Skip to main content
Log in

EPIRK-W and EPIRK-K Time Discretization Methods

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

Abstract

Exponential integrators are special time discretization methods where the traditional linear system solves used by implicit schemes are replaced with computing the action of matrix exponential-like functions on a vector. A very general formulation of exponential integrators is offered by the Exponential Propagation Iterative methods of Runge–Kutta type (EPIRK) family of schemes. The use of Jacobian approximations is an important strategy to drastically reduce the overall computational costs of implicit schemes while maintaining the quality of their solutions. This paper extends the EPIRK class to allow the use of inexact Jacobians as arguments of the matrix exponential-like functions. Specifically, we develop two new families of methods: EPIRK-W integrators that can accommodate any approximation of the Jacobian, and EPIRK-K integrators that rely on a specific Krylov-subspace projection of the exact Jacobian. Classical order conditions theories are constructed for these families. Practical EPIRK-W methods of order three and EPIRK-K methods of order four are developed. Numerical experiments indicate that the methods proposed herein are computationally favorable when compared to a representative state-of-the-art exponential integrator, and a Rosenbrock–Krylov integrator.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  1. Al-Mohy, A.H., Higham, N.J.: Computing the action of the matrix exponential, with an application to exponential integrators. SIAM J. Sci. Comput. 33(2), 488–511 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  2. Allen, S.M., Cahn, J.W.: A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening. Acta Metall. 27(6), 1085–1095 (1979). https://doi.org/10.1016/0001-6160(79)90196-2

    Article  Google Scholar 

  3. Berland, H., Owren, B., Skaflestad, B.: B-series and order conditions for exponential integrators. SIAM J. Numer. Anal. 43(4), 1715–1727 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  4. Berland, H., Skaflestad, B., Wright, W.M.: Expint—a matlab package for exponential integrators. ACM Trans. Math. Softw. (TOMS) 33(1), 4 (2007)

    Article  Google Scholar 

  5. Butcher, J.: Numerical Methods for Ordinary Differential Equations, 2nd edn. Wiley, New York (2008). https://doi.org/10.1002/9780470753767

    Book  MATH  Google Scholar 

  6. Butcher, J.: Trees, B-series and exponential integrators. IMA J. Numer. Anal. 30, 131–140 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  7. Caliari, M., Ostermann, A.: Implementation of exponential rosenbrock-type integrators. Appl. Numer. Math. 59(3–4), 568–581 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  8. D’Augustine, A.F.: MATLODE: A MATLAB ODE solver and sensitivity analysis toolbox. Masters Thesis, Virginia Tech (2018)

  9. Hairer, E., Norsett, S., Wanner, G.: Solving Ordinary Differential Equations I: Nonstiff Problems. Springer Series in Computational Mathematics, vol. 8. Springer, Berlin (1993)

    MATH  Google Scholar 

  10. Hairer, E., Wanner, G.: Solving Ordinary Differential Equations II: Stiff and Differential-Algebraic Problems. Springer Series in Computational Mathematics, vol. 14, 2nd edn. Springer, Berlin (1996)

    Book  MATH  Google Scholar 

  11. Heineken, W., Warnecke, G.: Partitioning methods for reaction–diffusion problems. Appl. Numer. Math. 56(7), 981–1000 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  12. Hochbruck, M., Lubich, C.: On Krylov subspace approximations to the matrix exponential operator. SIAM J. Numer. Anal. 34(5), 1911–1925 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  13. Hochbruck, M., Lubich, C., Selhofer, H.: Exponential integrators for large systems of differential equations. SIAM J. Sci. Comput. 19(5), 1552–1574 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  14. Hochbruck, M., Ostermann, A.: Explicit exponential Runge–Kutta methods for semilinear parabolic problems. SIAM J. Numer. Anal. 43(3), 1069–1090 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hochbruck, M., Ostermann, A.: Exponential integrators. Acta Numer. 19, 209–286 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  16. Loffeld, J., Tokman, M.: Comparative performance of exponential, implicit, and explicit integrators for stiff systems of odes. J. Comput. Appl. Math. 241, 45–67 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  17. Lorenz, E.N.: Predictability—a problem partly solved. In: Palmer, T., Hagedorn, R. (eds.) Predictability of Weather and Climate, pp. 40–58. Cambridge University Press (CUP), Cambridge (1996). https://doi.org/10.1017/cbo9780511617652.004

