Skip to main content
Log in

Optimal Monotonicity-Preserving Perturbations of a Given Runge–Kutta Method

  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

Perturbed Runge–Kutta methods (also referred to as downwind Runge–Kutta methods) can guarantee monotonicity preservation under larger step sizes relative to their traditional Runge–Kutta counterparts. In this paper we study the question of how to optimally perturb a given method in order to increase the radius of absolute monotonicity (a.m.). We prove that for methods with zero radius of a.m., it is always possible to give a perturbation with positive radius. We first study methods for linear problems and then methods for nonlinear problems. In each case, we prove upper bounds on the radius of a.m., and provide algorithms to compute optimal perturbations. We also provide optimal perturbations for many known methods.

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

Similar content being viewed by others

Notes

  1. Although the results in [24] are given in the context of contractivity, they are also relevant to the preservation of monotonicity. In [24, Thm. 5.1], quotients \(m_\tau [x,y]\) and one-sided Gateaux variations (\(m_+[ x,y], m_-[ x,y]\)), are used for \(x=u-{{\tilde{u}}}\) and \(y=f(u)-f({\tilde{u}})\) to obtain the contractivity property \(\Vert u(t)-\tilde{u}(t)\Vert \le \Vert u(t_0)-{\tilde{u}}(t_0)\Vert \) for \(t\ge t_0\). In the context of monotonicity, we simply take \(x=u\) and \(y=f(u)\) to obtain the monotonicity property \(\Vert u(t)\Vert \le \Vert u(t_0)\Vert \) for \(t\ge t_0\).

References

  1. Bogacki, P., Shampine, L.F.: An efficient Runge–Kutta (4, 5) pair. Comput. Math. Appl. 32(6), 15–28 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  2. Calvo, M., Montijano, J.I., Rández, L.: A new embedded pair of Runge–Kutta formulas of orders 5 and 6. Comput. Math. Appl. 20(1), 15–24 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  3. Conde, S., Gottlieb, S., Grant, Z.J., Shadid, J.N.: Implicit and implicit–explicit strong stability preserving Runge–Kutta methods with high linear order. J. Sci. Comput. 73(2), 667–690 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  4. Donat, R., Higueras, I., Martínez-Gavara, A.: On stability issues for IMEX schemes applied to hyperbolic equations with stiff reaction terms. Math. Comput. 80, 2097–2126 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  5. Dormand, J.R., Prince, P.J.: A family of embedded Runge–Kutta formulae. J. Comput. Appl. Math. 6(1), 19–26 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  6. Fehlberg, E.: Klassische Runge–Kutta–Formeln fünfter und siebenter Ordnung mit Schrittweiten–Kontrolle. Computing 4(2), 93–106 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  7. Ferracina, L., Spijker, M.N.: Stepsize restrictions for the total-variation-diminishing property in general Runge–Kutta methods. SIAM J. Numer. Anal. 42, 1073–1093 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  8. Gottlieb, S., Ketcheson, D.I., Shu, C.W.: Strong Stability Preserving Runge–Kutta and Multistep Time Discretizations. World Scientific Publishing Company, Singapore (2011)

    Book  MATH  Google Scholar 

  9. Gottlieb, S., Ruuth, S.J.: Optimal strong-stability-preserving time-stepping schemes with fast downwind spatial discretizations. J. Sci. Comput. 27, 289–303 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  10. Gottlieb, S., Shu, C.W.: Total variation diminishing Runge–Kutta schemes. Math. Comput. 67(221), 73–85 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  11. Hairer, E., Wanner, G.: Solving Ordinary Differential Equations II. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  12. Heun, K.: Neue methoden zur approximativen integration der differentialgleichungen einer unabhängigen veränderlichen. Z. Math. Phys. 45, 23–38 (1900)

    MATH  Google Scholar 

  13. Higueras, I.: Representations of Runge–Kutta methods and strong stability preserving methods. SIAM J. Numer. Anal. 43, 924–948 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  14. Higueras, I.: Strong stability for additive Runge–Kutta methods. SIAM J. Numer. Anal. 44(4), 1735–1758 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  15. Higueras, I.: Positivity properties for the classical fourth order Runge–Kutta methods. Monografías de la Real Academia de Ciencias de Zaragoza 33, 125–139 (2010)

