Skip to main content
Log in

Elementary Thermo-mechanical Systems and Higher Order Constraints

  • Published:
Qualitative Theory of Dynamical Systems Aims and scope Submit manuscript

Abstract

In this paper we study a class of physical systems that combine a finite number of mechanical and thermodynamic observables. We call them elementary thermo-mechanical systems (ETMS). We introduce these systems by means of simple examples, and we obtain their (time) evolution equations by using, essentially, the Newton’s laws and the First Law of Thermodynamics only. We show that such equations are similar to those defining certain constrained mechanical systems. From that, and addressing the main aim of the paper, we give a general definition of ETMS, in a variational formalism, as a particular subclass of the Lagrangian higher order constrained systems.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Physical systems whose states depend on mechanical and thermodynamic variables are more common in the domain of continuous media or fluid dynamics, but not in contexts where only a finite number of degrees of freedom are involved.

  2. By external force we mean any non-conservative force, or a conservative force whose corresponding potential energy is not included in what we consider the mechanical energy of the system.

  3. The thermal conductivity of the material, of which the wagon is made, is supposed to be big enough to ensure that the temperature is well defined and is uniformly distributed for all time.

  4. Note that \(E_{mec}\) is constituted, besides the kinetic energy, just by the potential energy related to \(F_{g}=-mg\). We are not including the potential energy related to \(F_{e}=kA^{-\gamma +1}x^{-\gamma }\) [see Eq. (22)] because, as previously mentioned, we are taking it as an external force.

  5. Using NDSolve in Wolfram Mathematica.

  6. The Newton’s equation has been derived by our generalization of D’Alembert Principle without using the second order kinematic constraint [see Eq. (41)]. Using Newton’s equation, we can replace such a constraint by a first order constraint: \(-\frac{\mu ({\dot{x}})^{2}}{2}+U=0\). Thus, we can describe our ETMS also as a HOCS of order (1, 1).

  7. Recall that, given a Lagrangian function \(L:TQ\rightarrow {\mathbb {R}}\), its energy \(E:TQ\rightarrow {\mathbb {R}}\) is given by

    $$\begin{aligned} E\left( v\right) =\left\langle {\mathbb {F}}L\left( v\right) ,v\right\rangle -L\left( v\right) ,\quad \forall v\in TQ, \end{aligned}$$

    being \({\mathbb {F}}L:TQ\rightarrow T^{*}Q\) the fiber derivative of L, i.e. the Legendre transformation related to L.

  8. The local representative of \(j^{\left( 2\right) }\) (in the above mentioned local charts) is

    $$\begin{aligned} j^{\left( 2\right) }\left( q,{\dot{q}},\ddot{q}\right) =\left( q,{\dot{q}},{\dot{q}},\ddot{q}\right) . \end{aligned}$$

References

  1. Abraham, R., Marsden, J.E.: Foundations of Mechanics. The Benjamin/Cummings Publishing Company, San Francisco (1978)

    MATH  Google Scholar 

  2. Abraham, R., Marsden, J.E., Ratiu, T.: Manifolds, Tensor Analysis and Aplications. Addison-Wesley, London (1983)

    MATH  Google Scholar 

  3. Amaya, M.P.: Introducción a los sistemas termo-mecánicos desde una perspectiva simpléctica. Ph.D. thesis, Universidad Nacional del Sur, Bahía Blanca, Argentina (2014)

  4. Arnold, V.I.: Mathematical Methods of Classical Mechanics. Springer, New York (1989)

    Book  Google Scholar 

  5. Bloch, A.M., Marsden, J.E., Zenkov, D.V.: Nonholonomic dynamics. Not. Am. Math. Soc. 52, 324–333 (2005)

    MathSciNet  MATH  Google Scholar 

  6. Boothby, W.M.: An Introduction to Differentiable Manifolds and Riemannian Geometry. Academic Press, New York (1985)

    MATH  Google Scholar 

  7. Callen, H.B.: Thermodynamics and an Introduction to Thermostatistics, 2nd edn. Wiley, New York (1985)

    MATH  Google Scholar 

  8. Cendra, H., Ferraro, S., Grillo, S.: Lagrangian reduction of generalized nonholonomic systems. J. Geom. Phys. 58, 1271–1290 (2008)

