., Ehrenfest-Afanassjewa, T.: The Conceptual Foundations of the Statistical Approach in Mechanics. Cornell University Press, New York (1912)

    Google Scholar 

  13. Feynman, R., Leighton, R., Sands, M.: The Feynman Lectures on Physics. Addison Wesley, Reading (1964)

    Google Scholar 

  14. Illner, R., Pulvirenti, M.: Global validity of the Boltzmann equation for a two-dimensional rare gas in a vacuum. Commun. Math. Phys. 105, 189–203 (1986); erratum ibidem 121 143–146 (1989)

  15. Lanford, O.E.: Time evolution of large classical systems. In: Moser, J. (ed.) Dynamical Systems, Theory and Applications. Lecture Notes in Theoretical Physics, vol. 38, pp. 1–111. Springer, Berlin (1975)

    Chapter  Google Scholar 

  16. Lanford, O.E.: On the derivation of the Boltzmann equation. Asterisque 40, 117–137 (1976)

    MathSciNet  Google Scholar 

  17. Lanford, O.E.: The hard sphere gas in the Boltzmann-Grad limit. Physica 106A, 70–76 (1981)

    Article  ADS  Google Scholar 

  18. Lebowitz, J.L.: Microscopic dynamics and macroscopic laws. In: Horton Jr, C.W., Reichl, L.E., Szebehely, V.G. (eds.) Long-Time Prediction in Dynamics, pp. 220–233. Wiley, New York (1983)

    Google Scholar 

  19. Loschmidt, J.: Über die Zustand des Wärmegleichgewichtes eines Systems von Körpern mit Rücksicht auf die Schwerkraft. Wien. Ber. 73, 139 (1877)

  20. Goldstein, S.: Boltzmann’s approach to statistical mechanics. In: Bricmon, J., et al. (eds.) Chance in Physics: Foundations and Perspectives. Lecture Notes in Physics, pp. 39–54. Springer, Berlin (2001)

    Chapter  Google Scholar 

  21. Schrödinger, E.: Irreversibility. Proc. R. Ir. Acad. 53, 189–195 (1950)

    Google Scholar 

  22. Spohn, H.: Kinetic equations from Hamiltonian dynamics: markovian limits. Rev. Mod. Phys. 53, 569–615 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  23. Spohn, H.: Large Scale Dynamics of Interacting Particles. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  24. Spohn, H.: Loschmidt’s reversibility argument and the H-theorem. In: Fleischhacker, W., Schlonfeld, T. (eds.) Pioneering Ideas for the Physical and Chemical Sciences, Loschmidt’s Contributions and Modern Developments in Structural Organic Chemistry, Atomistics and Statistical Mechanics, pp. 153–158. Plenum Press, New York (1997)

    Google Scholar 

  25. Spohn, H.: On the integrated form of the BBGKY hierarchy for hard spheres. http://arxiv.org/pdf/math-ph/0605068 (2006). Accessed 25 May 2006

  26. Uffink, J.: Compendium of the foundations of classical statistical physics. In: Butterfield, J., Earman, J. (eds.) Handbook for the Philosophy of Physics, pp. 923–1074. Elsevier, Amsterdam (2008)

    Google Scholar 

  27. Valente, G.: The approach towards equilibrium in Lanford’s theorem. Eur. J. Philos. Sci. 4(3), 309–335 (2014)

    Article  MathSciNet  Google Scholar 

  28. Villani, C.: (Ir)reversibilité et entropie. Séminaire Poincaré XV, Le Temps, pp. 17–75. Institut Henri Poincaré, Paris (2010)

    Google Scholar 

Download references

Acknowledgments

The authors would like to thank Herbert Spohn for extensive comments on the technical and conceptual subtleties of Lanford’s theorem, as well as two anonymous referees for helpful remarks.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giovanni Valente.

