Skip to main content
  • 1229 Accesses

Abstract

In this work, we are interested in a molecular theory (i.e., statistical mechanics) of time- and space-dependent nonequilibrium (irreversible) processes in matter regarded as composed of many discrete particles.

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 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
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.

    This viewpoint toward volume (per mole) should be modified because it does not have a molecular representation free from macroscopic variables in its definition. It will be found more appropriate to regard it as a mean volume allowed to molecules in a randomly distributed assembly of particles. Especially, if the fluid is away from equilibrium the latter concept is more appropriate since specific volume is not the same as the mean volume per molecule. For the notion of molecular representation of mean volume per molecule, see Ref. [7] and also Ref. [8]. Modification of manifold \(\mathfrak {P}\) will be discussed in a later chapter in which volume transport phenomena are discussed.

  2. 2.

    A conservation law can be best stated by means of a cyclic process since the state of the system must be exactly restored on completion of the cyclic process and so must be the value of internal energy of the system. Vanishing cyclic integrals consequently play important roles in formulation of thermodynamics of irreversible processes.

  3. 3.

    In some textbooks in thermodynamics inexact differentials are denoted with a bar on the differential sign, e.g., or . In this work we do not adopt such a notation.

  4. 4.

    This modification of the viewpoint can be implemented if the notion of volume allotted to a molecule is adopted [7] via the Voronoi volume. This notion will be more easily understood if the theory is formulated as a molecular theory as will be done in a later chapter on irreversible processes in condensed phase where volume transport is explicitly taken into account. In phenomenological formulation, the notion must be formally accepted, but unfortunately would remain abstract.

  5. 5.

    It should be noted that fluid particle velocity is not the velocity of a molecule comprising the fluid. It is the velocity of the packet of fluid in an elementary volume of fluid at position \(\mathbf {r}\) and time t. From the molecular theory viewpoint it is the ensemble average of molecular velocities in a sufficiently small volume around the position \(\mathbf {r}\) at time t. For its molecular theory definition, see Chaps. 3, 5, 6, and 7 of this work.

  6. 6.

    This theorem is proved in Chap. 4, Ref. [3]. See also H. Poincaré, Thermodynamique (Georges Carré, Paris, 1892).

  7. 7.

    It must be recognized that the presence and notion of unavailable work in an irreversible cycle is not as apparent in the Clausius and Kelvin principles as in the Carnot theorem, although they are equivalent to the Carnot theorem. In the end, it was the Carnot theorem that Clausius used to formulate the inequality named after him.

  8. 8.

    In this connection, I would like to quote a passage in an article by I. Prigogine in Science 201, 777 (1978) where he states “150 years after its formulation the second law of thermodynamics still appears to be more a program than a well-defined theory in the usual sense, as nothing precise (except the sign) is said about the S production”. In view of the fact that entropy was not defined for irreversible processes by Clausius, we interpret that the S production here is meant for the uncompensated heat despite the notation S used for entropy in the aforementioned paper. In fact, Prigogine’s use of terminology appears to be oblivious to the genesis of the notion of Clausius entropy S, as is evident from the discussion given below.

  9. 9.

    Reversible processes are traditionally defined as those of quasistatic processes which are in continuous equilibrium with the surroundings, but it is more mathematically precise to define them as quasistatic processes for which \(\mathrm {N}=0.\) This will be the definition of reversible processes used throughout in this work.

  10. 10.

    This interpretation of differentials is usually made in linear theory of irreversible thermodynamics. See [17, 30, 31].

  11. 11.

    There are basically two different versions of extended irreversible thermodynamics: one class of versions can be found in Refs. [36, 37, 41, 43, 44] and the other in Refs. [6, 35, 45, 46] and this work. In the former, it is assumed that there exists a nonequilibrium entropy which is a state function in the thermodynamic space and the nonequilibrium entropy is statistically represented by approximations of the Boltzmann entropy or its dense fluid generalization or the information entropy for dynamical systems, without a support of a kinetic equation. In the latter class (i.e., in Refs. [6, 45, 46]), it is shown from the second law of thermodynamics that there exists a quantity called compensation function (renamed to calortropy in this and earlier work) and its differential is an exact differential in the thermodynamics space by virtue of the second law. The compensation function, however, is not the same as the Boltzmann entropy appearing in the kinetic theory of dilute gases by Boltzmann. Since thermodynamics of irreversible processes must be securely founded on the thermodynamic laws, it is crucial to show that the basic equations are consistent with the thermodynamic laws and, for example, the extended Gibbs relation is equivalent to the second law of thermodynamics. An assumption for such a basic equation is not acceptable if the resulting theory will have anything to do with the thermodynamic laws in accounting for macroscopic processes in nature. Neither can the thermodynamic laws afford approximate representations. The basic thermodynamic equation in the formulation made in Refs. [6, 46] and in this work is a rigorous consequence of the second law of thermodynamics which mathematically extends equilibrium thermodynamics that is certainly endowed with a physical basis supported by the second law; it is not an assumption as in Refs. [3544].

