Skip to main content
Log in

Numerical instability and dynamical systems

  • Paper in Philosophy of Science in Practice
  • Published:
European Journal for Philosophy of Science Aims and scope Submit manuscript

Abstract

In philosophical studies regarding mathematical models of dynamical systems, instability due to sensitive dependence on initial conditions, on the one side, and instability due to sensitive dependence on model structure, on the other, have by now been extensively discussed. Yet there is a third kind of instability, which by contrast has thus far been rather overlooked, that is also a challenge for model predictions about dynamical systems. This is the numerical instability due to the employment of numerical methods involving a discretization process, where discretization is required to solve the differential equations of dynamical systems on a computer. We argue that the criteria for numerical stability, as usually provided by numerical analysis textbooks, are insufficient, and, after mentioning the promising development of backward analysis, we discuss to what extent, in practice, numerical instability can be controlled or avoided.

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

Similar content being viewed by others

Notes

  1. A model of a dynamical system can be stable under some respects, and unstable under others. For example, as Corless and Fillion (2013, [554]) show, the Lorenz attractor is unstable with respect to the position, but stable with respect to the dimensions of the attractor. In the paper, the (in)stability in models of dynamical systems we discuss mainly concerns the systems’ trajectories, and, for the sake of readability, this specification will remain implicit.

  2. See also Moir (2013) for discussions on errors and stability in mathematical modeling.

  3. We can define the dynamical system as \(<\mathbb {R}, S, \varphi >\) with S the state space.

  4. Practitioners may find the idea of relative error more useful than absolute error. Although the definition of sensitivity to initial conditions usually involves absolute error, it could certainly be revised.

  5. The definition of the structural instability is a question in itself, which is discussed in Frigg et al. (2014) and Nabergall et al. (2019), Winsberg (2018, 232–246). It is out of the scope of this paper to address this question.

  6. The small differences are non-eliminable since the discretization process changes the nature of the model, in that the continuous differential equation is replaced by a discrete algebraic equation.

  7. Advection is the horizontal transport of properties such as heat, temperature, salinity or polluting material, by displacement of a fluid such as water or air. \(u\cdot \frac {\partial u}{\partial t}\) is the advected quantity over time; in the context of fluid dynamics, it is the fluid velocity. \(\frac {\partial u}{\partial x}\) is its partial derivative against the spatial coordinate x. The fluid velocity, like any quantity in climate models, is decomposed into a Fourier series, i.e. into a sum of cosines and sines. In other words, it is represented as a wave packet, each wave being a perturbation.

  8. So-named because, for the multi-step methods of ordinary differential equations, “zero-stability” can be determined by considering only the differential equation y = 0 (Süli and Mayers 2003, 332). It is also sometimes called D-stability in honor of Dalquist (Hairer et al.1987, 380), or just stability.

  9. This presentation is simplified. In particular, it leaves aside small roundoff errors. For more details, see e.g., (Bréhier 2016, 14; Palais and Palais 2009, 137). Moreover, for linear multi-step methods, one needs to take into account the initial values at the different steps (see Süli and Mayers 2003, 332).

  10. Roughly speaking, the relative forward error (δy = Δy/y) is bounded by the product of condition number with the relative backward error (δx = Δx/x), so that:

    $$ \delta y \leq \kappa \cdot \delta x $$
    (6)

    A problem, or a dynamical system, is called ill-conditioned if κ > 1, and well-conditioned if κ ≤ 1.

  11. There are, however, cases where the use of numerical solutions is preferred although scientists know the exact solutions (Ardourel and Jebeile 2017).

  12. We are in debt to an anonymous reviewer for providing such insights.

  13. Having recourse to adaptive time steps is compatible with resorting to geometric integration (e.g., see Ardourel and Barberousse 2017, 180).

  14. For a recent attempt at showing how tuning the parameters can fix structural instability, see Baldissera Pacchetti (2020).

  15. They borrow this example from Curtiss and Hirschfelder (1952).

  16. There are mathematical results that state the conditions for which a finite difference formula can be zero-stable. This is for instance the case of the Courant-Friedrichs-Lewy (CFL) condition. Informally, the CFL condition states that a finite difference formula can be stable only if its numerical domain of dependence is at least as large as the mathematical domain of dependence (Trefethen 1994, 162).

