Skip to main content

The Representation of Time in Discrete Mechanics

  • Chapter
  • First Online:
Book cover Time of Nature and the Nature of Time

Part of the book series: Boston Studies in the Philosophy and History of Science ((BSPS,volume 326))

Abstract

The starting point of the chapter is a twofold observation:

  1. (i)

    most current physical theories make use of a continuous parameter t that plays the role of time;

  2. (ii)

    the current practice of modeling, namely, of computational modeling, makes use of a discrete parameter t k (k = 1, 2, …, n) that plays the role of time, because computers cannot handle continuous quantities.

A “parameter” here is a symbol that does not play the same role as the other symbols in the physical theories or models: whereas the other symbols represent physical quantities, t (or t k ) just plays the role of that within which other quantities evolve. It is “time”, but in a rather shallow sense.

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 119.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 159.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 159.99
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.

    We would like to emphasize that, at the beginning of the 1980s, Tsung-Dao Lee (1983, 1987) developed a Discrete Mechanics for another reason still. It was to solve the well-known divergence problems of Quantum Field Theory. He wanted to build up a Discrete Mechanics conceived as a first step toward fully discrete fundamental theories in which divergences could not occur. The idea was first to try to develop one discrete theory in order to generalize it to others. To our knowledge, it did not go very far toward QFT.

  2. 2.

    We would like to emphasize that there is no single way to define a discrete Lagrangian for a mechanical system. The choice of another discrete Lagrangian for the simple pendulum leads to a different equation of motion that corresponds to another variational integrator . In other words, it can be admitted that there are as many DMs – or versions of DM – as ways to define discrete Lagrangians. We go back to this point in the conclusion of the paper.

  3. 3.

    We emphasize that there is not a unique DM but a family of DM depending on the (initial) value of the discrete time step. Thus, when we claim that DM and continuous are empirically equivalent, we mean more precisely that there exists at least one DM that is empirically equivalent to continuous mechanics.

  4. 4.

    Our emphasis.

  5. 5.

    For details on the Noether theorem, see (Butterfield 2006b).

  6. 6.

    Newton-Smith then asks: “How ought we to respond to this undecidability result?” (1980, p. 126). According to him, the statement “time is continuous” can be interpreted as an empirical statement whose truth value is impossible to assess because there is no evidence to be had about it – this is “the Ignorance response”. However, it can also be interpreted as a non-empirical statement describing a theoretical framework within which to describe empirical facts – this is “the Arrogance response”. We do not discuss these interpretations since, as we make clear below in this Section, we deny that any conclusion about the undecidabilty of the structure of time can be drawn from physical theories, like DM, which represent time as discrete.

  7. 7.

    His line of inquiry thus continues Carnap’s, who examined the following question: “Because irrational numbers are always the results of calculations, never the result of direct measurement, might it not be possible in physics to abandon irrational numbers altogether and work only with the rational numbers? That is certainly possible, but it would be a revolutionary change.” (Carnap 1966, p. 88–89) As clear in the main text, this question is irrelevant to our own inquiry about the implications of the discrete representation of time in DM.

  8. 8.

    “Given the importance of conserving integrals of motion and the important role played by the Hamiltonian structure in the reduction procedure for a system with symmetry, one might hope to find an algorithm that combines all of the desirable properties: conservation of energy , conservation of momenta (and other independent integrals), and conservation of the symplectic structure. However, one cannot do all three of these things at once unless one relaxes one or more of the conditions. [...] It is interesting to note that when adaptive time steps are used, the arguments above no longer apply and indeed in this case it is possible to find integrators that are, in an appropriate sense, symplectic , energy preserving and momentum preserving, as shown in (Kane et al. 1999).” (Marsden 2009, p. 178).

  9. 9.