  12. 12.

    A simple, albeit somewhat ideal, example would be a cyclic process consisting of the irreversible segment where \(dN<0\) and a reversible segment for the remainder of the cycle in which \(dN=0\). This cycle, if possible to construct, would violate the second law.

  13. 13.

    In his work, Clausius made no mention of embedding the states A and B in the equilibrium part of the thermodynamic manifold. He simply constructed a cycle consisting of an irreversible segment starting from state A and ending at state B and a reversible segment starting from state B and restoring the system to state A. Since states A and B are also part of a reversible process, they must be embedded in the equilibrium part of the thermodynamic manifold. It now may be stated that there was no clear notion of thermodynamic manifold in the Clausius formulation of the second law of thermodynamics.

  14. 14.

    Inequality (2.60), together with (2.54) and (2.55), means that for an isolated system \(\Delta S_{\text {e}}\ge 0\) as is generally taken in the thermodynamics literature. However, this inequality for an isolated system can be misleading and cause confusion. In applying this inequality, it must be remembered that \(\Delta S_{\text {e}}\) is for the reversible segment complementary to the irreversible step making up the cycle in question and the inequality (2.60) simply means that \(\Delta S_{\text {e}}\) for the reversible segment is always larger than the irreversible compensated heat change. If the system is isolated, the compensated heat \(dQ=0\) everywhere in the interval of integration and hence \(\Delta S_{\text {e}}=0\) identically. This conclusion is also consistent with the definition of \(dS_{\text {e}}\) in (2.44). It is clearly convenient to think in terms of \(\Delta \Psi \ge 0\) instead of \(\Delta S_{\text {e}}\ge 0\).

  15. 15.

    If the volume transport is included, the range of index s may be taken \(s\ge 0\), if we adopt the system of notation used in the kinetic theory chapters in this work. See Chaps. 6 and 7.

  16. 16.

    Although (2.73) and (2.74) are sufficient for \(\mathbf {J}_{\text {c}}\) and \(\Xi _{\text {c}}\) to yield a one-form for \(d_{t}\hat{\Psi }\), it is not unique. For example, a vector \(\mathbf {A}\) and its divergence \(\nabla \varvec{\cdot A}\) can be added to \(\mathbf {J}_{\text {c}}\) and \(\Xi _{\text {c}}\), respectively, with no effect at all on the one-form that can be obtained from the balance equation for the calortropy. In fact, the vector \(\mathbf {A}\) can be taken with \(\sum _{a=1}^{r}\mathbf {J}_{a}/m_{a}\) as can be seen later in Chap. 3. The formulas for \(\mathbf {J}_{\text {c}}\) and \(\Xi _{\text {c}}\) proposed in Proposition 4 therefore are minimal in the sense that they are sufficient to produce a one-form for \(d_{t}\hat{\Psi }.\)

  17. 17.

    It must be noted that a Pfaffian differential form is not necessarily an exact differential unless it satisfies a set of integrability conditions [5053]. In the case of thermodynamics, the second law of thermodynamics preempts the integrability conditions which are partial differential equations not easy to solve in general.

  18. 18.

    These phenomenological transport coefficients may be identified with their moecular theory (i.e., kinetic theory) formulas in the forms of collision bracket integrals, which appear in the kinetic theory chapters.

  19. 19.

    Originally, the Rayleigh dissipation function was quadratic with respect to velocities. Here we are generalizing not only the original definition for fluxes (velocities) to \(\Phi _{a}^{(s)}\), but also the flux dependence of the dissipation function to a highly nonlinear form from a quadratic function of velocities.

  20. 20.

    In actual applications of the present theory, the nonconserved variables require closure relations. For example, in the case of single-component fluids it will be found useful to take the first thirteen moments (i.e., density, pressure, three velocity components, three components of heat flux, two normal stresses, three components of shear stress) for macroscopic variables and then assume the closure relations \( {\varvec{\psi }}_{a}^{(1)}= {\varvec{\psi }}_{a}^{(3)}=0\). Applications discussed in a later chapter take this kind of closure relations.