    Google Scholar 

  18. Lu, Y.Y.: Computing a matrix function for exponential integrators. J. Comput. Appl. Math. 161(1), 203–216 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  19. Minchev, B.V., Wright, W.: A review of exponential integrators for first order semi-linear problems. Technical report (2005)

  20. Niesen, J., Wright, W.M.: Algorithm 919: a krylov subspace algorithm for evaluating the \(\varphi \)-functions appearing in exponential integrators. ACM Trans. Math. Softw. 38(3), 22:1–22:19 (2012). https://doi.org/10.1145/2168773.2168781

    Article  MathSciNet  MATH  Google Scholar 

  21. Rainwater, G., Tokman, M.: A new approach to constructing efficient stiffly accurate Epirk methods. J. Comput. Phys. 323, 283–309 (2016). https://doi.org/10.1016/j.jcp.2016.07.026

    Article  MathSciNet  Google Scholar 

  22. Roberts, S., Popov, A., Sandu, A.: ODE Test Problems. https://sibiu.cs.vt.edu/ode-test-problems/index.html. Accessed 16 Dec 2017

  23. Saad, Y.: Iterative Methods for Sparse Linear Systems. SIAM, Philadelphia (2003)

    Book  MATH  Google Scholar 

  24. Schulze, J.C., Schmid, P.J., Sesterhenn, J.L.: Exponential time integration using Krylov subspaces. Int. J. Numer. Meth. Fluids 60(6), 561–609 (2008)

    MathSciNet  MATH  Google Scholar 

  25. Sidje, R.B.: Expokit: a software package for computing matrix exponentials. ACM Trans. Math. Softw. 24(1), 130–156 (1998). https://doi.org/10.1145/285861.285868

    Article  MATH  Google Scholar 

  26. Steihaug, T., Wolfbrandt, A.: An attempt to avoid exact Jacobian and nonlinear equations in the numerical solution of stiff differential equations. Math. Comput. 33(146), 521–521 (1979). https://doi.org/10.1090/s0025-5718-1979-0521273-8

    Article  MathSciNet  MATH  Google Scholar 

  27. Tokman, M.: Efficient integration of large stiff systems of ODEs with exponential propagation iterative (EPI) methods. J. Comput. Phys. 213(2), 748–776 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  28. Tokman, M.: A new class of exponential propagation iterative methods of Runge–Kutta type (EPIRK). J. Comput. Phys. 230, 8762–8778 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  29. Tokman, M., Loffeld, J., Tranquilli, P.: New adaptive exponential propagation iterative methods of Runge–Kutta type. SIAM J. Sci. Comput. 34(5), A2650–A2669 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  30. Tranquilli, P., Glandon, S.R., Sarshar, A., Sandu, A.: Analytical Jacobian-vector products for the matrix-free time integration of partial differential equations. J. Comput. Appl. Math. (2016). https://doi.org/10.1016/j.cam.2016.05.002

    MATH  Google Scholar 

  31. Tranquilli, P., Sandu, A.: Exponential–Krylov methods for ordinary differential equations. J. Comput. Phys. 278, 31–46 (2014). https://doi.org/10.1016/j.jcp.2014.08.013

    Article  MathSciNet  MATH  Google Scholar 

  32. Tranquilli, P., Sandu, A.: Rosenbrock–Krylov methods for large systems of differential equations. SIAM J. Sci. Comput. 36(3), A1313–A1338 (2014). https://doi.org/10.1137/130923336

    Article  MathSciNet  MATH  Google Scholar 

  33. Trefethen, L.N.: Evaluating matrix functions for exponential integrators via Carathéodory–Fejér approximation and contour integrals. Electron. Trans. Numer. Anal. 29, 1–18 (2007)