Appendix

Appendix

In this appendix we consider in more detail the issue of time-reversal invariance for the Boltzmann equation and two versions of the BBGKY hierarchy equations and prove the claims concerning this issue in the body of the paper. The commonly accepted criterium for judging time-reversal invariance of such evolution equations, describing a density function (either \(f\) or \(\rho _k\)) over particle configurations is as follows: the equation determines a class \(S\) of allowed solutions, where each solution can be seen as a ‘history’ of the density function, either \( \mathcal{H}:= \{ f_t, \; t\in {\mathbb {R}}\} \in S \) or \(\mathcal{H} := \{ \rho _{k,t}, \; t\in {\mathbb {R}}\}\), carving out, so to say, a trajectory in their respective spaces of all density functions that solve the evolution equation.

Now consider the time-reversal of such a history, defined as \(\mathcal{TH} := \{\tilde{f}_{t} \; t\in {\mathbb {R}}\}\) or \(\mathcal{TH}:= \{\tilde{\rho }_{k,t} \; t\in {\mathbb {R}}\}\) , where \(\tilde{f}_t (\vec {q}, \vec {p})= f_{-t} (\vec {q},- \vec {p})\) and \(\tilde{\rho }_{k,t} (\vec {q}_1, \vec {p}_1; \ldots ; \vec {q}_k , \vec {p}_k) = \rho _{k,-t} (\vec {q}_1, -\vec {p}_1; \ldots ; \vec {q}_k ,- \vec {p}_k)\) are obtained from \(f_t\) and \(\rho _{k,t}\) by reversing \(t\) and the momenta in their arguments. The question is then whether such a time-reversed history \(\mathcal{TH}\) is a solution of the equation too, whenever \(H\) is a solution of the equation in question. In other words: Is \(\mathcal{T} H \in S\) whenever \(T \in S\)? If the answer is yes, the equation is time-reversal invariant, otherwise not.

Our strategy will be the same in all three cases. We consider an arbitrary solution of the equation and construct from this a time-reversed history, and derive the equation it obeys. If the resulting equation is equivalent to the original equation we have proved that the original equation is time-reversal invariant. But if it is not, the original equation is not time-reversal invariant. To be sure, our first two propositions below are not surprising: every author in the field recognizes that the Boltzmann equation is not time-reversal invariant, whereas the BBGKY hierarchy is, even if detailed proofs of these assertions are hard to find. We have added these proofs mainly to show that that by the same standard of rigour, one can obtain Proposition 3, which we believe to be of more interest.

Proposition 1

The Boltzmann equation (8) is not time-reversal invariant.

Proof

Since nothing interesting happens to the position variables in the Boltzmann equation (8) we will suppress them in the notation below, and also put the mass \(m=1\). Note that \(\vec {p}_1\) is the only independent momentum variable in the equation: \(\vec {p}_2\) appears only in the right-hand side as a mere integration variable and the outgoing momenta variables \(\vec {p}_1^{\ \prime }, \vec {p}_2^{\ \prime }\) in the collision integral are functions of \(\vec {p}_i\): that is, \(\vec {p}_i^{\ \prime } =\vec {p}_i^{\ \prime }(\vec {p}_1, \vec {p}_2) = T_{\omega _{12}} (\vec {p}_1, \vec {p}_2)\), where \(T_{\omega _{12}}\) is defined by (7). So, let \( f_t (\vec {q}, \vec {p}) , t\in {\mathbb {R}}\), by an arbitrary solution of the Boltzmann equation (8), and consider the joint transformation of \(t\longrightarrow -t\) and \(\vec {p}_1 \longrightarrow - \vec {p}_1\). Applying this transformation to the left-hand side of (8) leads to:

$$\begin{aligned} - \displaystyle \frac{\partial }{\partial t} {f}_{-t}(- \vec {p}_{1}) \displaystyle -{\vec {p}_{1}} \cdot \frac{\partial }{\partial \vec {q}} \,{f}_{-t}(- \vec {p}_{1}) = \displaystyle \frac{\partial }{\partial t} \tilde{f}_{t}( \vec {p}_{1}) \displaystyle +{\vec {p}_{1}} \cdot \frac{\partial }{\partial \vec {q}} \,\tilde{f}_{t}( \vec {p}_{1}), \end{aligned}$$
(56)