References

  1. J.G. Kirkwood, I. Oppenheim, Chemical Thermodynamics (McGraw-Hill, New York, 1961)

    Google Scholar 

  2. H.B. Callen, Thermodynamics (Wiley, New York, 1960)

    MATH  Google Scholar 

  3. B.C. Eu, M. Al-Ghoul, Chemical Thermodynamics (World Scientific, Singapore, 2010)

    Book  Google Scholar 

  4. J.F. Schooley (ed.) in chief, Temperature, Parts 1 and 2, vol. 5 (American Institute of Physics, New York, 1982)

    Google Scholar 

  5. G.N. Lewis, M. Randall, Thermodynamics (McGraw-Hill, New York, 1961)

    Google Scholar 

  6. B.C. Eu, Generalized Thermodynamics (Kluwer, Dordrecht, 2002)

    Book  MATH  Google Scholar 

  7. B.C. Eu, J. Chem. Phys. 129, 094502, 134509 (2008)

    Google Scholar 

  8. J.S. Hunjan, B.C. Eu, J. Chem. Phys. 132, 134510 (2010)

    Article  ADS  Google Scholar 

  9. Y. Elkana, The Discovery of Energy (Hutchinson Educational, London, 1974)

    MATH  Google Scholar 

  10. R.B. Lindsay (ed.), Energy: Historical Development of the Concept (Dowden, Hutchinson, and Ross, Stroudburg, PA, 1975)

    Google Scholar 

  11. S.G. Brush, The Kind of Motion We Call Heat, vols. 1, 2 (North-Holland, Amsterdam, 1976)

    Google Scholar 

  12. R. Clausius, Ann. Phys. 125, 313 (1965). (Leipzig)

    Google Scholar 

  13. M. Planck, Thermodynamics (Dover, New York, 1945)

    Google Scholar 

  14. B.C. Eu, J. Phys. Chem. 103, 8583 (1999)

    Article  Google Scholar 

  15. B.C. Eu, J. Phys. Chem. 91, 1184 (1987)

    Article  Google Scholar 

  16. A.C. Eringen (ed.), Continuum Physics, vol. 2 (Academic, New York, 1975)

    Google Scholar 

  17. For example, see P. Mazur, S.R. de Groot, Nonequilibrium Thermodynamics (North-Holland, Amsterdam, 1962)

    Google Scholar 

  18. R. Haase, Thermodynamics of Irreversible Processes (Dover, New York, 1969)

    Google Scholar 

  19. H.A. Barnes, J.F. Hutton, K. Walters, An Introduction to Rheology (Elsevier, Amsterdam, 1989). For some examples of applications to Rheological Phenomena see Chapter 14 of Ref. [46]

    Google Scholar 

  20. D.K. Ferry, S.M. Goodnick, Transport in Nanostructure (Cambridge University Press, London, 1997). C.S. Ting (ed.), Physics of Hot Electron Transport in Semiconductors (World Scientific, Singapore, 1992)

    Google Scholar 

  21. For rarefied gas dynamic phenomena, see M. N. Kogan, Rarefied Gas Dynamics (Plenum, New York, 1967). Some examples for application of the present line of theory are also discussed in B.C. Eu, Nonequilibrium Statistical Mechanics (Kluwer, Dordrecht, 1998)

    Google Scholar 

  22. E.A. Mason, E.W. McDaniel, Transport Properties of Ions in Gases (Wiley, New York, 1988). J.M.G. Barthel, H. Krienke, W. Kunz, Physical Chemistry of Electrolyte Solutions (Springer, New York, 1998)

    Google Scholar 

  23. J.C. Maxwell, Philos. Trans. R. Soc. Lond. 157, 49 (1867)

    Article  Google Scholar 

  24. S. Chapman, Proc. R. Soc. Lond. A 93, 1 (1916–17)

    Google Scholar 

  25. J. Kestin (ed.), The Second Law of Thermodynamics (Dowden, Hutchinson & Ross, Stroudburg, PA, 1976)

    Google Scholar 

  26. S. Carnot, Refléxions sur la Puissance Motrice du Feu et sur les Machines (Bachelier, Paris, 1824). English translation (Peter Smith, Gloucester, MA, 1977)

    Google Scholar 

  27. R. Clausius, Philos. Mag. (Ser. 4) 2, 1, 102 (1851). 24, 81, 201 (1862)

    Google Scholar 

  28. L. Kelvin (W. Thomson), Mathematical and Physical Papers of William Thomson, vol. 1 (Cambridge University Press, Cambridge, 1882), pp. 100–106, 113–140, 174–200

    Google Scholar 

  29. L. Onsager, Phys. Rev. 37, 405 (1931). 38, 2265 (1931)

    Google Scholar 

  30. J. Meixner, Ann. Phys. (5) 39, 333 (1941). J. Meixner, H.G. Reik, Thermodynamik der irreversiblen Prozesses, in Handbuch der Physik, vol. 3, ed. by S. Flügge (Springer, Berlin, 1959)