    MathSciNet  MATH  Google Scholar 

  34. Verwer, J.G., Spee, E.J., Blom, J.G., Hundsdorfer, W.: A second-order rosenbrock method applied to photochemical dispersion problems. SIAM J. Sci. Comput. 20(4), 1456–1480 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  35. van der Vorst, H.A.: Iterative Krylov Methods for Large Linear Systems. Cambridge University Press, Cambridge (2003). https://doi.org/10.1017/cbo9780511615115

    Book  MATH  Google Scholar 

Download references

Acknowledgements

This work has been supported in part by NSF through Awards NSF DMS-1419003, NSF DMS-1419105, NSF CCF-1613905, by AFOSR through Award AFOSR FA9550-12-1-0293-DEF, and by the Computational Science Laboratory at Virginia Tech.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mahesh Narayanamurthi.

Appendices

A Derivation of K-Methods

The epirkk method is derived from epirkw method, where a specific Krylov-subspace approximation of the Jacobian is used instead of an arbitrary approximate Jacobian as admitted by the W-method. We start the derivation by first stating the general form of the epirkw method:

$$\begin{aligned} \mathbf {Y}_i&= \mathbf {y}_{n} + a_{i,1}\, \varvec{\psi }_{i,1}(g_{i,1}\,\,h\,\mathbf {A}_n)\, h\mathbf {f}(\mathbf {y}_{n}) + \displaystyle \sum _{j = 2}^{i} a_{i,j}\, \varvec{\psi }_{i,j}(g_{i,j}\,h\,\mathbf {A}_{n})\, h\Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_{n}),\nonumber \\&\quad i = 1, \ldots , s - 1, \nonumber \\ \mathbf {y}_{n+1}&= \mathbf {y}_{n}\, + b_{1}\, \varvec{\psi }_{s,1}(g_{s,1}\,h\,\mathbf {A}_{n})\, h\mathbf {f}(\mathbf {y}_{n}) + \displaystyle \sum _{j = 2}^{s} b_{j}\, \varvec{\psi }_{s,j}(g_{s,j}\,h\,\mathbf {A}_{n})\, h\Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_{n}). \end{aligned}$$

In the above equation, the \(\varvec{\psi }\) function is as defined in Eq. (8) and the following simplifying assumption is made about it:

$$\begin{aligned} \varvec{\psi }_{i,j}(z)= \varvec{\psi }_{j}(z) = \displaystyle \sum _{k=1}^{j} p_{j,k}\, \varvec{\varphi }_k(z), \end{aligned}$$

where \(\varvec{\varphi }\) function is as defined in Eqs. (5) and (6). Additionally, the remainder function (\(\mathbf {r}(\mathbf {y})\)) and the forward difference operator (\(\Delta ^{(j)}\mathbf {r}(\mathbf {Y}_i)\)) are defined accordingly below:

$$\begin{aligned} \mathbf {r}(\mathbf {y})= & {} \mathbf {f}(\mathbf {y}) - \mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n} \, (\mathbf {y} - \mathbf {y}_{n}),\nonumber \\ \Delta ^{(j)}\mathbf {r}(\mathbf {Y}_i)= & {} \Delta ^{(j - 1)}\mathbf {r}(\mathbf {Y}_{i+1}) - \Delta ^{(j - 1)}\mathbf {r}(\mathbf {Y}_{i}),\nonumber \\ \Delta ^{(1)}\mathbf {r}(\mathbf {Y}_i)= & {} \mathbf {r}(\mathbf {Y}_{i+1}) - \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$
(33)

The K-method uses a specific Krylov-subspace based approximation of the Jacobian (\(\mathbf {J}_{n}\)). An M-dimensional Krylov-subspace is built as,

$$\begin{aligned} \mathcal {K}_M = \text {span}\{\mathbf {f}_n, \mathbf {J}_{n}\mathbf {f}_n, \mathbf {J}_{n}^2\mathbf {f}_n, \ldots , \mathbf {J}_{n}^{M-1}\mathbf {f}_n\}, \end{aligned}$$
(34)

whose basis is the orthonormal matrix \(\mathbf {V}\) and \(\mathbf {H}\) is the upper-Hessenberg matrix obtained from Arnoldi iteration defined as

$$\begin{aligned} \mathbf {H}= \mathbf {V}^{T} \mathbf {J}_{n} \,\mathbf {V}. \end{aligned}$$
(35)