while the right-hand side of (8) becomes

$$\begin{aligned}&Na^{2} \displaystyle \int \! d \vec {p}_2 \int _{\vec {\omega }_{12} \cdot (\vec {p}_{1} +\vec {p}_{2}) \leqslant 0} \! d \vec {\omega }_{12} \,(-\vec {p}_{1} - \vec {p}_{2}) \cdot \vec {\omega }_{12} \,\nonumber \\&\quad \left[ {f}_{-t}( \vec {p}_{1}^{\prime }) {f}_{-t}( \vec {p}_{2}^{\prime } )-{f}_{-t}(-\vec {p}_{1})f_{-t}(\vec {p}_2) \right] \end{aligned}$$
(57)

Here, the notation \(\vec {p}_i^{\ \prime \prime }\) is used to indicate that these momenta have to be thought of as functions of \((-\vec {p}_1, \vec {p}_2)\): \( (\vec {p}_1^{\ \prime \prime }, \vec {p}_2^{\ \prime \prime }) = T_{\omega _{12}} (-\vec {p}_1, \vec {p}_2)\).

If we now perform an additional (cosmetic) transformation of the integration variables \(\vec {p}_2 \longrightarrow - \vec {p}_2\) and \(\vec {\omega }_{12} \longrightarrow - \vec {\omega }_{12}\), (57) can also be written as

$$\begin{aligned}&Na^{2} \displaystyle \int \! d \vec {p}_2 \int _{\vec {\omega }_{12} \cdot (\vec {p}_{1} - \vec {p}_{2}) \geqslant 0} \! d \vec {\omega }_{12} \,(\vec {p}_{1} - \vec {p}_{2}) \cdot \vec {\omega }_{12} \,\nonumber \\&\quad \left[ {f}_{t}( \vec {p}_{1}^{\ \prime \prime \prime }) {f}_{t}( \vec {p}_{2}^{\ \prime \prime \prime }) -{f}_{t}(-\vec {p}_{1}) {f}_{t}(- \vec {p}_{2}) \right] , \end{aligned}$$
(58)

where \((\vec {p}_1^{\ \prime \prime \prime }, \vec {p}_2^{\ \prime \prime \prime }) := T_{\omega _{12}} (-\vec {p}_1, -\vec {p}_2)\). But \(T_{\omega _{12}}\) is a linear operator, and therefore \(\vec {p}_i^{\ \prime \prime \prime } = -\vec {p}_i^{\ \prime }\). If we now substitute \(f_{-t}(-\vec {p}) = \tilde{f}_t( \vec {p})\) in (58) and equate the transformed left-hand side (55) of (8) to the transformed right-hand side (58) of (8), we find that \(\tilde{f}_t\) satisfies the equation

$$\begin{aligned}&\displaystyle \frac{\partial }{\partial t} \tilde{f}_{t}( \vec {p}_{1}) \displaystyle +{\vec {p}_{1}} \cdot \frac{\partial }{\partial \vec {q}} \,\tilde{f}_{t}( \vec {p}_{1}) = \nonumber \\&-Na^{2} \displaystyle \int \! d \vec {p}_2 \int _{\vec {\omega }_{12} \cdot (\vec {p}_{1} - \vec {p}_{2}) \geqslant 0} \! d \vec {\omega }_{12} \,(\vec {p}_{1} - \vec {p}_{2}) \cdot \vec {\omega }_{12} \, \left[ \tilde{f}_{t}( \vec {p}_{1}^{\prime }) \tilde{f}_{t}( \vec {p}_{2}^{\prime }) \!-\!\tilde{f}_{t}(\vec {p}_{1}) \tilde{f}_{t}(\vec {p}_{2}) \right] ,\nonumber \\ \end{aligned}$$
(59)

also known as the anti-Boltzmann equation. We conclude: whenever the solution \(\{ f_t \;, t\in {\mathbb {R}}\}\) satisfies the Boltzmann equation, the time reversed solution \(\{ \bar{f}_{-t} \;, t\in {\mathbb {R}}\}\) solves the (inequivalent) anti-Boltzmann equation, and therefore, the Boltzmann equation is not time-reversal invariant. \(\square \)

Proposition 2

The BBGKY hierarchy with the collision term given by (29) is time-reversal invariant.