    There are further aspects to the use of time symbols in discrete models. On the one hand, there are numerical methods with “stage” like the Runge-Kutta methods. Here, intermediate discrete instants t k,i , where i is an integer, are introduced between two consecutive instants t k and t k +h. Two levels of time discretization are thus used. On the other hand, there are discrete models with adaptive discrete time symbols: the time step between two consecutive discrete time symbols t k and t k+1 is variable . At each discrete instant, the time step is adapted according to criteria of computational accuracy.

  10. 10.

    All these notions differ from “the time of the simulation”, i.e., the time spent by the computer to solve the discrete model’s equations. In the remaining part of the paper, we shall put this notion aside.

  11. 11.

    We interpret k in a temporal way, as the “time of the model”. However, it seems that one might also interpret it in a spatial way since the change “k → k + 1” governs also the evolution of the position x k to x k+1 . Nevertheless, at least within DM, k seems to be more a temporal parameter than a spatial one. One reason would be that, to our knowledge, the parameter k shares with t k an order property: when k&amp;amp;amp;amp;amp;lt;k+1 one cannot have t k >t k+1 . In other words, k and t k have the same direction. On contrary, while k&amp;amp;amp;amp;amp;lt;k+1, one can have x k <x k+1 as well as x k > x k+1 . Besides, when discrete equations are implemented on computers, there might be other kinds of ``clocks”. For example, the computation of the coordinates (x k+1 , t k+1 ) might be implemented in a device before the computation of the coordinates (x k , t k ). There would be a clock that commands to use k+1 before k. Since we do not discuss implementation issues and focus on the equations of DM, we leave aside this point.

  12. 12.

    In the nineteenth century, pear is known in France to be a metaphor for bourgeois monarchy. This caricature shows thus the affinity of Thiers with this political regime.

References

  • Butterfield, J. 2006a. On symplectic reduction in classical mechanics. In The handbook of philosophy of physics, ed. J. Earman and J. Butterfield, 1–131. North Holland: Elsevier.

    Google Scholar 

  • ———. 2006b. On symmetries and conserved quantities in classical mechanics. In Physical theory and its interpretation, ed. W. Demopoulos and I. Pitowsky, 43–99. Dordrecht: Springer.

    Chapter  Google Scholar 

  • Carnap, R. 1966. Philosophical foundations of physics. New York/London: Basic Books.

    Google Scholar 

  • Cieslinski, J.L., and B. Ratkiewicz. 2006. On simulations of the classical harmonic oscillator equation by difference equations. Adv. Difference Equ.: 40171.

    Google Scholar 

  • Colyvan, M. 2001. The indispensability of mathematics. Oxford: Oxford University Press.

    Book  Google Scholar 

  • D’Innocenzo, A., L. Renna, and P. Rotelli. 1987. Some studies in discrete mechanics. European Journal of Physics 8: 245–252.

    Article  Google Scholar 

  • Feng, K., and M. Qin. 2010. Symplectic geometric algorithms for hamiltonian systems. Heidelberg/Dordrecht/London/New York: Springer.

    Book  Google Scholar 

  • Ge, Z. 1991. Equivariant symplectic difference schemes and generating functions. Physica D 49: 376–386.

    Article  Google Scholar 

  • Ge, Z., and J.E. Marsden. 1988. Lie-Poisson Hamilton-Jacobi theory and Lie-Poisson integrators. Physics Letters A 133 (3): 134–139.

    Article  Google Scholar 

  • Greenspan, D. 1973. Discrete models. London: Addison-Wesley Publishing Company.

    Google Scholar 

  • Hairer, E., C. Lubich, and G. Wanner. 2006. Geometric numerical integration: Structure-preserving algorithms for ordinary differential equations. Berlin/Heidelberg: Springer.

    Google Scholar 

  • Humphreys, P. 2004. Extending ourselves: computational science, empiricism, and scientific method. New-York: Oxford University Press.

    Book  Google Scholar 

  • Kane, C., J.E. Marsden, and M. Ortiz. 1999. Symplectic-energy-momentum preserving variational integrators. Journal of Mathematical Physics 40: 3353–3371.