    Article  MathSciNet  Google Scholar 

  9. Cendra, H.: Generalized nonholonomic mechanics, servomechanisms and related brackets. J. Math. Phys. 47, 022902 (2006)

    Article  MathSciNet  Google Scholar 

  10. Cendra, H., Grillo, S.: Lagrangian systems with higher order constraints. J. Math. Phys. 48, 052904 (2007)

    Article  MathSciNet  Google Scholar 

  11. Cendra, H., Ibort, A., de León, M., Martín de Diego, D.: A generalization of Chetaev’s principle for a class of higher order nonholonomic constraints. J. Math. Phys. 45(7), 2785 (2004)

    Article  MathSciNet  Google Scholar 

  12. Crampin, M., Sarlet, W., Cantrijn, F.: Higher-order differential equations and higher-order Lagrangian mechanics. Math. Proc. Camb. Philos. Soc. 99, 565–587 (1986)

    Article  MathSciNet  Google Scholar 

  13. de León, M., Rodrigues, P.R.: Generalized Classical Mechanics and Field Theory. North-Holland, Amsterdam (1985)

    MATH  Google Scholar 

  14. Dobronravov, V.V.: The fundamentals of the mechanics of nonholonomic systems. Vysshaya Shkola 271 (1970)

  15. Eberard, D., Maschke, B.M., van der Schaft, A.J.: An extension of hamiltonian systems to the thermodynamic phase space: towards a geometry of nonreversible processes. Rep. Math. Phys. 60(2), 175–198 (2007)

    Article  MathSciNet  Google Scholar 

  16. Fermi, E.: Thermodynamics. Dover Publications, Mineola (1956)

    MATH  Google Scholar 

  17. Gay-Balmaz, F., Yoshimura, H.: A lagrangian variational formulation for nonequilibrium thermodynamics. Part I: discrete systems. J. Geom. Phys. 111, 169–193 (2017)

    Article  MathSciNet  Google Scholar 

  18. Gay-Balmaz, F., Yoshimura, H.: Dirac structures in nonequilibrium thermodynamics. J. Math. Phys. 58, 012701 (2018)

    Article  MathSciNet  Google Scholar 

  19. Goldstein, H., Poole, C., Safko, P.: Classical Mechanics, 3rd edn. Wiley, New York (2000)

    MATH  Google Scholar 

  20. Grillo, S.: Higher order constrained Hamiltonian systems. J. Math. Phys. 50, 082901 (2009)

    Article  MathSciNet  Google Scholar 

  21. Grillo, S., Zuccalli, M.: Variational reduction of Lagrangian systems with general constraints. J. Geom. Mech. 4(1), 49–88 (2012)

    Article  MathSciNet  Google Scholar 

  22. Kobayashi, S., Nomizu, K.: Foundations of Differential Geometry. Wiley, New York (1963)

    MATH  Google Scholar 

  23. Marsden, J.E., Ratiu, T.: Introduction to Mechanics and Symmetry. Springer, New York (1994)

    Book  Google Scholar 

  24. Moran, M.J., Shapiro, H.N., Boettner, D.D., Bailey, M.B.: Fundamentals of Engineering Thermodynamics, 7th edn. Wiley, New York (2011)

    Google Scholar 

  25. Mrugala, R.: On contact and metric structures on thermodynamic spaces. Departmental Bulletin Paper, pp. 167–181 (2000)

  26. Mrugala, R., Nulton, J.D., Schön, J.C., Salamon, P.: Contact structure in thermodynamic theory. Rep. Math. Phys. 29, 109–121 (1991)