  17. In the context of partial differential equations, this is linked to the main theorem of numerical analysis, called the Lax Equivalence Theorem. Informally, the Lax Equivalence Theorem says that, as soon as the numerical scheme is consistent and based on a linear finite difference model, stability is a necessary and sufficient condition for convergence (Trefethen 1994, 157). The notion of convergence is connected to consistency (e.g. Trefethen1994, chap. 1, 53) but also to conditionning (e.g. Corless and Fillion 2013, 598; see Section 4.3).

  18. There are mathematical results that state under which conditions a linear multistep numerical method is absolute stable (see Fasshauer 2007, chap. 4, Hairer et al. 1987). The notion of L-stability can be introduced to further elaborate the idea of absolute stability (Bui and Bui 1970, 356).

References

  • Arakawa, A. (1977). Computational aspects of numerical models for weather prediction and climate simulation. In Methods in computational physics: advances in research and applications (p. 17).

  • Ardourel, V., & Barberousse, A. (2017). The representation of time in discrete mechanics. In Bouton, C., & Huneman, P. (Eds.) Time of nature, nature of time. Philosophical perspectives of time in natural sciences. Boston studies in the philosophy and history of science, (Vol. 326 pp. 173–208). Dordrecht: Springer.

  • Ardourel, V., & Jebeile, J. (2017). On the presumed superiority of analytical solutions over numerical methods. European Journal for Philosophy of Science, 7, 201–220.

    Article  Google Scholar 

  • Baer, F. (1961). The extended numerical integration of a simple barotropic model. Journal of Meteorology, 18, 319–339.

    Article  Google Scholar 

  • Baldissera Pacchetti, M. (2020). Structural uncertainty through the lens of model building. Synthese. https://doi.org/10.1007/s11229-020-02727-8.

  • Batterman, R.W. (1993). Defining chaos. Philosophy of Science, 60(1), 43–66.

    Article  Google Scholar 

  • Bradley, S., Frigg, R., Du, H., & Smith, L.A. (2014). Model error and ensemble forecasting: A cautionary tale. In Guichun, C.G., & Chuang, L. (Eds.) Scientific explanation and methodology of science (pp. 58–66). Singapore: World Scientific.

  • Bréhier, C.E. (2016). Introduction to numerical methods for ordinary differential equations. Lecture notes. https://hal.archives-ouvertes.fr/cel-01484274/document.

  • Bui, T.D., & Bui, T.R. (1970). Numerical methods for extremely stiff systems of ordinary differential equations. Applied Mathematical Modelling, 3, 355–358.

    Article  Google Scholar 

  • Corless, R.M. (1992). Defect-controlled numerical methods and shadowing for chaotic differential equations. Physica D: Nonlinear Phenomena, 60 (1-4), 323–334.

    Article  Google Scholar 

  • Corless, R.M., & Fillion, N. (2013). A graduate introduction to numerical methods from the viewpoint of backward error analysis. New York: Springer.

    Book  Google Scholar 

  • Corless, R.M., & Fillion, N. (2019). Backward error analysis for perturbation methods. In Fillion, N., CorlessIS. R.M., & Kotsireas, I.S. (Eds.) Proceedings of 2015 and 2016 ACMES conferences, Fields Institute Communications, (Vol. 82 pp. 35–79). New York: Springer.

  • Curtiss, C., & Hirschfelder, J. (1952). Integration of stiff equations. Proceedings of the National Academy of Sciences, 38, 235–243.

    Article  Google Scholar 

  • Fasshauer, G. (2007). Numerical methods for differential equations/computational mathematics ii (spring 07). handouts and worksheets. http://www.math.iit.edu/~fass/478_578_handouts.html.

  • Fillion, N., & Bangu, S. (2015). Numerical methods, complexity and epistemic hierarchies. Philosophy of Science, 82(5), 941–955.

    Article  Google Scholar 

  • Fillion, N., & Corless, R. (2014). On the epistemological analysis of modeling and computational error in the mathematical sciences. Synthese, 191, 1451–1467.

    Article  Google Scholar 

  • Fillion, N., & Corless, R.M. (2021). Concepts of solution and the finite element method: a philosophical take on variational crimes. Philosophy of Technology, 34, 129–148. https://doi.org/10.1007/s13347-019-00371-w.