Proof

Recall that the BBGKY hierarchy has the form, for \(k=1, \ldots N\):

$$\begin{aligned} \frac{\partial \rho _{k,t}^{(a)}(x_1,\ldots ,x_k)}{\partial t} - \mathcal{H}_k \rho _{k,t}^{(a)}(x_1, \ldots , x_k) = \left( { \mathcal C}^{(a)} _{k, k+1} \rho _{k+1, t}^{(a)}\right) (x_1, \ldots , x_k). \end{aligned}$$
(60)

In this equation, we deal with \(k\) particles (the momentum of the \(k+1\)th particle appears in (29) only as an integration variable). If we reverse sign of the momenta \(\vec {p}_1, \ldots \vec {p}_k\) and the sign of \( t\), it is easy to see the the left hand side of (66) changes sign. But here, the right-hand side (29) clearly changes sign too when we change sign of all momenta \(\vec {p}_1, \ldots , \vec {p}_{k+1}\), due to the fact that the integration over the antisymmetric factor \(\left( \vec {\omega }_{i, k+1}\cdot \big (\vec {p}_{k+1} - \vec {p}_i\big ) \right) \) in the integrand is extended over the entire surface of theunit sphere. More explicitly, if we use the notation \(\bar{x}_i = (\vec {q}_i,- \vec {p}_i)\) along \(x_i = (\vec {q}_i, \vec {p}_i)\), the transformed version of the left-hand side of Eq. (60) is

$$\begin{aligned}&-\frac{\partial }{\partial t} \rho _{k,-t}^{(a)}(\bar{x}_1,\ldots ,\bar{x}_k)+ \mathcal{H}_k \rho _{k,-t}^{(a)}(\bar{x}_1, \ldots , \bar{x}_k)= -\frac{\partial }{\partial t}\tilde{\rho }_{k,t}^{(a)}({x}_1,\ldots ,{x}_k)\nonumber \\&\quad +\, \mathcal{H}_k \tilde{\rho }_{k,t}^{(a)}({x}_1, \ldots , {x}_k), \end{aligned}$$
(61)

while the right-hand side transforms into:

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho _{k+1,-t}^{(a)}\right) (\bar{x}_1, \ldots \bar{x}_k) = Na^2 \sum _{i=1}^k \int _{{\mathbb {R}}^3} d\vec {p}_{k+1} \int _{S^2} d\vec {\omega }_{i,k+1}\nonumber \\&\quad \left( \vec {\omega }_{i, k+1}\cdot \big (\vec {p}_{k+1} + \vec {p}_i\big ) \right) \rho _{k+1,-t}^{(a)}(\bar{x}_1, \ldots , \bar{x}_k , \vec {q}_{i} + a \vec {\omega }_{i,k+1},\vec {p}_{k+1}). \end{aligned}$$
(62)

Hence, if we rewrite the integration variable \(\vec {p}_{k+1}\) as \(-\vec {p}_{k+1}\), we obtain from (62):

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho _{k+1,-t}^{(a)}\right) (\bar{x}_1, \ldots , \bar{x}_k) = \nonumber \\&\quad - Na^2 \sum _{i=1}^k \int _{{\mathbb {R}}^3}\!\!\!d\vec {p}_{k+1} \int _{S^2}\!\!\! d\vec {\omega }_{i,k+1} \left( \vec {\omega }_{i, k+1}\cdot \big ( \vec {p}_{k+1} - \vec {p}_i\big ) \right) {\rho }_{k+1,-t}^{(a)}(\bar{x}_1, \ldots , \bar{x}_k , \vec {q}_{i}\nonumber \\&\quad +\, a \vec {\omega }_{i,k+1} ,-\vec {p}_{k+1}) \, = \nonumber \\&\quad - Na^2 \sum _{i=1}^k \int _{{\mathbb {R}}^3}\!\!\!d\vec {p}_{k+1} \int _{S^2}\!\!\! d\vec {\omega }_{i,k+1} \left( \vec {\omega }_{i, k+1}\cdot \big ( \vec {p}_{k+1} - \vec {p}_i\big ) \right) \tilde{\rho }_{k+1,-t}^{(a)}({x}_1, \ldots , {x}_k , \vec {q}_{i}\nonumber \\&\quad +\, a \vec {\omega }_{i,k+1} ,\vec {p}_{k+1}). \end{aligned}$$
(63)