The corresponding Krylov-subspace based approximation of the Jacobian is built as

$$\begin{aligned} \mathbf {A}_{n} = \mathbf {V}\,\mathbf {H}\,\mathbf {V}^{T} = \mathbf {V}\mathbf {V}^{T} \mathbf {J}_{n} \,\mathbf {V}\mathbf {V}^{T}. \end{aligned}$$
(36)

The use of Krylov-subspace based approximation of the Jacobian reduces the \(\varvec{\varphi }\) and \(\varvec{\psi }\) function in accordance with Lemmas 2 and 3 to the following

$$\begin{aligned} \varvec{\varphi }_k(h\,\gamma \,\mathbf {A}_{n})= & {} \frac{1}{k!} \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) + \mathbf {V}\varvec{\varphi }_k(h\,\gamma \,\mathbf {H}) \mathbf {V}^{T}, \end{aligned}$$
(37)
$$\begin{aligned} \varvec{\psi }_{j}(h\gamma \mathbf {A}_{n})= & {} \widetilde{p}_j \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) + \mathbf {V}\,\varvec{\psi }_{j}(h\gamma \mathbf {H}) \mathbf {V}^{T}, \end{aligned}$$
(38)

where \(\widetilde{p}_j\) is defined as

$$\begin{aligned} \widetilde{p}_j = \displaystyle \sum _{k=1}^{j}\frac{p_{j,k}}{k!}. \end{aligned}$$
(39)

In order to derive the reduced stage formulation of the epirkk method we need to resolve the vectors in the formulation into components in the Krylov-subspace and orthogonal to it. Repeating the splittings from the main text:

  • Splitting the internal stage vectors noting that \(\mathbf {Y}_0 \equiv \mathbf {y}_n\):

    $$\begin{aligned} \mathbf {Y}_i = \mathbf {V}\varvec{\lambda }_i + \mathbf {Y}_i^{\bot } \qquad \text {where} \quad \mathbf {V}^{T} \mathbf {Y}_i = \varvec{\lambda }_i, \quad \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \,\mathbf {Y}_i = \mathbf {Y}_i^{\bot }. \end{aligned}$$
    (40)
  • Splitting the right-hand side function evaluated at internal stage vectors while noting that \(\mathbf {f}_0 \equiv \mathbf {f}(\mathbf {y}_n)\):

    $$\begin{aligned} \mathbf {f}_i := \mathbf {f}(\mathbf {Y}_i) = \mathbf {V}\varvec{\eta }_i + \mathbf {f}_i^{\bot } \qquad \text {where} \quad \mathbf {V}^{T} \mathbf {f}_i = \varvec{\eta }_i, \quad \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \mathbf {f}_i = \mathbf {f}_i^{\bot }. \end{aligned}$$
    (41)
  • Splitting the non-linear Taylor remainder of the right-hand side functions:

    $$\begin{aligned}&\mathbf {r}(\mathbf {Y}_i) = \mathbf {f}(\mathbf {Y}_i) - \mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n}\, (\mathbf {Y}_i - \mathbf {y}_{n}) = \mathbf {f}_i - \mathbf {f}_0 - \mathbf {V}\, \mathbf {H}\, \mathbf {V}^{T}\, (\mathbf {Y}_i - \mathbf {y}_{n}), \nonumber \\&\quad \text {where} \quad \mathbf {V}^{T}\, \mathbf {r}(\mathbf {Y}_i) = \varvec{\eta }_i - \varvec{\eta }_0 - \mathbf {H}\, (\varvec{\lambda }_i - \varvec{\lambda }_0), \nonumber \\&\quad \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \, \mathbf {r}(\mathbf {Y}_i) = \mathbf {f}_i^{\bot } - \mathbf {f}_0^{\bot }. \end{aligned}$$
    (42)
  • Splitting the forward differences of the non-linear remainder terms:

    $$\begin{aligned}&\mathbf {\widetilde{r}}_{(j-1)}:=\Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_{n}) = \mathbf {V}\, \mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }, \nonumber \\&\quad \text {where} \quad \mathbf {V}^{T}\,\mathbf {\widetilde{r}}_{(j-1)} = \mathbf {d}_{(j-1)}, \nonumber \\&\quad \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \, \mathbf {\widetilde{r}}_{(j-1)} = \mathbf {\widetilde{r}}_{(j-1)}^{\bot }. \end{aligned}$$
    (43)