    Article  Google Scholar 

  • Fillion, N., & Moir, R.H.C. (2018). Explanation and abstraction from a backward-error analytic perspective. European Journal for Philosophy of Science, 8, 735–759.

    Article  Google Scholar 

  • Frigg, R., Bradley, S., Du, H., & Smith, L.A. (2014). Laplace’s demon and the adventures of his apprentices. Philosophy of Science, 81, 31–59.

    Article  Google Scholar 

  • Frigg, R., Smith, L.A., & Stainforth, D.A. (2015). An assessment of the foundational assumptions in high-resolution climate projections: the case of ukcp09. Synthese, 192, 3979–4008.

    Article  Google Scholar 

  • Hairer, E., Lubich, C., & Wanner, G. (2002). Geometric numerical integration: Structure-preserving algorithms for ordinary differential equations. Springer Series in Computational Mathematics 31, first ed. 2002 second edition 2006.

  • Hairer, E., Nørsett, S., & Wanner, G. (1987). Solving ordinary differential equations i : Nonstiff problems. Springer Series in Computational Mathematics 8, second ed. 1993 third ed 2008.

  • Hairer, E., & Wanner, G. (1991). Solving ordinary differential equations ii : Stiff and differential-algebraic problems. Springer Series in Computational Mathematics 14, first ed. 1991 revised edition 2010.

  • Higham, N.J. (1996). Accuracy and stability of numerical algorithms. Philadelphia Siam: Society for Industrial and Applied Mathematics.

    Google Scholar 

  • Hollingsworth, A., Kållberg, P., Renner, V., & Burridge, D.M. (1983). An internal symmetric computational instability. Quaterly Journal of the Royal Meteorological Society, 109, 417–428.

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Kellert, S.H. (1993). In the wake of chaos. Chicago: University of Chicago Press.

    Book  Google Scholar 

  • Krivine, H., Lesne, A., & Treiner J. (2007). Discrete-time and continuous-time modelling: some bridges and gaps. Mathematical Structural in Comparative Science, 17, 1–16.

    Article  Google Scholar 

  • Lenhard, J. (2007). Computer simulation: The cooperation between experimenting and modeling. Philosophy of Science., 74(2), 176–194.

    Article  Google Scholar 

  • Lenhard, J. (2019). Calculated surprises: A philosophy of computer simulation. Oxford: Oxford Studies in Philosophy of Science.

    Book  Google Scholar 

  • Lenhard, J., & Küster, U. (2019). Reproducibility and the concept of numerical solution. Minds and Machines, 29(1), 19–36.

    Article  Google Scholar 

  • Mayo-Wilson, C. (2015). Structural chaos. Philosophy of Science, 82, 1236–1247.

    Article  Google Scholar 

  • Moir, R.H.C. (2010). Reconsidering backward error analysis for ordinary differential equations. Unpublished master’s thesis, Faculty of Science, Department of Applied Mathematics, The University of Western Ontario London, Ontario, Canada.

  • Moir, R.H.C. (2013). Structures in real theory application: A study in feasible epistemology.Unpublished doctoral dissertation, Faculty of Science, Department of Applied Mathematics, The University of Western Ontario London, Ontario, Canada. https://ir.lib.uwo.ca/etd/1578.

  • Moir, R.H.C. (2019). Effective validity: a generalized logic for stable approximate inference. In Fillion, N., Corless, R.M., & Kotsireas, IS (Eds.) Proceedings of 2015 and 2016 ACMES conferences, Fields institute communications, (Vol. 82 pp. 225–268). New York: Springer.

  • Nabergall, L., Navas, A., & Winsberg E. (2019). An antidote for hawkmoths: on the prevalence of structural chaos in non-linear modeling. European Journal for Philosophy of Science, 9, 21.

    Article  Google Scholar 

  • Orszag, S.A. (1971). On the elimination of aliasing in finite difference schemes by filtering high-wavenumber components. Journal of Atmospheric Science, 28, 1074.

    Article  Google Scholar 

  • Palais, R.S., & Palais, R.A. (2009). Differential equations, mechanics, and computation. Student Mathematical Library, 51. IAS/Park City Mathematical Subseries. American Mathematical Society, Providence RI; Institute for Advanced Study (IAS), Princeton, NJ.

  • Phillips, N.A. (1959). An example of non-linear computational instability. In Bolin, B. (Ed.) The atmosphere and the sea in motion (pp. 501–504). Oxford: Oxford University Press.