Comparing this with (29), we conclude

$$\begin{aligned} \left( \mathcal{C}^{(a)}_{k, k+1} \rho _{k+1,-t}^{(a)}\right) (\bar{x}_1, \ldots \bar{x}_k) = - ({ \mathcal C}^{(a)} _{k, k+1}\tilde{ \rho }_{k+1, t}^{(a)})({x}_1, \ldots , {x}_k). \end{aligned}$$
(64)

Putting (61) and (64) together, we see that the time-reversed version \(\tilde{\rho }^{(a)}_{k,t}\) of an arbitrary solution \(\rho ^{(a)}_{k,t}\) of (60) obeys the equivalent equation

$$\begin{aligned} -\frac{\partial }{\partial t}\tilde{\rho }_{k,t}^{(a)} + \mathcal{H}_k \tilde{\rho }_{k,t}^{(a)} = - { \mathcal C}^{(a)} _{k, k+1}\tilde{ \rho }_{k+1, t}^{(a)}. \end{aligned}$$
(65)

This shows that if \(\{ \rho _{k+1, t}^{(a)}\, , t \in {\mathbb {R}}\}\) solves Eq. (60), then \(\{ \tilde{ \rho }_{k+1, t}^{(a)}\, , t \in {\mathbb {R}}\}\) is a solution of the same equation, so that we can conclude that (60) is time-reversal invariant. \(\square \)

Proposition 3

Let the continuity across collisions condition (53) hold. Then, the BBGKY hierarchy with the collision term expressed by (32) is equal to the BBGKY hierarchy with the collision term expressed by (52). Furthermore, it is time-reversal invariant.

Proof

Recall that, after adopting the incoming configuration for collision points, the BBGKY hierarchy takes the form

$$\begin{aligned} \frac{\partial \rho _{k,t}^{(a)}(x_1,\ldots ,x_k)}{\partial t} - \mathcal{H}_k \rho _{k,t}^{(a)}(x_1, \ldots , x_k) = \left( { \mathcal C}^{(a)} _{k, k+1} \rho _{k+1, t}^{(a)}\right) (x_1, \ldots , x_k), \end{aligned}$$
(66)

where the left-hand side is the same as in (60), but the collision operator is now expressed by (32). On the other hand, when adopting the outgoing configuration for collision points, the collision operator is expressed by (52).

We first show that the BBGKY hierarchy with the collision term expressed by (32) is equal to the BBGKY hierarchy with the collision term expressed by (52). Since the only apparent difference between these hierarchies lies in the collision term, one just needs to show that the operator (32) is the same as (52). Let us focus on the former. Notice that the phase point \((x_{1}, \ldots , x_{i-1},\vec {q}_{i} ,\vec {p}_{i}^{\ \prime }, x_{i+1}, \ldots x_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1} , \vec {p}_{k+1}^{\ \prime })\) as well as the phase point \((x_{1}, \ldots , x_{i-1},\vec {q}_{i} ,\vec {p}_{i}, x_{i+1}, \ldots x_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1} , \vec {p}_{k+1})\) are pre-collision coordinates in the hemisphere \(\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0 \). Therefore, if the continuity at collisions condition holds, it follows that (32) can be written as

