Next Article in Journal
An Algorithm for the Numbers of Homomorphisms from Paths to Rectangular Grid Graphs
Previous Article in Journal
A Markov Decision Process with Awareness and Present Bias in Decision-Making
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On Reusing the Stages of a Rejected Runge-Kutta Step

by
Vladislav N. Kovalnogov
1,
Ruslan V. Fedorov
1,
Tamara V. Karpukhina
1,
Theodore E. Simos
1,2,3,4,5,6,* and
Charalampos Tsitouras
7
1
Laboratory of Inter-Disciplinary Problems of Energy Production, Ulyanovsk State Technical University, 32 Severny Venetz Street, 432027 Ulyanovsk, Russia
2
Center for Applied Mathematics and Bioinformatics, Gulf University for Science and Technology, West Mishref 32093, Kuwait
3
Department of Medical Research, China Medical University Hospital, China Medical University, Taichung City 40402, Taiwan
4
Data Recovery Key Laboratory of Sichun Province, Neijiang Normal University, Neijiang 641100, China
5
Section of Mathematics, Department of Civil Engineering, Democritus University of Thrace, 67100 Xanthi, Greece
6
Department of Mathematics, University of Western Macedonia, 50100 Kozani, Greece
7
General Department, National & Kapodistrian University of Athens, Euripus Campus, 34400 Evia, Greece
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(11), 2589; https://doi.org/10.3390/math11112589
Submission received: 9 May 2023 / Revised: 2 June 2023 / Accepted: 4 June 2023 / Published: 5 June 2023
(This article belongs to the Section Engineering Mathematics)

Abstract

:
Runge-Kutta (RK) pairs are amongst the most popular methods for numerically solving Initial Value Problems. While using an RK pair, we may experience rejection of some steps through the interval of integration. Traditionally, all of the evaluations are then dropped, and we proceed with a completely new round of computations. In this work, we propose avoiding this waste and continuing by reusing the rejected RK stages. We focus especially on an RK pair of orders six and five. After step rejection, we reuse all the previously evaluated stages and only add three new stages. We proceed by evaluating the output using a smaller step. By this technique, we manage to significantly reduce the cost in a set of problems that are known to pose difficulties in RK algorithms with changing step sizes.
MSC:
65L05; 65L06