  • Platzman, G.W. (1961). An approximation to the product of discrete functions. Journal of Meteorology, 18, 31–37.

    Article  Google Scholar 

  • Smith, P. (1998). Explaining chaos. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Söderlind, G. (2006). Time-step selection algorithms: Adaptivity, control, and signal processing. Applied Numerical Mathematics, 56(3), 488–502.

    Article  Google Scholar 

  • Stern, A., & Desbrun, M. (2008). Discrete geometric mechanics for variational time integrators. In Discrete differential geometry : An applied introduction (chap. 15). Siggraph Course Note.

  • Süli, E., & Mayers, D. (2003). An introduction to numerical analysis. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Trefethen, L.N. (1994). Finite difference and spectral methods for ordinary and partial differential equations. Available at http://people.maths.ox.ac.uk/trefethen/pdetext.html.

  • Werndl, C. (2009). What are the new implications of chaos for unpredictability? British Journal for the Philosophy of Science., 60(1), 195–220.

    Article  Google Scholar 

  • Wilkinson, J.H. (1971). Modern error analysis. SIAM Review, 13 (4), 548–568.

    Article  Google Scholar 

  • Winsberg, E. (2018). Philosophy and climate science. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Winsberg, E., & Goodwin, W.M. (2016). The adventures of climate science in the sweet land of idle arguments. Studies in History and Philosophy of Modern Physics, 54, 9–17.

    Article  Google Scholar 

Download references

Funding

JJ is funded by the Swiss National Science Foundation project PP00P1_170460 “The Epistemology of Climate Change”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincent Ardourel.

Ethics declarations

- Disclosure of potential conflicts of interest: none.- Research involving human participants and/or animals: not applicable.- Informed consent: all the co-authors gave explicit consent to submit the paper.

Additional information

Publisher’s note

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

This article belongs to the Topical Collection: Dimensions of Applied Mathematics Guest Editors: Matthew W. Parker, Davide Rizza

Appendix: A

Appendix: A

This Appendix provides examples and mathematical details on some notions discussed in the paper, viz. stiff problems, zero-stability and absolute stability.

1.1 A.1 Stiff problems

Here is an one-dimensional example of stiff problem provided by Hairer and Wanner (1991):Footnote 15

$$ {\frac{dy}{dx}=-50(y-\cos{x})} $$
(7)

On the left of Fig. 5 are the solution curves of the equation. They show a smooth solution, close to \(y=\cos \limits {x}\), as well as a bunch of solutions that reach this smooth solution after a transient phase. However, some other numerical solutions do not behave the same way. On the right of Fig. 5 are two solutions with initial value y(0) = 1, and respectively step sizes h = 1.974/50 and 1.875/50. These solutions have no transient phase, and enable us to infer that, as soon as the step size is larger than 2/50, the numerical solution goes too far beyond the equilibrium and violent oscillations occur (Hairer and Wanner 1991, 2).

Fig. 5
figure 5

On the left, solution curves with implicit Euler solution. On the right, explicit Euler solution for y(0) = 0, h = 1.974/50 and 1.875/50. (Hairer and Wanner 1991, 2)

Specific numerical methods have been developed in order to solve stiff problems (see Appendix Appendix).

1.2 A.2 Zero-stability

In order to understand the importance of a numerical method being zero-stable, let us take the example of a dynamical system defined by the following partial differential equation:

$$ {\frac{\partial u}{\partial t}=\frac{\partial u}{\partial x}} $$
(8)

with \(u(x,0)=\cos \limits ^{2}x\) if \( \lvert x \rvert \leq \pi /2\), u(x, 0) = 0 otherwise.

Let us take the leap frog numerical method, viz. the discretization \(u^{n+1}_{j}=u^{n-1}_{j}+\lambda (u^{n}_{j+1}-u^{n}_{j-1})\), with the space step h = 0.04π and time step k = λh, where λ is the mesh ratio (see Trefethen 1994, 148). Figure 6 shows the two numerical solutions with respectively λ = 0.9 and λ = 1.1. As we can see, for λ < 1 there is a propagation of the wave as predicted by the theory. This is the regime where the numerical method is zero-stable. On the contrary, for λ > 1 numerical artefacts appear in the propagation of the wave. This is the regime where the numerical method is not zero-stable.