    MathSciNet  Google Scholar 

  16. Higueras, I., Ketcheson, D.I.: Reproducibility repository for computations of optimal perturbations to Runge-Kutta methods, version 0.2 (2016). https://doi.org/10.5281/zenodo.1146916

  17. Horváth, Z.: On the positivity step size threshold of Runge–Kutta methods. Appl. Numer. Math. 53, 341–356 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  18. Hundsdorfer, W., Koren, B., van Loon, M., Verwer, J.C.: A positive finite-difference advection scheme. J. Comput. Phys. 117(1), 35–46 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  19. Ketcheson, D.I.: Highly efficient strong stability preserving Runge–Kutta Methods with low-storage implementations. SIAM J. Sci. Comput. 30, 2113–2136 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  20. Ketcheson, D.I.: Computation of optimal monotonicity preserving general linear methods. Math. Comput. 78, 1497–1513 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  21. Ketcheson, D.I.: High Order Strong Stability Preserving Time Integrators and Numerical Wave Propagation Methods for Hyperbolic PDEs. Doctoral thesis, University of Washington (2009)

  22. Ketcheson, D.I.: Step sizes for strong stability preservation with downwind-biased operators. SIAM J. Numer. Anal. 49(4), 1649–1660 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  23. Ketcheson, D.I.: Nodepy software version 0.6.1 (2015). Available from http://github.com/ketch/nodepy

  24. Kraaijevanger, J.F.B.M.: Contractivity of Runge–Kutta methods. BIT 31, 482–528 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  25. LeVeque, R.J., Yee, H.C.: A study of numerical methods for hyperbolic conservation laws with stiff source terms. J. Comput. Phys. 210, 187–210 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  26. Merson, R.H.: An operational method for the study of integration processes. In: Proceedings of the Symposium on Data Processing, pp. 1–25 (1957)

  27. Prince, P.J., Dormand, J.R.: High order embedded Runge–Kutta formulae. J. Comput. Appl. Math. 7(1), 67–75 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  28. Ruuth, S.J.: Global optimization of explicit strong-stability-preserving Runge–Kutta Methods. Math. Comput. 75, 183–207 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Ruuth, S.J., Spiteri, R.J.: High-order strong-stability-preserving Runge–Kutta methods with downwind-biased spatial discretizations. SIAM J. Numer. Anal. 42, 974–996 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  30. Shu, C.W., Osher, S.: Efficient implementation of essentially non-oscillatory shock-capturing schemes. J. Comput. Phys. 77(2), 439–471 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  31. Zhang, X., Shu, C.W.: On maximum-principle-satisfying high order schemes for scalar conservation laws. J. Comput. Phys. 229(9), 3091–3120 (2010)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Inmaculada Higueras.

Additional information

The Inmaculada Higueras was supported by Ministerio de Economía y Competividad, Spain, Projects MTM2014-53178-P and MTM2016-77735-C3-2-P. The David I. Ketcheson and Tihamér A. Kocsis were supported by KAUST Award No. FIC/2010/05-2000000231. The Tihamér A. Kocsis was also supported by TÁMOP-4.2.2.A-11/1/KONV-2012-0012: Basic research for the development of hybrid and electric vehicles, supported by the Hungarian Government and co-financed by the European Social Fund.

Appendix

Appendix

In this section we give additional details on SSP coefficients and optimal perturbations of second order 2-stage Runge–Kutta methods and the classical 4-stage fourth order Runge–Kutta method.

1.1 Second Order 2-Stage Methods

We consider the family of 2-stage second order methods (26). In Example 1 we studied perturbations that increase the SSP coefficient for the linear case. For nonlinear problems, in Example 3, Fig.  2 shows the values of \(R^{\mathrm{opt}}(K)\) for \(\alpha \in [-3,3]\).