1. Introduction

A fundamental issue in mathematical analysis, the Initial Value Problem (IVP) has applications in a variety of disciplines, including physics, engineering, and economics. The IVP has the following form
y = f ( x , y ) , y ( x 0 ) = y 0 R m , x [ x 0 , x e ] ,
with f : R × R m R m continuously differentiable.
The IVP’s purpose is to determine the solution to a differential equation given the starting condition, which is the value of the dependent variable at a specified initial time. The most frequent approach for solving the IVP is to use numerical methods, which entail estimating the solution of the differential equation using discrete time steps.
The Runge-Kutta (RK) method is one of the most prominent and commonly used numerical methods for solving the IVP. The RK technique was developed in the late 1800s by German mathematicians Carl Runge and Martin Kutta. It is a class of numerical methods that estimate the value of the solution at the end of a time step by taking a weighted average of function values at different points in the time step. The algebraic order of the RK method relates to the number of function evaluations for each time step, which impacts the technique’s accuracy.
For solving the IVP, the RK approach offers significant benefits over other numerical methods. It is a basic and easy-to-implement approach that is understandable to non-mathematicians. It is also a flexible approach that may be used to solve a broad variety of differential equations, including stiff problems that are difficult to solve with other numerical methods. Furthermore, the RK approach is a self-starting technique, which implies that no extra equations or conditions must be solved.
However, the RK method’s accuracy is dependent on the sequence of the process, which increases the technique’s computing cost and complexity. To overcome this issue, the Runge-Kutta pair, a version of the RK technique, was devised. A Runge-Kutta pair is a pair of RK methods using the same body of function evaluations but distinct weight coefficients that are used to estimate the solution and error at each time step. The estimated error is utilised to adjust the step size in the following time step, leading to a more accurate and efficient technique than the single RK method.
The general form of an s-stage Runge-Kutta pair of orders p and p 1 can be written as follows (remark that clearly s > p for p > 4 ):
y n + 1 = y n + h n i = 1 s b i k n i , p th order aproximation , y ^ n + 1 = y n + h n i = 1 s b ^ i k n i , ( p 1 ) th order aproximation ,
with
k n i = f ( x n + c i h n , y n + h n j = 1 i 1 a i j k n j ) , i = 1 , 2 , , s ,
where y n is the approximate solution at time x n , h n is the time step size, k n i are the intermediate function values, and b i , b ^ i , c i , and a i j are the RK coefficients. We assume that the coefficients form the column vectors b T , b ^ T c R s and the matrix A R s × s . The error estimate can be calculated as follows:
e n + 1 = y n + 1 y ^ n + 1 = h n i = 1 s ( b i b ^ i ) k n i .
By comparing the error estimate to a specified tolerance level T O L and modifying the time-step size accordingly, the step size may be adjusted. This enables a more efficient and precise technique for resolving the IVP. For achieving this, we use the formula
h n + 1 = 0.9 · h n · T O L e n + 1 ( 1 / p ) .
In case e n + 1 > T O L , the step is rejected, i.e., we do not move forward from x n to x n + 1 . Then, we again use (3) but replace the left-hand side with h n . This is a classical procedure: see [1] (p. 167).
There have been several studies on the performance and accuracy of Runge-Kutta pairs in solving the IVP. Among the first such studies, Fehlberg compared the performance of different RK pairs of order 5 and 6 on a set of test problems and found that a particular RK pair, known as the Fehlberg pair, was the most efficient and accurate pair for a wide range of problems [2]. These pairs by Fehlberg are quadrature defective. Thus, for quadratic problems, the lower and higher order results coincide, and error control fails.
In another study, Prince and Dormand introduced a new RK pair of order 6 and 5 that avoided this defect and showed improved performance compared to other RK pairs [3].
In recent years, there have been developments in using RK pairs in combination with adaptive time-stepping algorithms to solve the IVP. One such algorithm is the Dormand–Prince method, which uses an embedded RK pair of order 5 and 4 with adaptive time-step size control to improve the accuracy and efficiency of the method [4]. Another algorithm is the Tsitouras method, which uses an embedded RK pair of order 5 and 6 with adaptive time-step size control; it has been shown to be more efficient and accurate than other RK pairs on certain types of problems [5].
In conclusion, the Runge-Kutta pair is a powerful and versatile numerical method for solving the IVP. It provides a more accurate and efficient method than using a single RK method by using a pair of methods of the same order with different coefficients to estimate the value of the solution and the error at each time step. There have been several studies on the performance and accuracy of RK pairs, with some pairs designed specifically for stiff equations and others optimized for efficiency and accuracy on a wide range of problems. Adaptive time-stepping algorithms, such as the Dormand–Prince and Tsitouras pairs, have been developed to further improve the accuracy and efficiency of the RK pair method.

2. The New Step-Size Control Scheme

After using the devised Equations (2) and (3) for changing the step size, we may experience step rejection. Then, traditionally, all the stages are completely wasted and we proceed by evaluating a new smaller step. This unjustified loss will be discussed here. We intend to suggest extension of existing pairs in order to handle these stages of a rejected step again and reduce the cost.
Thus, after an s-stage RK pair is applied and the error estimation fails to stay below ϵ n + 1 , we reduce the current step to have a new length, say τ h n , τ < 1 . We then extend the pair by adding some more stages. After having at hand the s stages of the underlying pair, given by (1), we additionally compute
k n i = f ( x n + c i h n , y n + h n j = 1 i 1 a i j k n j ) , i = s + 1 , s + 2 , , s ˜ ,
and combine all these s ˜ stages in a new pair of formulas
y n + 1 * = y n + h n i = 1 s ˜ b i * k n i , p th order aproximation , y ^ n + 1 * = y n + h n i = 1 s ˜ b ^ i * k n i , ( p 1 ) th order aproximation .
i.e., the new weights b * , b ^ * are used in order to approximate y n + 1 * y ( x n + τ h n ) .
The error estimate now becomes
e n + 1 * = y n + 1 * y ^ n + 1 * = h n i = 1 s ˜ ( b i * b ^ i * ) k n i ,
and the slightly modified formula
h n + 1 = 0.9 · h n · T O L e n + 1 * ( 1 / p ) .
is used for advancing the integration.
In consequence, we spend only ( s ˜ s ) stages after a step rejection instead of the s-stages lost with the standard step-size-control algorithm. Obviously, we clearly demand ( s ˜ s ) < s .
For safety reasons, we only apply the new algorithm in case
T O L < ϵ n + 1 < λ · T O L ,
for some λ > 1 . Otherwise, we continue with the standard policy (2)–(3).
We assume that the new extended set of coefficients now forms the column vectors b * T , b ^ * T c R s ˜ and the new extended matrix A R s ˜ × s ˜ .