    Article  MathSciNet  Google Scholar 

  27. Neimark, J.I., Fufaev, N.A.: Dynamics of Non-holonomic Systems. Translations of Mathematical Monographs, vol. 33. American Mathematical Society, Providence (1972)

    Google Scholar 

  28. Planck, M.: Treatise on Thermodynamics. Dover Publications, Mineola (1917)

    Google Scholar 

  29. Śniatycki, J., Tulczyjew, W.M.: Generating forms of Lagrangian submanifolds. Indiana Univ. Math. J. 22(3), 267 (1972)

    Article  MathSciNet  Google Scholar 

  30. Stueckelberg, E.C.G., Scheurer, P.B.: Thermocinétique phénoménologique galiléenne. Birkhäuser, Basel (1974)

    MATH  Google Scholar 

  31. Tulczyjew, W.M.: The Legendre transformation. Ann. l’Inst. Henri Poincaré Sect. A XXVII, 101 (1977)

    MathSciNet  MATH  Google Scholar 

  32. Tulczyjew, W.M., Urbański, P.: A slow and careful Legendre transformation for singular Lagrangians. Acta Phys. Pol. Ser. B 30(10), 2909–2978 (1999)

    MathSciNet  MATH  Google Scholar 

  33. Zemansky, M.: Heat and Thermodynamics: An Intermediate Textbook, 5th edn. McGraw-Hill, New York (1967)

    MATH  Google Scholar 

Download references

Acknowledgements

The authors thank CONICET for its financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sergio Grillo.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A: Background on Thermodynamics

Appendix A: Background on Thermodynamics

Below, we introduce the basic notation and terminology on thermodynamics that we shall use along all of the paper, and recall some fundamental concepts on the subject (see [7, 16, 24, 28]).

\(*\) A thermodynamic system is typically defined by \(2d+1\) observables, which we shall denote \(x_{1},\ldots ,x_{d-1},T,y_{1},\ldots ,y_{d-1},S\) and U. The \(x_{i}\)’s and T (resp. \(y_{i}\)’s, S and U) are called intensive (resp. extensive) variables. The extensive variables depend on the “size” of the system, while the intensive ones do not. U is the internal energy, T is the temperature and S the entropy. For example, consider a mixture of \(r\in {\mathbb {N}}\) different chemical components. In this case, we have that:

  • \(d=r+2\);

  • for \(i=1,\ldots ,r\), each variable \(y_{i}:=N_{i}\) (resp. \(-x_{i}:=\mu _{i}\)) represents the number of moles (resp. the chemical potential) of a given type;

  • \(x_{r+1}=:P\) is the pressure and \(y_{r+1}=:V\) the volume.

Remark I

Sometimes, systems can be seen as composed by “simpler” ones, i.e. those defined by a smaller number of variables. We say in this case that such a system is a composite system. In the last example, each chemical compound can be seen as a part of a composite system.

\(*\) The possible values of the variables \(x_{1},\ldots ,x_{d-1},T,y_{1},\ldots ,y_{d-1},S\) and U give rise to a manifold \({\mathcal {T}}\subset {\mathbb {R}}^{2d+1}\) (that we shall assume to be open), usually called the thermodynamical phase space (TPS): the set of states of the system. We can see these variables as coordinates for \({\mathcal {T}}\).

\(*\) The manifold \({\mathcal {T}}\) is a contact manifold with contact form

$$\begin{aligned} \theta :=dU-T\ dS+{\displaystyle \sum \limits _{i=1}^{d-1}}x_{i}\,dy_{i}. \end{aligned}$$
(A.1)

Thus, \(\left( x_{1},\ldots ,x_{d-1},-T,y_{1},\ldots ,y_{d-1},S,U\right) \) defines a global Darboux system for \(\left( {\mathcal {T}},\theta \right) \), see [4, 26].