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,t} \right) (x_1, \ldots , x_k) \nonumber \\&\quad = Na^{2} \sum _{i=1}^{k}\int _{{\mathbb {R}}^3} \!d\vec {p}_{k+1} \int _{\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0} \! \!\!d \vec {\omega }_{i,k+1}\, \, \, \vec {\omega }_{i,k+1}\cdot \left( \vec {p}_{i}- \vec {p}_{k+1} \right) \nonumber \\&\quad \times \big [ \rho _{k+1,t}^{(a)}(x_{1}, \ldots , x_{i-1},\vec {q}_{i} ,\vec {p}_{i}, x_{i+1}, \ldots x_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1} , \vec {p}_{k+1})\nonumber \\&\quad - \rho _{k+1,t}^{(a)}(x_{1}, \ldots , x_{i-1}, \vec {q}_{i} , \vec {p}_{i}^{\ \prime }, x_{i+1}, \ldots x_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1}, \vec {p}_{k+1}^{\ \prime }) \big ], \end{aligned}$$
(67)

which is just equal to the collision term (52).

Next, we demonstrate that the BBGKY hierarchy with the collision term expressed by (32) is time-reversal invariant. Here, we are again dealing with \(k\) particles, but in this case both incoming and outgoing momenta appear in the same formula, just as in the Boltzmann equation. And just as in the Boltzmann equation, one ought to take the outgoing momenta variables here as (implicit) functions of the incoming momenta: \((\vec {p}_i^{\ \prime }, \vec {p}_{k+1}^{\ \prime }) = T_{\omega _{i,k+1}} (\vec {p}_i, \vec {p}_{k+1})\). We now apply a combination of the arguments we used above to judge the time-reversal invariance of the Boltzmann equation and the BBGKY hierarchy in the version (60): we replace \( t \) by \(- t \) and \( (x_1 , \ldots , x_k)\) by \((\bar{x}_1 , \ldots , \bar{x}_k)\). Since the left-hand side is the same as in (60), we draw the same conclusion: this side transforms into (61). But we have to scrutinize the behaviour of the right-hand side in more detail. This side transforms to:

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k) \nonumber \\&\quad = d\vec {p}_{k+1} d \vec {\omega }_{i,k+1}\, \, \vec {\omega }_{i,k+1}\cdot \left( \vec {p}_{i}+\vec {p}_{k+1} \right) \nonumber \\&\quad \times \big [ \rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,\vec {p}_{i}^{\ \prime }, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1}, \vec {p}_{k+1}^{\ \prime })\nonumber \\&\quad -\rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1}, \vec {q}_{i} , -\vec {p}_{i}, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1}, -\vec {p}_{k+1}) \big ], \end{aligned}$$
(68)

where, as before, the variables \((\vec {p}_i^{\ \prime }, \vec {p}_{k+1}^{\ \prime })\) are defined as \( (\vec {p}_i^{\ \prime }, \vec {p}_{k+1}^{\ \prime }) = T_{\omega _{i,k+1}} (-\vec {p}_i, \vec {p}_{k+1})\). Repeating a similar step of our first argument, we perform a conventional transformation on the integration variables \(\vec {p}_{k+1} \longrightarrow -\vec {p}_{k+1}\) and \( \vec {\omega }_{_i,k+1} \longrightarrow - \vec {\omega }_{i, k+1}\) and use that the primed momenta transform as \((\vec {p}_i^{\ \prime },\vec {p}_{k+1}^{\ \prime }) \longrightarrow T_{\omega _{i,k+1}} (-\vec {p}_i, -\vec {p}_{k+1}) = (-\vec {p}_i^{\ \prime },-\vec {p}_{k+1}^{\ \prime })\) to rewrite the integral (68) as

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k)\nonumber \\&\quad = Na^{2} \sum _{i=1}^{k}\int _{{\mathbb {R}}^3} \!d\vec {p}_{k+1} \int _{\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0} \! \!\!d \vec {\omega }_{i,k+1}\, \vec {\omega }_{i,k+1}\cdot \left( \vec {p}_{i}- \vec {p}_{k+1} \right) \nonumber \\&\quad \times \big [ \rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,-\vec {p}_{i}^{\ \prime }, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1} , -\vec {p}_{k+1}^{\ \prime }) \nonumber \\&\quad -\rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1}, \vec {q}_{i} , -\vec {p}_{i}, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1}, -\vec {p}_{k+1}) \big ]. \end{aligned}$$
(69)