Using these each internal stage of the epirkk method can be expressed as below:

$$\begin{aligned} \mathbf {V}\varvec{\lambda }_i + \mathbf {Y}_i^{\bot }= & {} \mathbf {y}_{n}\, + h\, a_{i,1} \bigg (\widetilde{p_1} \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) + \mathbf {V}\,\varvec{\psi }_{1}(h\,g_{i,1}\, \mathbf {H}) \mathbf {V}^{T}\bigg ) \bigg (\mathbf {V}\varvec{\eta }_0 + \mathbf {f}_0^{\bot }\bigg ) \nonumber \\&+\, \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \bigg (\widetilde{p}_j \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) + \mathbf {V}\,\varvec{\psi }_{j}(h\,g_{i,j}\, \mathbf {H}) \mathbf {V}^{T}\bigg ) \bigg (\mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }\bigg ) \nonumber \\= & {} \mathbf {y}_{n}\, + h\, a_{i,1} \bigg (\widetilde{p_1} \, \mathbf {f}_0^{\bot } + \mathbf {V}\,\varvec{\psi }_{1}(h\,g_{i,1}\, \mathbf {H}) \varvec{\eta }_0\bigg ) \nonumber \\&+\, \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \bigg (\widetilde{p}_j \, \mathbf {\widetilde{r}}_{(j-1)}^{\bot } + \mathbf {V}\,\varvec{\psi }_{j}(h\,g_{i,j}\, \mathbf {H}) \, \mathbf {d}_{(j-1)}\bigg ). \end{aligned}$$
(44)

The reduced stage formulation of the epirkk method is obtained by multiplying the above equation by \(\mathbf {V}^{T}\) from the left

$$\begin{aligned} \varvec{\lambda }_i= & {} \mathbf {V}^{T} \mathbf {y}_{n}\, + h \, a_{i,1} \varvec{\psi }_{1}(h\,g_{i,1}\,\mathbf {H}) \varvec{\eta }_0 + \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \varvec{\psi }_{j}(h\,g_{i,j}\, \mathbf {H}) \, \mathbf {d}_{(j-1)} \nonumber \\= & {} \varvec{\lambda }_0 + h \, a_{i,1} \varvec{\psi }_{1}(h\,g_{i,1}\,\mathbf {H}) \varvec{\eta }_0 + \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \varvec{\psi }_{j}(h\,g_{i,j}\, \mathbf {H}) \, \mathbf {d}_{(j-1)}. \end{aligned}$$
(45)

And the full stage vector can be recovered by first computing the reduced stage vector and adding the orthogonal piece \(\mathbf {Y}_i^{\bot }\) obtained when multiplying Eq. (44) by \(\left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \)

$$\begin{aligned} \mathbf {Y}_i^{\bot }= & {} \left( \mathbf {I} - \mathbf {V}\mathbf {V}^{T}\right) \mathbf {y}_{n}\, + h \, a_{i,1} \, \widetilde{p_1} \, \mathbf {f}_0^{\bot } + \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \widetilde{p}_j \, \mathbf {\widetilde{r}}_{(j-1)}^{\bot } \nonumber \\= & {} (\mathbf {y}_{n}\, - \mathbf {V}\varvec{\varvec{\lambda }}_0) + h \, a_{i,1} \, \widetilde{p_1} \, \mathbf {f}_0^{\bot } + \displaystyle \sum _{j = 2}^{i} h\, a_{i,j}\, \widetilde{p}_j \, \mathbf {\widetilde{r}}_{(j-1)}^{\bot }. \end{aligned}$$
(46)