Remark II

For a composite system (see Remark I) formed out by two simpler ones, the TPS is a product manifold \({\mathcal {T}}={\mathcal {T}}_{1}\times {\mathcal {T}}_{2}\) with contact form

$$\begin{aligned} \theta :=dU_{1}-T_{1}\ dS_{1}+dU_{2}-T_{2}\ dS_{2}+{\displaystyle \sum \limits _{i=1}^{d_{1}-1}}x_{1,i}\,dy_{1,i}+{\displaystyle \sum \limits _{i=1}^{d_{2}-1}}x_{2,j}\,dy_{2,j}.\nonumber \\ \end{aligned}$$
(A.2)

Here \(\left( x_{k,1},\ldots ,x_{k,d_{1}-1},-T_{k},y_{k,1},\ldots ,y_{k,d_{1}-1},S_{k},U_{k}\right) \), with \(k=1,2\), are the global Darboux coordinates of \({\mathcal {T}}_{1}\) and \({\mathcal {T}}_{2}\).

\(*\) By process we shall mean every curve \(\Gamma :\left[ a,b\right] \rightarrow {\mathcal {T}}\). It represents a “continuum” of actions on the system that produce a “continuum” of changes on its states.

\(*\) Among the states, a special role is played by a subset \({\mathcal {N}}\subset {\mathcal {T}}\), known as the space of equilibrium states, which is defined by the following two conditions on U. The first one says that, on the equilibrium states, U and the rest of the extensive variables \(y_{i}\)’s and S must be related by the formula

$$\begin{aligned} U=\Phi \left( y_{1},\ldots ,y_{d-1},S\right) , \end{aligned}$$
(A.3)

for some function \(\Phi \) (typically homogeneous of degree one). Equation above is known as the Fundamental Equation of the system. The second condition says that, for any differentiable curve \(\Gamma :\left[ a,b\right] \rightarrow {\mathcal {N}}\subset {\mathcal {T}}\), the variation \(\Delta U:=U\left( \Gamma \left( b\right) \right) -U\left( \Gamma \left( a\right) \right) \) must satisfies

$$\begin{aligned} \Delta U={\mathbb {Q}}-W, \end{aligned}$$
(A.4)

where \({\mathbb {Q}}={\mathbb {Q}}\left( \Gamma \right) \) and \(W=W\left( \Gamma \right) \) are the heat and the mechanical work, respectively, interchanged by the system and the environment along the process \(\Gamma \). This is the First Law of Thermodynamics.

Remark

When \({\mathbb {Q}}=0\) along a process, one says that such a process is adiabatic.

At a differential level, Eq. (A.4) translates to

(A.5)

Here, and are 1-forms (non-necessarily closed) on \({\mathcal {T}}\) such that, given a process \(\Gamma \),

Idem for . For instance, for a mixture of chemical components (see above), and are given by

at least for some processes. Accordingly,

$$\begin{aligned} dU=T\ dS-P\ dV-{\displaystyle \sum \limits _{i=1}^{r}}x_{i}\,dy_{i}. \end{aligned}$$

In general, we must have

$$\begin{aligned} dU=T\ dS-{\displaystyle \sum \limits _{i=1}^{d-1}}x_{i}\,dy_{i}. \end{aligned}$$
(A.6)

Combining Eqs. (A.3) and (A.6), it follows that the subset \({\mathcal {N}}\) is defined by the equations

$$\begin{aligned} x_{i}= & {} -\frac{\partial \Phi }{\partial y_{i}}\left( y_{1},\ldots ,y_{d-1},S\right) ,\nonumber \\ T= & {} \frac{\partial \Phi }{\partial S}\left( y_{1},\ldots ,y_{d-1},S\right) \quad \text {and }U=\Phi \left( y_{1},\ldots ,y_{d-1},S\right) , \end{aligned}$$
(A.7)

known as state equations. This means that \({\mathcal {N}}\) is a Legendre submanifold of \(\left( {\mathcal {T}},\theta \right) \) (see [4, 26]).