3. Evaluation of a Mew-Extended Runge-Kutta Pair

In the present work, we concentrate on a pair of orders six and five. We choose the DLMP6(5) pair that appeared in [6]. This pair seems to be among the best choices for moderate tolerances (i.e., 10 4 > T O L > 10 9 , see [5]). For this case, s = 9 , but since
a 9 j = b j , j = 1 , 2 , , 9 ,
the last stage of each step can be used as the first stage of the next step. This is called FSAL (First Stage As Last), and then DLMP6(5) only actually spends 8 stages per step. The same number of stages is wasted after a step rejection using (2) and (3).
We set τ = 0.8 and s ˜ = 12 . The new extended pair (named DLMP6(5)ext) shares an enhanced set of coefficients, as shown in the previous subsection. When a step is rejected, we evaluate these three new stages, k n , 10 , k n , 11 and k n , 12 , and combine them with the new set of weights b * , b ^ * to form the new approximations for the point x n + τ h n .
Thus, we have to evaluate the coefficients
a 10 , 1 , a 10 , 3 , a 10 , 4 , , a 10 , 9 , a 11 , 1 , a 11 , 3 , a 11 , 4 , , a 11 , 10 , a 12 , 1 , a 12 , 3 , a 12 , 4 , , a 12 , 11 ,
c 10 , c 11 , c 12 , b 1 * , b 4 * , b 5 * , , b 12 * , b ^ 1 * , b ^ 4 * , b ^ 5 * , , b ^ 11 * ,
since for the considered type of pairs, it is appropriate to set
a 10 , 2 = a 11 , 2 = a 12 , 2 = b 2 * = b 3 * = b ^ 2 * = b ^ 3 * = b ^ 12 * = 0 .
The coefficients of DLMP6(5) along with the above ones have to satisfy the order conditions listed in Table 1 and Table 2.
In these Tables, we mention with d j i the j th-order conditions. By *, we denote component-wise multiplication among vectors. This multiplication has lesser priority than the dot product. We also use the notation c 2 = c c , c 3 = c c c , etc. By A 2 = A · A or A c = A · c , we denote the classical dot product. We also set I = [ 1 , 1 , , 1 ] T R s ˜ .
For solving the above equations, we make some simplifying assumptions. Namely
A I = c , ( A · c = 1 2 c 2 ) ( 3 + ) , ( A · c 2 = 1 3 c 3 ) ( 3 + ) ,
and
b · ( A + C τ I s ˜ ) = 0 s ˜ .
In the above, we use the symbolization v ( i + ) with the elements of some vector v but with its first ( i 1 ) elements dropped. We set C = diag ( c ) , and by I s ˜ and 0 s ˜ , we mean, respectively, the identity and zero matrix in R s ˜ × s ˜ .
Then, there is a severe reduction in the number of the order conditions to be solved. We are able to analytically solve these remaining order conditions with respect to the new coefficients. Indeed, when the above assumptions hold, only the order conditions
d 6 , 1 , d 5 , 1 , d 4 , 1 , d 3 , 1 , d 2 , 1 , d 1 , 1 , d 5 , 4 , d 5 , 5 , d 6 , 16 , d 6 , 8 , d 6 , 11 , d 6 , 19 ,
and
d ^ 5 , 1 , d ^ 4 , 1 , d ^ 3 , 1 , d ^ 2 , 1 , d ^ 1 , 1 , d ^ 5 , 4 , d ^ 5 , 5 ,
have to be solved. The task is becoming a lot easier. We even manage to solve the seventh-order equations of condition for the higher-order formula. The resulting pair is shown in Table 3. These coefficients can be retrieved in Mathematica [7] format from http://users.uoa.gr/~tsitourasc/dlmp65ext.m (accessed on 3 June 2023).
An interpolant of the sixth-order at the same cost can be constructed for this type of pair [8]. Then, we may freely choose the magnitude of τ and vary the lengths of the new step after rejection. By fixing τ = 0.8 and achieving locally seventh-order accuracy after rejection, we obtain better results overall.