The final stage can also be written in the above form with multipliers \(b_i\) in place of \(a_{ij}\). Notice that the expensive computations are performed in the reduced space, i.e. the \(\varvec{\psi }\) function is computed in the reduced space instead of the full space offering potential computational savings. In the above equations, the quantities \(\mathbf {d}_{(j-1)}\) and \(\mathbf {\widetilde{r}}_{(j-1)}^{\bot }\) can be shown to be

$$\begin{aligned} \mathbf {d}_{(j-1)}= & {} \displaystyle \sum _{k=0}^{j-1} \bigg ((-1)^k {j-1 \atopwithdelims ()k} \varvec{\eta }_{j-1-k} - \mathbf {H}\bigg ((-1)^k {j-1 \atopwithdelims ()k} \varvec{\varvec{\lambda }}_{j-1 -k}\bigg )\bigg ), \end{aligned}$$
(47)
$$\begin{aligned} \mathbf {\widetilde{r}}_{(j-1)}^{\bot }= & {} \displaystyle \sum _{k=0}^{j-1} \bigg ((-1)^k {j-1 \atopwithdelims ()k} \mathbf {f}_{j-1-k}^{\bot }\bigg ), \end{aligned}$$
(48)

as is done in the following “Appendix”.

B Proofs

In order to prove Eqs. (23a) and (23b), we start with the definition of the remainder function and forward difference.

$$\begin{aligned} \mathbf {r}(\mathbf {y}) = \mathbf {f}(\mathbf {y}) - \mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n} (\mathbf {y} - \mathbf {y}_{n}), \end{aligned}$$
(49)
$$\begin{aligned} \Delta ^{(j)}\mathbf {r}(\mathbf {Y}_i)= & {} \Delta ^{(j - 1)}\mathbf {r}(\mathbf {Y}_{i+1}) - \Delta ^{(j - 1)}\mathbf {r}(\mathbf {Y}_{i}), \end{aligned}$$
(50a)
$$\begin{aligned} \Delta ^{(1)}\mathbf {r}(\mathbf {Y}_i)= & {} \mathbf {r}(\mathbf {Y}_{i+1}) - \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$
(50b)

Lemma 4

\(\Delta ^{(j)}\mathbf {r}(\mathbf {Y}_i) = \displaystyle \sum _{k=0}^{j} (-1)^k {j \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{i+j-k})\).

Proof

In order to prove the lemma, we resort to mathematical induction. Base case \(j = 1\),

$$\begin{aligned} \Delta ^{(1)}\mathbf {r}(\mathbf {Y}_i)= & {} \displaystyle \sum _{k=0}^{1} (-1)^k {1 \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{i+1-k})\\= & {} \mathbf {r}(\mathbf {Y}_{i+1}) - \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$

The base case is true by definition. We now assume that the proposition holds true for all j up to \(k-1\). We have,

$$\begin{aligned} \Delta ^{(k-1)}\mathbf {r}(\mathbf {Y}_i)= & {} \displaystyle \sum _{l=0}^{k-1} (-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-1-l}), \end{aligned}$$
(51)
$$\begin{aligned} \Delta ^{(k-1)}\mathbf {r}(\mathbf {Y}_{i+1})= & {} \displaystyle \sum _{l=0}^{k-1} (-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}). \end{aligned}$$
(52)

Then for \(j = k\),

$$\begin{aligned} \Delta ^{(k)}\mathbf {r}(\mathbf {Y}_i)= & {} \Delta ^{(k - 1)}\mathbf {r}(\mathbf {Y}_{i+1}) - \Delta ^{(k - 1)}\mathbf {r}(\mathbf {Y}_{i}) \\= & {} \displaystyle \sum _{l=0}^{k-1}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) - \displaystyle \sum _{l=0}^{k-1}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k -1 -l})\\= & {} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) \\&-\, \displaystyle \sum _{l=0}^{k-2}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k -1 -l}) + (-1)^k \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$