Remark III

As explained in [16], the function \(\Phi \) or, equivalently, the internal energy U, is defined by the allowed mechanical work on the system. In other words, U is completely determined if we know the work done \(W\left( \Gamma \right) \) along any process \(\Gamma \). (This information, in fact, not only determines U, but also \({\mathbb {Q}}\)).

\(*\) For instance, the fundamental equation of the so-called ideal gas, with only one chemical component, is given by 

$$\begin{aligned} \Phi (N,S,V)=N\,u_{0}\left( \frac{N\,v_{0}\ e^{\frac{S-N\,s_{0}}{N\,R}}}{V}\right) ^{\frac{1}{\alpha }}, \end{aligned}$$
(A.8)

where \(\alpha \) is a dimensionless constant, R is the universal constant of ideal gases, and \(s_{0}\), \(v_{0}\) and \(u_{0}\) are constants with units of entropy, volume and energy per mole, respectively (see [7] for more details). Thus, the related state equations read

$$\begin{aligned} \mu =\frac{U}{N}\,\left( 1+\frac{1}{\alpha }\,\left( 1-\frac{S}{NR}\right) \right) ,\quad T=\frac{U}{NR\alpha },\quad P=\frac{U}{V\,\alpha }, \end{aligned}$$
(A.9)

with

$$\begin{aligned} U=N\,u_{0}\left( \frac{N\,v_{0}\ e^{\frac{S-N\,s_{0}}{N\,R}}}{V}\right) ^{\frac{1}{\alpha }}. \end{aligned}$$
(A.10)

Let us mention that the last two equations in (A.9) and the Eq. (A.10) are usually written as

$$\begin{aligned} U=\alpha NRT,\quad PV=NRT\quad \text {and }S=N\,s_{0}+NR\ \mathrm {ln}\left( \frac{T^{\alpha }V}{t_{0}^{\alpha }Nv_{0}}\right) , \end{aligned}$$
(A.11)

where \(t_{0}:=u_{0}/R\alpha \).

\(*\) The differentiable curves along the equilibrium states \(\Gamma :\left[ a,b\right] \rightarrow {\mathcal {N}}\) are usually called quasi-static processes (and we shall take this convention). Note that, along such curves, the state equations (A.7) are satisfied for every \(t\in \left[ a,b\right] \) (by definition of \({\mathcal {N}}\)).

Remark IV

In practice, in order to have a quasi-static process \(\Gamma \), the velocity of \(\Gamma \) must be small (w.r.t. certain characteristic lengths and times related to the microscopic properties of the system). That is to say, the action that defines the process must produce changes in the states at a very slow rate. This justifies the name “quasi-static.” However, in this paper, when we say that a process is quasi-static we will not be assuming that the rate of change of states is necessarily slow. We will only assume that the state equations are satisfied for all time along such a process.

\(*\) Not every process \(\Gamma :\left[ a,b\right] \rightarrow {\mathcal {N}}\) is allowed. According to the Second Law of Thermodynamics, a process \(\Gamma \) must satisfy

(A.12)

where

(A.13)

In infinitesimal terms

(A.14)

For adiabatic processes, since , we must have \(\Delta S\ge 0\).

\(*\) A process \(\Gamma :\left[ a,b\right] \rightarrow {\mathcal {N}}\) is say to be reversible if there exists another process \(\Gamma ^{-}:\left[ a,b\right] \rightarrow {\mathcal {N}}\) such that \(\Gamma ^{-}\left( t\right) =\Gamma \left( a+b-t\right) \). Otherwise, \(\Gamma \) is say to be irreversible. Then, a process is reversible if and only if the equation

(A.15)

holds.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cendra, H., Grillo, S. & Palacios Amaya, M. Elementary Thermo-mechanical Systems and Higher Order Constraints. Qual. Theory Dyn. Syst. 19, 39 (2020). https://doi.org/10.1007/s12346-020-00371-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12346-020-00371-8

Keywords

Navigation