In this section, for each \(\alpha \), we give the expressions for \(R^{\mathrm{opt}}(K)\) and we show optimal perturbations \(\widetilde{K}_{NL}\) such that \(R(K, \widetilde{K}_{NL})=R^{\mathrm{opt}}(K)\). It is important to point out the convenience of choosing \(\widetilde{K}_{NL}= \widetilde{K}_{L}\), where \(\widetilde{K}_{L}\) denotes the optimal perturbation for the linear case. In this case, we have not only \(R(K, \widetilde{K}_{L})=R^{\mathrm{opt}}(K)\) but also \({R}_{\mathrm{Lin}}(K, \widetilde{K}_{NL})=R_{\mathrm{Lin}}^{\mathrm{opt}}(K)\). The computations required to obtain the results in this section have been done with the symbolic computation program Mathematica.

If we denote by \(r=R^{\mathrm{opt}}(K)\), we have that

$$\begin{aligned} r= {\left\{ \begin{array}{ll} \displaystyle \frac{1}{|\alpha |}, &{} \qquad \displaystyle \hbox {if}\; \quad \alpha \in \left( -\infty , -\frac{1}{2} \left( 1+\sqrt{7}\right) \right] \bigcup \left[ \frac{1}{2} \left( -1+\sqrt{7}\right) , \infty \right) , \\ \displaystyle \frac{-1+\alpha +\sqrt{ 3 \alpha ^2-2 \alpha +1} }{|\alpha | }, &{} \qquad \displaystyle \hbox {if}\; \quad \alpha \in \left( -\frac{1}{2} \left( 1+\sqrt{7}\right) , 0\right) \bigcup \left( 0, \frac{1}{2} \left( -1+\sqrt{7}\right) \right) . \end{array}\right. }\nonumber \\ \end{aligned}$$
(72)

Next we give optimal perturbations \(\widetilde{K}_{NL}\).

For \(\alpha <0\), we obtain that it is not possible to obtain a perturbation of the form (26) with \({\tilde{b}}_2=0\) and \({\tilde{a}}_{21}=0\). Consequently, \(\widetilde{K}_{NL}\ne \widetilde{K}_{L}\) and we always have that \({R}_{\mathrm{Lin}}(K, \widetilde{K}_{NL})<R_{\mathrm{Lin}}^{\mathrm{opt}}(K)\). Optimal perturbations of the form (26) for different values of \(\alpha <0\) must satisfy the following conditions.

  • For \(-\frac{1}{2} \left( 1+\sqrt{7}\right) \le \alpha <0\), the coefficients \({\tilde{a}}_{21}, {\tilde{b}}_1\) and \({\tilde{b}}_2\) in \(\widetilde{K}_{NL}\) must satisfy

    $$\begin{aligned} -\,\alpha \le {\tilde{a}}_{21}\le \frac{1-r \, \alpha }{2 \, r}, \qquad {\tilde{b}}_1=-\frac{r\, {\tilde{a}}_{21}}{2\, \alpha } , \qquad {\tilde{b}}_2=-\frac{1}{2\, \alpha }, \end{aligned}$$

    where \(r=R^{\mathrm{opt}}(K)\).

  • For \(\alpha \le -\frac{1}{2} \left( 1+\sqrt{7}\right) \), we should have

    $$\begin{aligned} {\tilde{a}}_{21}=-\,\alpha ,\qquad -\frac{1}{2 \,\alpha }\le {\tilde{b}}_1\le \frac{-2 \,\alpha ^2-2 \, \alpha +1}{4\, \alpha }, \qquad -\,\frac{1}{2 \,\alpha }\le {\tilde{b}}_2\le \frac{2\, \alpha \,{\tilde{b}}_1-1}{4\, \alpha }. \end{aligned}$$

For \(\alpha >0\) we can find optimal perturbations with \({\tilde{b}}_2=0\) and \({\tilde{a}}_{21}=0\). Coefficient \({\tilde{b}}_1\) must satisfy the following conditions.

  • For \(0<\alpha \le \left( -\,1+\sqrt{7}\right) /2\), we have that

    $$\begin{aligned} \qquad {\tilde{b}}_1=\frac{\sqrt{3 \alpha ^2-2 \alpha +1}-\alpha }{2 \alpha } . \end{aligned}$$
    (73)

    Thus there is a unique \(\widetilde{K}_{NL}\) of the form (26). In this case, we have \(R(K)<R(K, \widetilde{K}_{NL})=R^{\mathrm{opt}}(K)\).