4. Numerical Results

We have chosen the detest problems D4, D5, and E2 as the set to perform our tests on. Integrating these problems by any 6 ( 5 ) pair, we observe that almost 25 % of the computational cost is wasted upon rejected steps. The problems chosen are given below.
Problem D4
This is the two-body problem with eccentricity e = 0.7 , and it has the following form [9]
y 1 = y 3 , y 2 = y 4 , y 3 = y 1 ( y 1 2 + y 2 2 ) 3 / 2 , y 4 = y 2 ( y 1 2 + y 2 2 ) 3 / 2 , y 1 ( 0 ) = e = 7 10 , y 2 ( 0 ) = 0 , y 3 ( 0 ) = 0 , y 4 ( 0 ) = 17 3 , x [ 0 , 20 ] .
Notice that subscripts above denote the component of the problem and not time steps. We solve the above equations in the interval 0 , 20 , and the theoretical solution is given in [10].
After we run this problem for tolerances T O L = 10 4 , 10 5 , , 10 9 and λ = 7 , we record in Table 4 and Table 5 all the characteristics of the results. In the final column, we present an efficiency measure according to the formula [11,12]
efficiency = ( total function evaluations ) · ( end point error ) 1 / 6 .
This information is perhaps more significant because it ensures that we do not simply decrease the number of rejected steps with a very conservative step-size algorithm, which, in reverse, decreases efficiency.
Problem D5
This is again the two-body problem (i.e., Kepler problem) as above but with eccentricity e = 0.9 , and the initial conditions are now
y 1 ( 0 ) = e = 9 10 , y 2 ( 0 ) = 0 , y 3 ( 0 ) = 0 , y 4 ( 0 ) = 19 , x [ 0 , 20 ] .
The corresponding results are given in Table 6 and Table 7.
Problem E2
This is the Van der Pol oscillator driven by the equations
y 1 = y 2 , y 2 = ( 1 y 1 2 ) y 2 y 1 , y 1 ( 0 ) = 2 , y 2 ( 0 ) = 0 , x [ 0 , 20 ] .
The corresponding results are given in Table 8 and Table 9.
Arenstorf orbit
A fascinating orbit displays a spacecraft’s steady journey around the Earth and Moon [1] (p. 297). This can be given in the form of (1) as follows
y 1 = y 3 , y 2 = y 4 , y 3 = ψ ψ cos x + y 1 P 1 + ψ ψ cos x y 1 P 2 , y 4 = ψ ψ sin x + y 2 P 1 + ψ ψ sin x y 2 P 2 ,
with
P 1 = ( y 1 + ψ cos x ) 2 + ( y 2 + ψ sin x ) 2 1.5 , P 2 = ( y 1 ψ cos x ) 2 + ( y 2 ψ sin x ) 2 1.5 , ψ = 0.012277471 , ψ = 0.987722529 ,
and initi · al values
y 1 ( 0 ) = 0.994 , y 3 ( 0 ) = 0 , y 2 ( 0 ) = 0 , y 4 ( 0 ) = 1.007585106379082 ,
and the solution is periodic with period x A = 17.0652165601579625589 . We solved the Arenstorf problem and recorded the cost (i.e., the stages used) and the errors observed for x A .
The corresponding results are given in Table 10 and Table 11.
Interpreting the results, we observe about 28%–29% improvement in the efficiency for the problems and accuracies chosen. This can be found by dividing the corresponding values in the last column of Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10 and Table 11: e.g., for problem D4 and T O L = 10 4 , we observe that the standard algorithm ((2) and (3)) is about 213.6 150.3 = 42.1 % more expensive than the new algorithm. In all 24 runs, we experience better efficiency with the new algorithm. Another interesting issue is that the error is decreasing, since after step rejection, we apply locally a seventh-order approximation.
The new algorithm proposed in the paper could be applied to fractional-order differential equations after some proper modifications [13,14].

5. Conclusions

A new step-control algorithm was presented in this article that, after the rejection of a Runge-Kutta step, avoids the unnecessary waste of the already-made computations. Numerical results on problems with a high percentage of rejected steps show the clear advantages of the new step-size policy.