    Article  Google Scholar 

  • Kibble, T.W.B., and F.H. Berkshire. 2009. Classical mechanics. London: Imperial College Press.

    Google Scholar 

  • Kuorikoski, J. 2012. Mechanisms, modularity and constitutive explanation. Erkenntnis 77 (3): 361–380.

    Article  Google Scholar 

  • LaBudde, R.A., and D. Greenspan. 1974. Discrete mechanics-A general treatment. Journal of Computational Physics 15: 134–167.

    Article  Google Scholar 

  • Lee, T.D. 1983. Can time be a discrete dynamical variable? Physics Letters 122B (3–4): 217–220.

    Article  Google Scholar 

  • ———. 1987. Difference equations and conservation laws. Journal of Statistical Physics 46 (5–6): 843–860.

    Article  Google Scholar 

  • Lee, T., M. Leok, and H. McClamroch. 2009. Discrete control systems. In Encyclopedia of complexity and systems science, ed. Robert A. Meyer, 2002–2019. New York: Springer.

    Chapter  Google Scholar 

  • Maddy, P. 1997. Naturalism in mathematics. Oxford: Clarendon Press.

    Google Scholar 

  • Marsden, J.E. 2009. Lectures on mechanics. Cambridge: Cambridge University Press.

    Google Scholar 

  • Marsden, J.E., and M. West. 2001. Discrete mechanics and variational integrators. Acta Numer 10: 357–514.

    Article  Google Scholar 

  • Newton-Smith, W.H. 1980. The structure of time. London: Routledge & Kegan Paul.

    Google Scholar 

  • North, J. 2009. The structure of physics: A case study. Journal of Philosophy 106 (2): 57–88.

    Article  Google Scholar 

  • Stern, A. and M. Desbrun. 2008. Discrete geometric mechanics for variational time integrators. In: Discrete differential geometry: An applied introduction, Siggraph 2006 Course Notes, Chap. 15.

    Google Scholar 

  • Suppes, P. 1957. Introduction to logic. New York: Van Nostrand Reinhold.

    Google Scholar 

  • Wilson, J. 2007. Newtonian forces. The British Journal for Philosophy of Science 58: 173–205.

    Article  Google Scholar 

Download references

Acknowledgements

We wish to thank the participants of the conference “The Time of Nature, The Nature of Time” for comments and discussion. We are most grateful to Christophe Bouton and Philippe Huneman for helpful comments on previous drafts of this paper. We would like also to thank Paul Humphreys for precious suggestions and remarks on the paper. The authors are grateful to Andrew McFarland for a thorough language check of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincent Ardourel .

Editor information

Editors and Affiliations

Appendices

Appendices

1.1 Appendix 1: Discrete Mechanics and Some Applications

Let us call A d the action in Discrete Mechanics (for details see (D’Innocenzo et al. 1987; Marsden and West 2001, Chap. 5)):

$$ {A}_d=\sum_{k=0}^{N-1}{L}_d.\left({t}_{k+1}-{t}_k\right)\kern5.5em $$
(9.15)

with L d the discrete Lagrangian. The principle of least action states that: δA d  = 0. It results in the following twofold discrete Euler-Lagrange equations (DEL). The first discrete Euler-Lagrange equation is:

$$ \left({t}_k-{t}_{k-1}\right)\frac{\partial {L}_d\left({t}_{k-1},{q}_{k-1},{t}_k,{q}_k\right)}{\partial {q}_k}+\left({t}_{k+1}-{t}_k\right)\frac{\partial {L}_d\left({t}_k,{q}_k,{t}_{k+1},{q}_{k+1}\right)}{\partial {q}_k}=0 $$
(9.16)