  • For \(\left( -1+\sqrt{7}\right) /2<\alpha < 1\), we also get \(R(K)<R^{\mathrm{opt}}(K)\), but in this case the optimal perturbation \(\widetilde{K}_{NL}\) is not unique. All the perturbations with \({\tilde{b}}_1\) satisfying

    $$\begin{aligned} \frac{1-\alpha }{\alpha } \le {\tilde{b}}_1\le \frac{2 \alpha ^2-2 \alpha +1}{4 \alpha } , \end{aligned}$$

    are optimal. In particular, we can take \(\widetilde{K}_{NL}=\widetilde{K}_{L}\). With this choice, \(R(K, \widetilde{K}_{L})=R^{\mathrm{opt}}(K)=1/\alpha \) and \({R}_{\mathrm{Lin}}(K, \widetilde{K}_{L})=R_{\mathrm{Lin}}^{\mathrm{opt}}(K)\approx 1.22\). Furthermore, \(\alpha =\left( -\,1+\sqrt{7}\right) /2\) provides the largest SSP coefficient within the family of 2-stage second order method (see Fig. 2).

  • For \(1\le \alpha \), we have \(R(K)=R^{\mathrm{opt}}(K)=1/\alpha \) and the optimal perturbation \(\widetilde{K}_{NL}\) is not unique. All the values

    $$\begin{aligned}0\le {\tilde{b}}_1\le \frac{2 \alpha ^2-2 \alpha +1}{4 \alpha } \end{aligned}$$

    give optimal perturbations. We can take \(\widetilde{K}_{NL}=0\), but in this case \({R}_{\mathrm{Lin}}(K, 0)< R_{\mathrm{Lin}}^{\mathrm{opt}}(K)\). A better choice is \(\widetilde{K}_{NL}=\widetilde{K}_{L}\). Observe that, for \(\alpha =1\), we get the optimal SSP coefficient \(R(K)=1\) that cannot be increased by perturbations.

Next, we consider some concrete values of \(\alpha \) to show the the expressions of the perturbations. For each value, we give the Butcher tableau of the perturbation and matrices \(\alpha ^{\mathrm{up}}\) and \(\alpha ^{\mathrm{down}}\) in (36).

  • For \(\alpha =1/2\) we get method RK2a in [18] with \(R(K)=0\). With perturbation

    $$\begin{aligned} \widetilde{K}&= \begin{pmatrix} 0 &{}\quad 0 &{} \quad 0 \\ 0 &{}\quad 0 &{}\quad 0\\ {\tilde{b}}_1 &{}\quad 0 &{}\quad 0 \end{pmatrix}, \quad \alpha ^{\mathrm{up}}= \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ {\tilde{b}}_1 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 2 {\tilde{b}}_1 &{}\quad 0 \end{pmatrix}, \quad \alpha ^{\mathrm{down}}= \begin{pmatrix} 0 &{} \quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0\\ 1-2{\tilde{b}}_1 &{}\quad 0 &{}\quad 0 \end{pmatrix},\nonumber \\ \gamma&= \begin{pmatrix} 1\\ 1-{\tilde{b}}_1\\ 0 \end{pmatrix}, \end{aligned}$$
    (74)

    where \({\tilde{b}}_1=\frac{1}{2} \left( \sqrt{3}-1\right) \), we get \(R(K,\widetilde{K}) ={R}_{\mathrm{Lin}}(K, \widetilde{K})= \sqrt{3}-1\).