Author Contributions

Conceptualization, V.N.K., R.V.F., T.V.K., T.E.S. and C.T.; Methodology, T.E.S. and C.T.; Software, T.E.S. and C.T.; Validation, V.N.K., R.V.F., T.V.K., T.E.S. and C.T.; Formal analysis, T.E.S. and C.T.; Investigation, V.N.K., R.V.F., T.V.K., T.E.S. and C.T.; Resources, T.E.S. and C.T.; Data curation, V.N.K., R.V.F., T.V.K., T.E.S. and C.T.; Writing—original draft, C.T.; Writing—review & editing, T.E.S. and C.T.; Visualization, V.N.K., R.V.F., T.V.K., T.E.S. and C.T.; Supervision, T.E.S.; Project administration, T.E.S.; Funding acquisition, V.N.K. All authors contributed equally. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by a Mega Grant from the Government of the Russian Federation within the framework of federal project No. 075-15-2021-584.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The coefficients of the method presented can be retrieved from http://users.uoa.gr/~tsitourasc/dlmp65ext.m (accessed on 3 June 2023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hairer, E.; Norsett, S.P.; Wanner, G. Solving Ordinary Differential Equations I: Nonstiff Problems; Springer: Berlin/Heidelberg, Germany, 1987. [Google Scholar]
  2. Fehlberg, C. Classical Fifth-, Sixth-, Seventh-, and Eighth-Order Runge-Kutta Formulas with Stepsize Control; NASA Technical Report-287; National Aeronautics and Space Administration (NASA): Washington, DC, USA, 1969.
  3. Prince, P.J.; Dormand, J.R. High order embedded Runge-Kutta Formulae. J. Comput. Appl. Math. 1981, 7, 67–75. [Google Scholar] [CrossRef] [Green Version]
  4. Dormand, J.R.; Prince, P.J. A family of embedded Runge-Kutta formulae. J. Comput. Appl. Math. 1980, 6, 19–26. [Google Scholar] [CrossRef] [Green Version]
  5. Tsitouras, C. A parameter study of explicit Runge-Kutta pairs of orders 6(5). Appl. Math. Lett. 1998, 11, 65–69. [Google Scholar] [CrossRef] [Green Version]
  6. Dormand, J.R.; Lockyer, M.A.; McGorrigan, N.E.; Prince, P.J. Global error estimation with Runge-Kutta triples. Comput. Math. Appl. 1989, 18, 835–846. [Google Scholar] [CrossRef] [Green Version]
  7. Wolfram Research, Inc. Mathematica; Version 11.3.0; Wolfram Research, Inc.: Champaign, IL, USA, 2018. [Google Scholar]
  8. Verner, J.H. Differentiable Interpolants for high-Order Runge-Kutta Methods. SIAM J. Numer. Anal. 1993, 30, 1446–1466. [Google Scholar] [CrossRef]
  9. Tsitouras, C. A tenth order symplectic Runge–Kutta–Nyström method. Celest. Mech. Dyn. Astron. 1999, 74, 223–230. [Google Scholar] [CrossRef]
  10. Enright, W.; Pryce, J.D. Two FORTRAN packages for assessing initial value methods. ACM Trans. Math. Softw. 1987, 13, 1–27. [Google Scholar] [CrossRef]
  11. Kovalnogov, V.N.; Fedorov, R.V.; Chukalin, A.V.; Simos, T.E.; Tsitouras, C. Evolutionary Derivation of Runge–Kutta Pairs of Orders 5(4) Specially Tuned for Problems with Periodic Solutions. Mathematics 2021, 9, 2306. [Google Scholar] [CrossRef]
  12. Shampine, L.F. Some practical Runge-Kutta formulas. Math. Comp. 1986, 46, 135–150. [Google Scholar] [CrossRef]