We perform a change of variable for the second summation, \(m = l + 1 \; \implies l = (m - 1)\),

$$\begin{aligned} \Delta ^{(k)}\mathbf {r}(\mathbf {Y}_i)= & {} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) \\&-\, \displaystyle \sum _{m=1}^{k-1}(-1)^{m-1} {k-1 \atopwithdelims ()m - 1} \mathbf {r}(\mathbf {Y}_{i+k -m}) + (-1)^k \mathbf {r}(\mathbf {Y}_i) \\= & {} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l {k-1 \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) \\&+\, \displaystyle \sum _{m=1}^{k-1}(-1)^{m} {k-1 \atopwithdelims ()m - 1} \mathbf {r}(\mathbf {Y}_{i+k -m}) + (-1)^k \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$

The summations run between the same start and end indices, and can be collapsed.

$$\begin{aligned} \Delta ^{(k)}\mathbf {r}(\mathbf {Y}_i)= & {} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l \bigg ({k-1 \atopwithdelims ()l} + {k-1 \atopwithdelims ()l - 1}\bigg ) \mathbf {r}(\mathbf {Y}_{i+k-l}) + (-1)^k \mathbf {r}(\mathbf {Y}_i). \end{aligned}$$

We use the identity,

$$\begin{aligned} {k-1 \atopwithdelims ()l} + {k-1 \atopwithdelims ()l - 1} = {k \atopwithdelims ()l}, \nonumber \end{aligned}$$

and arrive at the desired result,

$$\begin{aligned} \Delta ^{(k)}\mathbf {r}(\mathbf {Y}_i)= & {} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l {k \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) + (-1)^k \mathbf {r}(\mathbf {Y}_i) \nonumber \\= & {} (-1)^0 {k \atopwithdelims ()0} \mathbf {r}(\mathbf {Y}_{i+k}) + \displaystyle \sum _{l=1}^{k-1}(-1)^l {k \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}) + (-1)^k {k \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_i) \nonumber \\= & {} \displaystyle \sum _{l=0}^{k} (-1)^l {k \atopwithdelims ()l} \mathbf {r}(\mathbf {Y}_{i+k-l}). \end{aligned}$$
(53)

\(\square \)

Lemma 5

Given

$$\begin{aligned} \Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_{n}) = \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }, \end{aligned}$$
(54)

we need to prove

$$\begin{aligned} \mathbf {d}_{(j-1)}= & {} \displaystyle \sum _{k=0}^{j-1} \bigg ((-1)^k {j-1 \atopwithdelims ()k} \varvec{\eta }_{j-1-k} - \mathbf {H}\bigg ((-1)^k {j-1 \atopwithdelims ()k} \varvec{\varvec{\lambda }}_{j-1 -k}\bigg )\bigg ),\end{aligned}$$
(55)
$$\begin{aligned} \mathbf {\widetilde{r}}_{(j-1)}^{\bot }= & {} \displaystyle \sum _{k=0}^{j-1} \bigg ((-1)^k {j-1 \atopwithdelims ()k} \mathbf {f}_{j-1-k}^{\bot }\bigg ). \end{aligned}$$
(56)

Proof

We start with Lemma 4 where we have proven that

$$\begin{aligned} \Delta ^{(j)}\mathbf {r}(\mathbf {Y}_i) = \displaystyle \sum _{k=0}^{j} (-1)^k {j \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{i+j-k}). \end{aligned}$$

We plugin the value \(i = 0\) which corresponds to \(\Delta ^{(j)}\mathbf {r}(\mathbf {Y}_0) \equiv \Delta ^{(j)}\mathbf {r}(\mathbf {y}_n)\) and we get,

$$\begin{aligned} \Delta ^{(j)}\mathbf {r}(\mathbf {y}_n) = \displaystyle \sum _{k=0}^{j} (-1)^k {j \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{j-k}). \end{aligned}$$

WLOG replacing j by \(j-1\) yields

$$\begin{aligned} \Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_n) = \displaystyle \sum _{k=0}^{j-1} (-1)^k {j-1 \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{j-1-k}). \end{aligned}$$
(57)

Since the left-hand side of Eq. (57) can be written as

$$\begin{aligned} \Delta ^{(j-1)}\mathbf {r}(\mathbf {y}_{n}) = \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }, \end{aligned}$$

we have the following result

$$\begin{aligned} \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot } = \displaystyle \sum _{k=0}^{j-1} (-1)^k {j-1 \atopwithdelims ()k} \mathbf {r}(\mathbf {Y}_{j-1-k}). \end{aligned}$$
(58)

The remainder function \(\mathbf {r}(\mathbf {Y}_i)\) is defined as

$$\begin{aligned} \mathbf {r}(\mathbf {Y}_i) = \mathbf {f}(\mathbf {Y}_i) - \mathbf {f}(\mathbf {y}_n) - \mathbf {A}_{n} * (\mathbf {Y}_i - \mathbf {y}_n). \end{aligned}$$

Plugging in the definition of remainder function in (58) and observing that \(\mathbf {r}(\mathbf {y}_n) = 0\), we get

$$\begin{aligned} \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }= & {} \displaystyle \sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {f}(\mathbf {y}_n) - \mathbf {A}_{n} * \bigg (\mathbf {Y}_{j-1-k} - \mathbf {y}_n\bigg )\bigg )\nonumber \\= & {} \displaystyle \sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {A}_{n} * \mathbf {Y}_{j-1-k}\bigg ) \nonumber \\&\displaystyle -\,\sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n} * \mathbf {y}_{n}\bigg ). \end{aligned}$$
(59)

Consider the identity involving alternating sum and difference of binomial coefficients,

$$\begin{aligned} \displaystyle \sum _{i = 0}^{k} (-1)^{i} {k \atopwithdelims ()i} = 0. \end{aligned}$$
(60)

Applying (60) to (59) we get,

$$\begin{aligned} \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot }= & {} \displaystyle \sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {f}(\mathbf {y}_n) - \mathbf {A}_{n} * \bigg (\mathbf {Y}_{j-1-k} - \mathbf {y}_n\bigg )\bigg )\nonumber \\= & {} \displaystyle \sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {A}_{n} * \mathbf {Y}_{j-1-k}\bigg )\nonumber \\&- \,(-(-1)^{(j-1)}) \bigg (\mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n} * \mathbf {y}_{n}\bigg ) \nonumber \\= & {} \displaystyle \sum _{k=0}^{j-2} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {A}_{n} * \mathbf {Y}_{j-1-k}\bigg )\nonumber \\&+\, (-1)^{(j-1)} {j-1 \atopwithdelims ()j-1} \bigg (\mathbf {f}(\mathbf {y}_{n}) - \mathbf {A}_{n} * \mathbf {y}_{n}\bigg ) \nonumber \\= & {} \displaystyle \sum _{k=0}^{j-1} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {A}_{n} * \mathbf {Y}_{j-1-k}\bigg ). \end{aligned}$$
(61)

Additionally, since we are replacing the Jacobian by the approximation in the Krylov-subspace, i.e. \(\mathbf {A}_{n} = \mathbf {V}\,\mathbf {H}\,\mathbf {V}^T\) we have

$$\begin{aligned} \mathbf {V}\mathbf {d}_{(j-1)} + \mathbf {\widetilde{r}}_{(j-1)}^{\bot } = \displaystyle \sum _{k=0}^{j-1} (-1)^k {j-1 \atopwithdelims ()k} \bigg (\mathbf {f}(\mathbf {Y}_{j-1-k}) - \mathbf {V}\mathbf {H}\mathbf {V}^{T} * \mathbf {Y}_{j-1-k}\bigg ). \end{aligned}$$
(62)

We get the expression for \(\mathbf {d}_{(j-1)}\) by multiplying Eq. (62) from the left by \(V^{T}\) and that for \(\mathbf {\widetilde{r}}_{(j-1)}^{\bot }\) by multiplying by \((\mathbf {I} - \mathbf {V}\mathbf {V}^{T})\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Narayanamurthi, M., Tranquilli, P., Sandu, A. et al. EPIRK-W and EPIRK-K Time Discretization Methods. J Sci Comput 78, 167–201 (2019). https://doi.org/10.1007/s10915-018-0761-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-018-0761-3

Keywords

Mathematics Subject Classification

Navigation