The second discrete Euler-Lagrange equation is:

$$ \frac{\partial }{\partial {t}_k}\left[\left({t}_k-{t}_{k-1}\right){L}_d\left({t}_{k-1},{q}_{k-1},{t}_k,{q}_k\right)\right]+\frac{\partial }{\partial {t}_k}\left[\left({t}_{k+1}-{t}_k\right){L}_d\left({t}_k,{q}_k,{t}_{k+1},{q}_{k+1}\right)\right]=0\kern1em $$
(9.17)

Let us follow d’Innocenzo et al. for the choice of the discrete Lagrangian:

$$ {L}_d\left({q}_k,{q}_{k+1},{h}_{k+1}\right)=\frac{1}{2} m{\left(\frac{q_{k+1}-{q}_k}{h_{k+1}}\right)}^2- V\left(\frac{q_{k+1}+{q}_k}{2}\right) $$

with h k+1 =t k+1 −t k . Under these conditions, let us solve (i) the free particle system, (ii) the one dimension falling body problem, and (iii) the one dimension harmonic oscillator system.

(i) The free particle is the system with V=V 0 . The DEL become:

$$ m\frac{v_{x, k+1}-{v}_{x, k}}{h_{k+1}}=0\kern2.5em $$
(9.18)
$$ \frac{1}{2} m{v}_{x, k+1}^2=\frac{1}{2} m{v}_{x, k}^2\kern3.25em $$
(9.19)

with \( {v}_{x, k}=\frac{x_k-{x}_{k-1}}{t_k-{t}_{k-1}} \). The solution of the equations is:

$$ {x}_k={v}_{x, i}{t}_k+{x}_i\kern3em $$
(9.20)

with the initial conditions t 0 =0, x i =x 0 , v x,i =(x 1 −x 0 )/h 1 .

(ii) The one dimension falling body problem is the system where \( \left({z}_k,{z}_{k+1}\right)= mg\frac{z_{k+1}+{z}_k}{2} \). Thus, the DEL are:

$$ {v}_{z, k+1}-{v}_{z, k}=-\frac{g}{2}\left({t}_{k+1}-{t}_{k-1}\right)\kern1.5em $$
(9.21)
$$ \frac{1}{2} m{v}_{z, k+1}^2+ mg\frac{z_{k+1}+{z}_k}{2}=\frac{1}{2} m{v}_{z, k}^2+ mg\frac{z_k+{z}_{k-1}}{2}\kern0.5em $$
(9.22)

with \( {v}_{z, k}=\frac{z_k-{z}_{k-1}}{t_k-{t}_{k-1}} \). We follow the resolution of d’Innocenzo et al. (1987) with different notations. The solution of the equations is t k  − t k − 1 = h and:

$$ {z}_k=-\frac{1}{2} g{t}_k^2+\left({v}_{z, i}+ gh\right){t}_k+{z}_{i\ }\kern1.25em $$
(9.23)

with the initial conditions t 0 =0, z i =z 0 , v z,i =(z 1 −z 0 )/h 1 .

(iii) The one dimension harmonic oscillator system is the system where \( V\left({x}_k,{x}_{k+1}\right)=\frac{1}{2} K{\left(\frac{x_{k+1}+{x}_k}{2}\right)}^2 \). Thus, the DEL are:

$$ m\left({v}_{x, k+1}-{v}_{x, k}\right)=-\frac{K}{4}\left[\left({x}_k+{x}_{k-1}\right){h}_k+\left({x}_{k+1}+{x}_k\right){h}_{k+1}\right]\kern1.25em $$
(9.24)
$$ \frac{1}{2} m{v}_{x, k+1}^2+\frac{1}{2} K{\left(\frac{x_{k+1}+{x}_k}{2}\right)}^2=\frac{1}{2} m{v}_{x, k}^2+\frac{1}{2} K{\left(\frac{x_k+{x}_{k-1}}{2}\right)}^2 $$
(9.25)

Following d’Innocenzo et al. (1987) with different notations (in particular see (Cieslinski and Ratkiewicz 2006), we have:

$$ {x}_k={x}_0 \cos \left({\omega}_d{t}_k\right)+\frac{x_1-{x}_0 \cos \left({\omega}_d h\right)}{ \sin \left({\omega}_d h\right)} \sin \left({\omega}_d{t}_k\right) $$

with \( {\omega}_d=1/ h\ { \tan}^{-1}\left(\frac{\omega h}{1-{\omega}^2{h}^2/4}\right) a n d\ \omega =\sqrt{K/ m} \).