  13. Zhao, K. Solvability and GUH-stability of a nonlinear CF-fractional coupled Laplacian equations. AIMS Math. 2023, 8, 13351–13367. [Google Scholar] [CrossRef]
  14. Zhao, K. Existence and UH-stability of integral boundary problem for a class of nonlinear higher-order Hadamard fractional Langevin equation via Mittag-Leffler functions. Filomat 2023, 37, 1053–1063. [Google Scholar]
Table 1. Order conditions for the higher-order weights of DLMP6(5)ext, i.e., b * .
Table 1. Order conditions for the higher-order weights of DLMP6(5)ext, i.e., b * .
d 1 , 1 = b * I τ , d 2 , 1 = b * c τ 2 2 , d 3 , 1 = 1 2 b * c 2 τ 3 3 ,
d 3 , 2 = b * A c τ 3 6 , d 4 , 1 = 1 6 b * c 3 τ 4 4 , d 4 , 2 = b * A c 2 τ 4 12 ,
d 4 , 3 = b * c * A c 1 24 , d 4 , 4 = b * A 2 c τ 4 24 , d 5 , 1 = 1 24 b * c 4 τ 5 5 ,
d 5 , 2 = 1 2 b * c 2 * A c τ 5 10 , d 5 , 3 = 1 2 b * A c 2 τ 5 10 , d 5 , 4 = 1 2 b * c * A c 2 τ 5 15 ,
d 5 , 5 = b * c * A 2 c τ 5 10 , d 5 , 6 = 1 6 b * A c 3 τ 5 20 , d 5 , 7 = b * A c * A c τ 5 40 ,
d 5 , 8 = 1 2 b * A 2 c 2 τ 5 60 , d 5 , 9 = b * A 3 c τ 5 120 , d 6 , 1 = 1 120 b * c 5 τ 6 6 ,
d 6 , 2 = 1 6 b * c 3 * A c τ 6 12 , d 6 , 3 = 2 b * c * A c 2 τ 6 24 , d 6 , 4 = 1 4 b * c 2 * A c 2 τ 6 18 ,
d 6 , 5 = 1 2 b * c 2 * A 2 c τ 6 36 , d 6 , 6 = 1 2 b * A c * A c 2 τ 6 36 , d 6 , 7 = 1 2 b * A c * A 2 c τ 6 72 ,
d 6 , 8 = 1 6 b * c * A c 3 τ 6 24 , d 6 , 9 = b * c * A c * A c τ 6 48 , d 6 , 10 = 1 2 b * c * A 2 c 2 τ 6 72 ,
d 6 , 11 = b * c * A 3 c τ 6 144 , d 6 , 12 = 1 24 b * A c 4 τ 6 30 , d 6 , 13 = 1 2 b * A c 2 * A c τ 6 60 ,
d 6 , 14 = 1 2 b * A A c 2 τ 6 120 , d 6 , 15 = 1 2 b * A c * A c 2 τ 6 90 , d 6 , 16 = b * A c * A 2 c τ 6 180 ,
d 6 , 17 = 1 6 b * A 2 c 3 τ 6 120 , d 6 , 18 = b * A 2 c * A c τ 6 240 , d 6 , 19 = 1 2 b * A 3 c 2 τ 6 360 ,
d 6 , 20 = b * A 4 c τ 6 720 .
Table 2. Order conditions for the lower-order weights of DLMP6(5)ext, i.e., b ^ * .
Table 2. Order conditions for the lower-order weights of DLMP6(5)ext, i.e., b ^ * .
d ^ 1 , 1 = b * I τ , d ^ 2 , 1 = b ^ * c τ 2 2 , d ^ 3 , 1 = 1 2 b ^ * c 2 τ 3 3 ,
d ^ 3 , 2 = b ^ * A c τ 3 6 , d ^ 4 , 1 = 1 6 b ^ * c 3 τ 4 4 , d ^ 4 , 2 = b ^ * A c 2 τ 4 12 ,
d ^ 4 , 3 = b ^ * c * A c 1 24 , d ^ 4 , 4 = b ^ * A 2 c τ 4 24 , d ^ 5 , 1 = 1 6 b ^ * c 4 τ 5 5 ,
d ^ 5 , 2 = 1 2 b ^ * c 2 * A c τ 5 10 , d ^ 5 , 3 = 1 2 b ^ * A c 2 τ 5 10 , d ^ 5 , 4 = 1 2 b ^ * c * A c 2 τ 5 15 ,
d ^ 5 , 5 = b ^ * c * A 2 c τ 5 10 , d ^ 5 , 6 = 1 6 b ^ * A c 3 τ 5 20 , d ^ 5 , 7 = b ^ * A c * A c τ 5 40 ,
d ^ 5 , 8 = 1 2 b ^ * A 2 c 2 τ 5 60 , d ^ 5 , 9 = b ^ * A 3 c τ 5 120 .
Table 3. The coefficients of the new extended pair.
Table 3. The coefficients of the new extended pair.
c 2 = 1 9 , c 3 = 1 6 , c 4 = 1 4 , c 5 = 5 9 , c 6 = 1 2 , c 7 = 48 49 ,
c 8 = 1 , c 9 = 1 , c 10 = 4 139 , c 11 = 17 38 , c 12 = 4 5 ,
b 1 * = 0.06075441182658404 , b 2 * = b 3 * = 0 ,
b 4 * = 0.25108031811087983 , b 5 * = 0.59459248062264663 ,
b 6 * = 0.58130691768291823 , b 7 * = 0.01117792906462664 ,
b 8 * = 0.001953125 , b 9 * = 0.00453876219794998 ,
b 10 * = 0.18340955527240297 , b 11 * = 0.33291925465838509 ,
b 12 * = 0.08474576271186441 , b ^ 1 * = 0.0607545222182737630 ,