  • For \(\alpha =2/3\), we have a nontrivial SSP coefficient \(R^{\mathrm{opt}}(K)=1/2\), but we can increase this value to \(R(K,\widetilde{K}_1)=R^{\mathrm{opt}}(K)=1\) with perturbation

    $$\begin{aligned} \widetilde{K}_1&= \begin{pmatrix} 0 &{} \quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0\\ 1/4 &{}\quad 0 &{}\quad 0 \end{pmatrix}, \quad \alpha ^{\mathrm{up}}= \begin{pmatrix} 0 &{} \quad 0 &{}\quad 0 \\ 2/3 &{}\quad 0 &{}\quad 0\\ 0 &{}\quad 3/4 &{}\quad 0 \end{pmatrix} , \quad \alpha ^{\mathrm{down}}= \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0\\ 1/4 &{}\quad 0 &{}\quad 0 \end{pmatrix} ,\\ \gamma&= \begin{pmatrix} 1\\ 1/3\\ 0 \end{pmatrix}. \end{aligned}$$

    For this perturbation, \(R(\phi _K)={R}_{\mathrm{Lin}}(K, \widetilde{K}_1)=1\). We can take \(\gamma =(1,0,0)^t\) by modifying the first column of \(\alpha ^{\mathrm{up}}\) and \(\alpha ^{\mathrm{down}}\) according to (50),

    $$\begin{aligned} \widetilde{K}_2&= \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ 1/6 &{}\quad 0 &{}\quad 0\\ 3/8 &{}\quad 0 &{}\quad 0 \end{pmatrix}, \quad \alpha ^{\mathrm{up}}= \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 5/6 &{} \quad 0 &{} \quad 0\\ 0 &{} \quad 3/4 &{} \quad 0 \end{pmatrix} , \quad \alpha ^{\mathrm{down}}= \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 1/6 &{} \quad 0 &{} \quad 0\\ 1/4 &{} \quad 0 &{} \quad 0 \end{pmatrix} ,\\ \gamma&= \begin{pmatrix} 1\\ 0\\ 0 \end{pmatrix}. \end{aligned}$$
  • As it has been pointed out above, the largest value in the \(\alpha \)-family of 2-stage second order schemes is \(R^{\mathrm{opt}}(K)=(1+\sqrt{7})/3\) and it is obtained for \(\alpha =(\sqrt{7}-1)/2\). The perturbation is of the form (26) with \({\tilde{b}}_1= \left( \sqrt{7}-2\right) /2\), and

    $$\begin{aligned} \alpha ^{\mathrm{up}}= \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 1 &{} \quad 0 &{} \quad 0\\ 0 &{} \quad \frac{1}{9} \left( 4+\sqrt{7}\right) &{}\quad 0 \end{pmatrix} , \quad \alpha ^{\mathrm{down}}= \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0\\ \frac{1}{9} \left( 5-\sqrt{7}\right) &{} \quad 0 &{} \quad 0 \end{pmatrix} , \quad \gamma _r = \begin{pmatrix} 1\\ 0 \\ 0 \end{pmatrix} \end{aligned}$$

    This is the perturbation obtained in [9, Table V] by numerical search in the class of perturbations considered in [9].

1.2 Classical Fourth Order 4-Stage Method

For nonlinear problems, applying the analysis above, we find that the optimal perturbation of the classical method has SSP coefficient given by the real root of \(x^3+2 x^2+4 x-4=0\), which is approximately \(R^{\mathrm{opt}}(K)\approx 0.685016\). The corresponding perturbation is not unique. For instance, we can take \(\gamma _r=(1,0,0,0,0)\), and all entries of \(\alpha ^{\mathrm{down}}_r\) equal to zero except

$$\begin{aligned} \left( \alpha ^{\mathrm{down}}_r\right) _{31}&= \frac{r^2}{4}&\left( \alpha ^{\mathrm{down}}_r\right) _{42}&= \frac{r^2}{2}, \end{aligned}$$
(75)

where \(r=R^{\mathrm{opt}}(K)\). However, there exist other optimal perturbations with additionally \((\alpha ^{\mathrm{down}}_r)_{42}=\epsilon \) where \(0\le \epsilon \le 0.782\).

We remark that nearly-optimal perturbations for this method are given in [30, p. 448] and [15]. Interestingly, these different perturbed methods have different values of \({R}_{\mathrm{Lin}}(K, \widetilde{K})\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Higueras, I., Ketcheson, D.I. & Kocsis, T.A. Optimal Monotonicity-Preserving Perturbations of a Given Runge–Kutta Method. J Sci Comput 76, 1337–1369 (2018). https://doi.org/10.1007/s10915-018-0664-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-018-0664-3

Keywords

Mathematics Subject Classification

Navigation