1.2 Appendix 2: Trajectory of a Falling Body in DM

Since x k  = v x, i t k  + x i  (see Eq. (9.20) in Appendix) then, \( {t}_k=\frac{x_k-{x}_i}{v_{x, i}} \). Hence, put it in the Eq. (9.23) in Appendix, we derive the equation of the trajectory:

$$ {z}_k=-\frac{1}{2} g{t}_k^2+\left({v}_{z, i}+ gh\right){t}_k+{z}_i = -\frac{g}{2{v}_{x, i}^2}{\left({x}_k-{x}_i\right)}^2+\frac{v_{z, i}+ gh}{v_{x, i}}\left({x}_k-{x}_i\right)+{z}_{i\ } $$

Now, we can deduce the highest position of the body. It is the position where the partial derivative z k with respect to x k vanishes: \( \frac{\partial {z}_k}{\partial {x}_k}\left({x}_k^{\ast}\right)=0 \). Hence,

$$ {x}_k^{\ast }={v}_{x, i}\left({v}_{z, i}+ gh\right)/ g+{x}_i\kern1em and\kern0.75em {z}_k^{\ast }={z}_k\left({x}_k^{\ast}\right)={\left({v}_{z, i}+ gh\right)}^2/\left(2 g\right)+{z}_i $$

1.3 Appendix 3: Ge-Marsden Theorem

We report here the Ge-Marsden theorem and its proof as they are formulated in the original paper:

We recall that there are algorithms which exactly preserve energy, some of which also preserve other quantities […]. However, these algorithms cannot be symplectic , according to the following result of Ge:

Let H be a Hamiltonian which has no conserved quantities (in a given class \( \mathcal{H} \), for example analytic functions) other than functions of H. That is, if {F,H}=0, then F(z)=F 0 (H(z)) for a function F 0 . Let ΦΔt be an algorithm which is defined for small Δt and is smooth. If this algorithm is symplectic , and conserved H exactly, then it is the time advance map for the exact Hamiltonian system up to a reparametrization of time. In other words, approximate symplectic algorithms cannot preserve energy for nonintegrable systems.

This result is in fact easy to prove. The algorithm being symplectic , is generated by a dependent function F(z,t), which we assume belong to \( \mathcal{H} \). Since ΦΔt preserves H, and H is assume to be time independent, F commutes with H, and so F(z)=F 0 (H(z)). It follows that the hamiltonian vector fields of F and H are parallel, so their integral curves are related by a time reparametrization. (Ge and Marsden 1988, p. 135).

As far we understand this result, while the Hamiltonian system is assumed to be nonintegrable – i.e that an exact solution cannot be constructed – if the symplectic  algorithm exactly preserves energy, the solution of the algorithm would be the exact solution of the system modulo a reparametrization. This contradicts the assumption according to which the system is nonintegrable (see also (Ge 1991, p. 380; Marsden 2009, p. 178)). However, as Marsden (2009, p. 179) emphasizes, if the time step is variable , the previous result does not hold and symplectic algorithms can exactly preserve energy.

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing AG

About this chapter

Cite this chapter

Ardourel, V., Barberousse, A. (2017). The Representation of Time in Discrete Mechanics. In: Bouton, C., Huneman, P. (eds) Time of Nature and the Nature of Time. Boston Studies in the Philosophy and History of Science, vol 326. Springer, Cham. https://doi.org/10.1007/978-3-319-53725-2_9

Download citation

Publish with us

Policies and ethics