Fig. 6
figure 6

Stable and unstable leap frog approximations to \(\frac {\partial u}{\partial t}=\frac {\partial u}{\partial x}\). (Trefethen 1994, 149, Fig. 4.1.1)

As emphasized by Trefethen (1994), in the case where λ > 1: “The errors introduced at each step are not much bigger than before [i.e. for λ < 1] but they grow exponentially in subsequent time steps [...]. This rapid blow up of a sawtooth mode is typical of unstable finite difference formulas” (1994, chap 4, 148).

It is important to emphasize that the instability is not generated by round-off errors; it is generated by discretization errors which accumulate in the regime where the numerical method is not zero-stable. In both cases (λ < 1 and λ > 1) the numerical method is consistent. Nevertheless, the numerical solution to Eq. 8 would converge to the exact solution when h and k tend to zero in the regime of zero-stability, i.e. when λ < 1.Footnote 16 Thus, as we have just seen, in order to avoid numerical instability, it can be crucial to investigate whether a numerical method is zero-stable.

Zero-stability can be applied to numerical methods used for ordinary differential equations. In those cases, it concerns the behavior of the numerical method for a fixed time value t, as time step h tends to zero (see Hairer et al. 1987, 378, and Fasshauer 2007, chap. 4). This is a particularly important condition for multi-step methods (Süli and Mayers; Hairer et al. 1987, 398; Trefethen 1994, chap. 1, 40 2003; Hairer et al. 1987, 398; Trefethen 1994, chap. 1, 40, 332). Zero-stability also concerns the numerical methods used for partial differential equations. In this case, it also concerns the behavior of the numerical method for a fixed time value t, as time step k tends to zero, but also as spatial step h tends to zero.

Consistency is a property required for a numerical method. A numerical method is said to be consistent if the local errors introduced at each step tend to zero when the discretization steps tend to zero. However, consistency is a necessary but not sufficient condition for solution convergence, which means that the discrete solution tends to the exact continuous one when the time step tends to zero. Zero-stability is also required: if a numerical method is consistent and zero-stable, then it is convergent. In other words, being zero-stable allows a numerical method to be convergent.Footnote 17

1.3 A.3 Absolute stability and stiffness

In order to better understand what absolute stability is, and how it is connected with stiff problems, let us take the example of the initial value problem:

$$ \frac{dx}{dt}= -\mu x \ \ \ x(0)=a $$
(9)

The solution is an exponential function such as x(t) = aeμt with μ > 0.

The resolution of the initial value problem with the Euler forward method leads to the following discrete equation:

$$ x_{k+1}=x_{k}-h\mu x_{k}, \ \ \ x_{0}=a $$
(10)

The exact discrete solution is: xk = a(1 − μh)k. If \(h<\frac {2}{\mu }\), both continuous and discrete solutions tend to zero when t or k tends to infinity. Nevertheless, when \(h>\frac {2}{\mu }\), the discrete solution increases with k. This illustrates that the Euler method is not absolutely stable for all values of h. It is a conditionally stable numerical method.

Let us now use the Euler backward method in order to solve this same problem. In that case, the discrete equation becomes:

$$ x_{k+1}=x_{k}-h\mu x_{k+1}, \ \ \ x_{0}=a $$
(11)

The exact discrete solution is: xk = x0/(1 + μh)k. Here, both continuous and discrete equations tend to zero when t and k tend to infinity, whatever the value of h is. This illustrates that the Euler backward method is an unconditionally stable numerical method.

The Euler forward method is an explicit method: one just needs to know the value xk to directly infer the value xk+ 1 at the next time step by applying the discrete time evolution. On the contrary, the Euler backward method is implicit: one cannot directly derive the value of xk+ 1 with xk, instead one generally has to solve a non-linear equation with xk+ 1. Generally, explicit methods are not suitable to solve stiff problems – something which is sometimes part of the definition of stiffness itself! Thus, Hairer et al. (1987) assert that “stiff equations are problems for which explicit methods don’t work” (1987, 2). Implicit methods, on the contrary, are indeed generally suitable.Footnote 18

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ardourel, V., Jebeile, J. Numerical instability and dynamical systems. Euro Jnl Phil Sci 11, 49 (2021). https://doi.org/10.1007/s13194-021-00372-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13194-021-00372-7

Keywords

Navigation