b ^ 2 * = b ^ 3 * = 0 = b ^ 12 * ,          b ^ 4 * = 0.362681592201453867 ,
b ^ 5 * = 1.18886870906761734 , b ^ 6 * = 1.20278300666332157 ,
b ^ 7 * = 0.357600832335522983 ,   b ^ 8 * = 0.232809581363277529
b ^ 9 * = 0.0760545523116338381 b ^ 10 * = 0.163215379071331048
b ^ 11 * = 0.314851188060490077 b ^ 12 * = 0.0826573591413146190 , b 1 = 203 2880 ,
b 2 = b 3 = 0 = b 9 , b 4 = 30208 70785 ,
b 5 = 177147 164560 , b 6 = 536 705 ,
b 7 = 1977326743 3619661760 , b 8 = 259 720 ,
b ^ 1 = 36567 458800 , b ^ 4 = 9925984 27063465 ,
b ^ 5 = 85382667 117968950 , b ^ 6 = 310378 808635 ,
b ^ 7 = 262119736669 345979336560 , b ^ 8 = 1 2 , b ^ 8 = 101 2294 ,
a 21 = 0.111111111111111111 , a 31 = 0.0416666666666666667 ,
a 32 = 0.125 , a 41 = 0.0625 ,
a j 2 = 0 , j = 4 , 5 , , 12 , a 43 = 0.1875 ,
a 51 = 0.384087791495198903 , a 53 = 1.33744855967078189 ,
a 54 = 1.50891632373113855 , a 61 = 0.417370572207084469 ,
a 63 = 1.46730245231607629 , a 64 = 1.60862026257121625 ,
a 65 = 0.0586883824622244241 , a 71 = 0.906581932271243731 ,
a 73 = 1.98165828767968130 , a 74 = 0.967924991130227440 ,
a 75 = 7.90644976448593311 , a 76 = 8.96985927428990425 ,
a 81 = 1.23125466844812894 , a 83 = 2.33058398998453494 ,
a 84 = 1.69577556052661329 , a 85 = 10.8007435894539014 ,
a 86 = 12.5648566499630329 , a 87 = 0.0309918215538877730 ,
a 9 j = b j , j = 1 , 2 , , 8 , a 101 = 0.0276060694624219017 ,
a 103 = 0.18678058047598361 , a 104 = 0.391371551663676298 ,
a 105 = 1.09230024433914178 , a 106 = 1.22247349711209067 ,
a 107 = 0.556216395594661712 , a 108 = 0.356521739130434783 ,
a 109 = 0.126447847004327 , a 111 = 0.0192549367566782782 ,
a 114 = 0.496087246358859837 , a 115 = 1.18052838103602307 ,
a 116 = 1.29939201810168170 , a 117 = 0.586956521739130435 ,
a 118 = 0.367816091954022989 , a 119 = 0.142156862745098039 ,
a 1110 = 0.281632150794417543 ,          a 121 = 0.820970265019910839 ,
a 124 = 0.653270781790705787 ,          a 125 = 4.32243201762434916 ,
a 126 = 5.36952327363607790 ,          a 127 = 1.10690062359555245 ,
a 128 = 0.688006483439893015 ,          a 129 = 0.274081679397217048 ,
a 1210 = 0.562729086953349127 ,          a 1211 = 1.38529454069957502 ,
a 113 = 0.545453116962992122 ,          a 123 = 1.51812113592786359
Table 4. Results of DLMP6(5) over the problem D4 using the standard step-size-change algorithm.
Table 4. Results of DLMP6(5) over the problem D4 using the standard step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 60240 1.0 × 10 3 213.6
10 5 78290 2.1 × 10 4 208.5
10 6 108390 2.5 × 10 5 201.7
10 7 151480 2.6 × 10 6 186.5
10 8 215580 2.2 × 10 7 169.2
10 9 303100 2.2 × 10 8 132.1
Table 5. Results of DLMP6(5) over the problem D4 using the new step-size-change algorithm.
Table 5. Results of DLMP6(5) over the problem D4 using the new step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 58713 3.3 × 10 4 150.3
10 5 76513 6.7 × 10 5 141.0
10 6 107020 4.2 × 10 6 119.0
10 7 155026 2.0 × 10 6 151.4
10 8 224036 1.7 × 10 7 144.0
10 9 30709 2.1 × 10 8 131.2
Table 6. Results of DLMP6(5) over the problem D5 using the standard step-size-change algorithm.