    Google Scholar 

  31. I. Prigogine, Étude Thermodynamiques des Phénomènes Irreversible (Desor, Liège, 1947). Thermodynamics of Irreversible Processes (Interscience, New York, 1967)

    Google Scholar 

  32. L.D. Landau, E.M. Lifshitz, Fluid Mechanics (Pergamon, London, 1959)

    MATH  Google Scholar 

  33. G.K. Batchelor, Fluid Dynamics (Cambridge, London, 1967)

    Google Scholar 

  34. J. Meixner, On the foundation of thermodynamics of processes, in A Critical Review of Thermodynamics. E.B. Stuart, B. Gal-Or, A.J. Brainard (eds.) (Mono Book, Baltimore, 1970), p. 544. Rheol. Acta 12, 465 (1973)

    Google Scholar 

  35. B.C. Eu, J. Chem. Phys. 73, 2958 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  36. R. Nettleton, Phys. Fluids 4, 1488 (1961)

    Article  ADS  MathSciNet  Google Scholar 

  37. I. Müller, Z. Phys. 198, 329 (1967)

    Article  ADS  Google Scholar 

  38. G. Lebon, Bull. Acad. R. Belg. Cl. Sci. 64, 456 (1978)

    Google Scholar 

  39. G. Lebon, D. Jou, J. Casas-Vazquez, J. Phys. A Math. Gen. 13, 275 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  40. L.S. Garcia-Colín, G.J. Fuentes y Matinez, J. Stat. Phys. 29, 387 (1982)

    Google Scholar 

  41. D. Jou, J. Casas-Vazquez, G. Lebon, Extended Irreversible Thermodynamics (Springer, Berlin, 1993)

    Book  MATH  Google Scholar 

  42. I. Müller, T. Ruggeri, Extended Thermodynamics (Springer, Berlin, 1993)

    Book  MATH  Google Scholar 

  43. L.S. Garcia-Colín, M. Lopez de Haro, R.F. Rodriguez, J. Casas-Vazquez, D. Jou, J. Stat. Phys. 37, 465 (1984)

    Article  ADS  Google Scholar 

  44. R.M. Velasco, L.S. Garcia-Colín, J. Non-Equil. Thermodyn. 18, 157 (1993)

    Google Scholar 

  45. B.C. Eu, Kinetic Theory and Irreversible Thermodyanamics (Wiley, New York, 1992)

    Google Scholar 

  46. B.C. Eu, Nonequilibrium Statistical Mechanics (Kluwer, Dordrecht, 1998)

    Book  MATH  Google Scholar 

  47. B.C. Eu, Chem. Phys. Lett. 143, 65 (1988)

    Article  ADS  Google Scholar 

  48. W.T. Grandy, Phys. Rep. 62, 175 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  49. B.C. Eu, Phys. Rev. E 51, 768 (1995)

    Article  ADS  Google Scholar 

  50. I.N. Sneddon, Elements of Partial Differential Equations (McGraw-Hill, New York, 1957)

    MATH  Google Scholar 

  51. D.G.B. Edelen, Applied Exterior Calculus (Wiley, New York, 1985)

    MATH  Google Scholar 

  52. C. von Westenholz, Differential Forms in Mathematical Physics (North-Holland, Amsterdam, 1978)

    MATH  Google Scholar 

  53. H. Flanders, Differential Forms (Academic, New York, 1963)

    MATH  Google Scholar 

  54. M. Chen, B.C. Eu, J. Math. Phys. 34, 3012 (1993). For mathematical aspects of differential forms and their integrability see Refs. [51, 52, 53] quoted above

    Google Scholar 

  55. L. Rayleigh, Theory of Sound (Dover, New York, 1949)

    MATH  Google Scholar 

  56. P. Glansdorff, I. Prigogine, Thermodynamic Theory of Structure, Stability, and Fluctuations (Wiley, New York, 1971)

    MATH  Google Scholar 

  57. I. Procaccia, D. Ronis, M.A. Collins, J. Ross, I. Oppenheim, Phys. Rev. A 19, 1290 (1979)

    Article  ADS  Google Scholar 

  58. H. Brenner, Phys. A 349, 11 (2005). 349, 60 (2005). H. Brenner, Phys. Rev. E 70, 061201 (2004)

    Google Scholar 

  59. G.F. Voronoi, Zeit. Reine Angew. Math. 34, 198 (1908)

    MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Byung Chan Eu .

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Eu, B.C. (2016). Thermodynamic Theory of Irreversible Processes. In: Kinetic Theory of Nonequilibrium Ensembles, Irreversible Thermodynamics, and Generalized Hydrodynamics. Springer, Cham. https://doi.org/10.1007/978-3-319-41147-7_2

Download citation

Publish with us

Policies and ethics