Now, in analogy to the previous case (cf. Eqn. (64)), we must show that

$$\begin{aligned} \left( \mathcal{C}^{(a)}_{k, k+1} {\rho }_{k+1,-t}^{(a)}\right) (\bar{x}_{1}, \ldots , \bar{x}_{k}) \left. \displaystyle { =} \right. - \left( \mathcal{C}^{(a)}_{k, k+1} \tilde{\rho }^{(a)}_{k+1,t}\right) ({x}_{1}, \ldots , {x}_{k}). \end{aligned}$$
(70)

For this purpose, notice that the phase point \((\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,-\vec {p}_{i}^{\ \prime }, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1} , -\vec {p}_{k+1}^{\ \prime })\) as well as the phase point \((\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,-\vec {p}_{i}, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1} , -\vec {p}_{k+1})\) are pre-collision coordinates in the hemisphere \(\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0 \), and thus one can apply the continuity at collisions condition (53) to both such points. The collision term in (67) can thus be written as

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k) \nonumber \\&\quad = Na^{2} \sum _{i=1}^{k}\int _{{\mathbb {R}}^3} \!d\vec {p}_{k+1} \int _{\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0} \! \!\!d \vec {\omega }_{i,k+1}\, \, \, \vec {\omega }_{i,k+1}\cdot \left( \vec {p}_{i}- \vec {p}_{k+1} \right) \nonumber \\&\quad \times \big [ \rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,-\vec {p}_{i}, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1} , -\vec {p}_{k+1})\nonumber \\&\quad -\rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1}, \vec {q}_{i} , -\vec {p}_{i}^{\ \prime }, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1}, -\vec {p}_{k+1}^{\ \prime }) \big ]. \end{aligned}$$
(71)

If one then performs the (cosmetic) transformation of the integration variable \(\vec {\omega }_{i, k+1}\) into \( - \vec {\omega }_{i, k+1}\), one obtains

$$\begin{aligned}&\left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k)\nonumber \\&\quad = Na^{2} \sum _{i=1}^{k}\int _{{\mathbb {R}}^3} \!d\vec {p}_{k+1} \int _{\vec {\omega }_{i,k+1} \cdot (\vec {p}_{i} -\vec {p}_{k+1}) \geqslant 0} \! \!\!d \vec {\omega }_{i,k+1}\, \vec {\omega }_{i,k+1}\cdot \left( \vec {p}_{i}- \vec {p}_{k+1} \right) \nonumber \\&\quad \times \big [ \rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1},\vec {q}_{i} ,-\vec {p}_{i}, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} - a \vec {\omega }_{i, k+1} , -\vec {p}_{k+1}) \nonumber \\&\quad -\rho _{k+1,-t}^{(a)}(\bar{x}_{1}, \ldots , \bar{x}_{i-1}, \vec {q}_{i} , -\vec {p}_{i}^{\ \prime }, \bar{x}_{i+1}, \ldots \bar{x}_k, \vec {q}_{i} + a \vec {\omega }_{i, k+1}, -\vec {p}_{k+1}^{\ \prime }) \big ], \end{aligned}$$
(72)

which is equal to \(- \left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k) \), as one can see by contrasting the above equation with (67). By recalling the definition of \(\tilde{\rho }_{k, t}^{(a)}\), we conclude that

$$\begin{aligned} \left( \mathcal{C}^{(a)}_{k, k+1} \rho ^{(a)}_{k+1,-t} \right) (\bar{x}_1, \ldots , \bar{x}_k)&= - \left( \mathcal{C}^{(a)}_{k, k+1} \tilde{\rho }^{(a)}_{k+1,t} \right) (x_1, \ldots , x_k). \end{aligned}$$
(73)

It has thus been shown that the BBGKY hierarchy with the collision term expressed by (32) is time-reversal invariant. Equivalently, the BBGKY hierarchy with the collision term expressed by (52) is time-reversal invariant too. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Uffink, J., Valente, G. Lanford’s Theorem and the Emergence of Irreversibility. Found Phys 45, 404–438 (2015). https://doi.org/10.1007/s10701-015-9871-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-015-9871-z

Keywords

Navigation