Table 6. Results of DLMP6(5) over the problem D5 using the standard step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 90290 4.3 × 10 3 384.1
10 5 118450 3.0 × 10 4 337.0
10 6 160580 3.8 × 10 5 319.4
10 7 223750 3.7 . 1 × 10 6 296.8
10 8 316920 2.9 × 10 7 265.9
10 9 449640 2.6 × 10 8 224.1
Table 7. Results of DLMP6(5) over the problem D5, using the new step-size-change algorithm.
Table 7. Results of DLMP6(5) over the problem D5, using the new step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 891810 4.0 × 10 3 356.9
10 5 1172313 4.9 × 10 5 224.3
10 6 158327 2.2 × 10 5 232.8
10 7 227040 2.4 × 10 6 229.0
10 8 327052 2.0 × 10 7 215.7
10 9 457036 2.2 × 10 8 200.5
Table 8. Results of DLMP6(5) over the problem E2 using the standard step-size-change algorithm.
Table 8. Results of DLMP6(5) over the problem E2 using the standard step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 50250 2.5 × 10 4 150.9
10 5 71310 2.2 × 10 5 136.9
10 6 102320 1.7 × 10 6 117.4
10 7 146420 1.8 × 10 7 113.5
10 8 207410 1.9 × 10 8 102.1
10 9 298330 2.0 × 10 9 94.2
Table 9. Results of DLMP6(5) over the problem E2 using the new step-size-change algorithm.
Table 9. Results of DLMP6(5) over the problem E2 using the new step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 51414 5.7 × 10 5 97.3
10 5 73616 3.2 × 10 5 123.9
10 6 104021 2.4 × 10 6 105.7
10 7 149324 1.7 × 10 7 97.4
10 8 210419 1.4 × 10 8 88.1
10 9 302119 1.7 × 10 9 86.2
Table 10. Results of DLMP6(5) over the Arenstorf orbit using the standard step-size-change algorithm.
Table 10. Results of DLMP6(5) over the Arenstorf orbit using the standard step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 61230 5.4 × 10 1 605.9
10 5 81290 2.6 × 10 1 704.6
10 6 111420 3.4 × 10 3 474.6
10 7 155490 4.4 × 10 4 450.0
10 8 219470 3.6 × 10 5 387.4
10 9 315440 4.3 × 10 6 366.6
Table 11. Results of DLMP6(5) over the Arenstorf orbit using the new step-size-change algorithm.
Table 11. Results of DLMP6(5) over the Arenstorf orbit using the new step-size-change algorithm.
TOL Accepted StepsRejected StepsExtended StepsEnd Point ErrorEfficiency (4)
10 4 62165 6.6 × 10 1 595.8
10 5 81167 2.7 × 10 1 640.2
10 6 1131218 5.1 × 10 4 297.6
10 7 157026 1.6 × 10 4 310.4
10 8 225030 9.0 × 10 6 272.7
10 9 320025 3.0 × 10 6 316.9
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kovalnogov, V.N.; Fedorov, R.V.; Karpukhina, T.V.; Simos, T.E.; Tsitouras, C. On Reusing the Stages of a Rejected Runge-Kutta Step. Mathematics 2023, 11, 2589. https://doi.org/10.3390/math11112589

AMA Style

Kovalnogov VN, Fedorov RV, Karpukhina TV, Simos TE, Tsitouras C. On Reusing the Stages of a Rejected Runge-Kutta Step. Mathematics. 2023; 11(11):2589. https://doi.org/10.3390/math11112589

Chicago/Turabian Style

Kovalnogov, Vladislav N., Ruslan V. Fedorov, Tamara V. Karpukhina, Theodore E. Simos, and Charalampos Tsitouras. 2023. "On Reusing the Stages of a Rejected Runge-Kutta Step" Mathematics 11, no. 11: 2589. https://doi.org/10.3390/math11112589

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop