Runge–Kutta methods determined from extended phase space methods for Hamiltonian systems
Abstract
We study two existing extended phase space integrators for Hamiltonian systems, the midpoint projection method and the symmetric projection method, showing that the first is a pseudosymplectic and pseudosymmetric Runge–Kutta method and the second is a monoimplicit symplectic Runge–Kutta method.
1 Introduction
Many commonly used numerical methods for the time integration of differential equations can be expanded in B-series which elucidate their geometric and numerical properties [2, 10]. However, symplectic integrators with a B-series are implicit, the implicit midpoint rule being a central example [5]. Explicit symplectic integrators exist for some systems, such as separable classical mechanical systems [13]. To avoid this restriction, Pihajoki [15] introduced extended phase space methods: a new Hamiltonian, defined on the product of two copies of the original phase space, is constructed, that is amenable to explicit symplectic integration in the extended phase space.
In place of the Hamiltonian system associated with the Hamiltonian and canonical symplectic form ,
(1) |
(), Pihajoki considered the extended system
() with initial condition
such that the solution obeys and for all and satisfies the original system (1) with initial condition . The extended system is Hamiltonian with extended Hamiltonian , , and symplectic form .
As and (resp. and ) commute, the flow of Hamilton’s equations for is given explicitly by Euler’s method:
(and analogously for ). The integrator
is therefore explicit, second order, and preserves the extended symplectic form .
Two key issues, however, immediately arise: the ‘duplicate’ point may move away from the ‘base’ point ; and it is not clear how symplecticity in the extended phase space is advantageous in the original phase space. Tao [16] addressed the first point by adding a coupling term to the extended Hamiltonian, finding that this could suppress the growth of . Other authors have addressed the second point by projecting the solution to the original phase space in different ways. Two of these, the symmetric projection method of Ohsawa [14] and the midpoint projection method of Luo et al. [7], are the subject of this paper.
The midpoint projection method, considered in Section 2, is shown to be equivalent to an explicit Runge–Kutta method that is pseudosymplectic (that is, approximately symplectic) and pseudosymmetric up to surprisingly high order – order 5 for the leapfrog-based method of classical order 2, and order 9 for the methods of classical order 4. We suggest that these properties account for the methods’ good performance in astrophysical applications.
The symmetric projection methods, considered in Section 3, are shown to be equivalent to monoimplicit symplectic Runge–Kutta methods, revealing their affine equivariance and generality.
2 The midpoint projection method
Luo et al. [7] composed such an extended phase space integrator with the midpoint projection111Called the midpoint permutation in [7].
to yield an explicit integrator on the original phase space. These methods were called ‘symplectic-like’ ‘because they, like standard implicit symplectic integrators, show no drift in the energy error’ [7]. This lack of energy drift has been observed in many astrophysical simulations with nonseparable Hamiltonians without explanation, and the method has become quite popular [6, 18]. The order can be increased using composition methods [13]. The following result accounts for the greatly reduced energy drift.
Proposition 1.
The -stage midpoint projected methods of the form
are equivalent to -stage explicit Runge–Kutta methods of at least the same classical order as the underlying composition method. In the three cases
-
1.
, (the standard extended phase space integrator with midpoint projection, of classical order 2);
-
2.
, , (classical order 4);
-
3.
, , (classical order 4).
the methods have pseudosymplecticity order , , and respectively, and pseudosymmetry order , , and respectively. That is, and .
Proof.
We begin by noting that the extended phase space methods do not rely on the partitioning into variables, but can be written in an affine-equivariant way that exhibits how they can be applied to any ordinary differential equation. For the ODE , , we consider the duplicated (i.e. extended) system
(2) |
which is separable and can be integrated by splitting and composition. When , , and , this yields the method above. When preserves , (2) preserves . We write out the method first in variables, with initial conditions :
Imposing , this can be written in Runge–Kutta form as
This is a 3-stage explicit Runge–Kutta method with Butcher tableau
For this method we compute its B-series and check the pseudosymplecticity conditions222We evaluated the symplecticity conditions where and are Butcher trees in Mathematica using
Ω<<NumericalDifferentialEquationAnalysis‘ΩButcherProduct[u_, v_] := If[ByteCount[v]==0, \[FormalF][u], ReplacePart[v, 1->v[[1]] u]]ΩSymplectic[u_, v_] := ButcherPhi[u] ButcherPhi[v] - ButcherPhi[ButcherProduct[u, v]]Ω- ButcherPhi[ButcherProduct[v, u]]Ω\end{verbatim}Ω
[5, VI.7.3]. There are 1, 1, 1, 3, and 6 conditions respectively at orders 1, …, 5; these are all satisfied. Of the 16 order 6 conditions, 13 are satisfied and 3 are not, thus the method is pseudosymplectic of order 5. For pseudosymmetry, we expand in B-series similarly.
The calculation for the other methods proceeds similarly. For the Butcher tableau is
∎
Pseudosymplecticity was introduced by Aubry and Chartier [1], who derive various explicit pseudosymplectic Runge–Kutta methods; the second order method above appears there as a member of a 1-parameter family of 3-stage methods of pseudosymplectic order 4, it being the unique member of that family that is pseudosymplectic of order 5.
The numerical examples in the astrophysics literature do not show energy drift. However, the potential drift effect is rather small and we have confirmed numerically in planar Hamiltonian systems that the energy does drift proportionally to resp. for the 2nd resp. 4th order methods above. Symmetry also affects time integration and can moderate energy drift [12], so it is possible that pseudosymmetry is having some positive effect as well. The 3-stage 4th order method is known to have large error constants and it is possible that the 5-stage method given here, or some other explicit pseudosymplectic method, may be have advantages in these applications. On the other hand, energy behaviour is not the only manifestation of symplecticity and the choice of a symplectic vs a pseudosymplectic integrator may depend on the application.
3 The symmetric projection method
Ohsawa [14] defined the extended phase space integrator with symmetric projection as follows. Let
Given an extended phase space integrator and initial condition , the integrator is the map defined by
-
1.
-
2.
Find such that
-
3.
-
4.
-
5.
-
6.
The method is well-defined, symmetric, and symplectic [14]. Jayawardana and Ohsawa [8] further show that the method preserves arbitrary quadratic invariants. Together these results give a strong indication that the method may be a symplectic B-series method. We show below that this is true and that it is in fact equivalent to a monoimplicit Runge–Kutta method, a class introduced by Cash [3] that have only a single implicit stage. (The Simpson–AVF method is an example [4].) In a monoimplicit Runge–Kutta method the matrix of the Butcher tableau takes the form of a rank one matrix plus a matrix of zero spectral radius.
The extended phase space integrator with symmetric projection is also an instance of the class of extended Runge–Kutta methods that we now define.
Definition 1.
For ODEs , , an extended Runge–Kutta method is defined by the equations
where the matrix has full rank.
Recall the central result of Munthe-Kaas et al. [9]: Let be an integration method, defined for all vector fields on all dimensions . Then is a B-series method if and only if the property of affine equivariance is fulfilled: if is an affine map from to , a vector field on , and a vector field on such that , then .
Proposition 2.
Extended Runge–Kutta methods are affine equivariant.
Proposition 3.
Let be the matrix defined by
for , where we have defined for . Let be an matrix whose columns form a basis for the nullspace of . If for , and , then the extended Runge–Kutta method with parameters , , and , is quadratic-preserving and symplectic.
Proof.
Suppose has a quadratic first integral . Following the standard proof for Runge–Kutta methods,
Using the condition on the , and expressing in the basis whose columns are , i.e. , gives the result. ∎
Proposition 4.
Extended phase space integrators with symmetric projection, where the extended method is a composition method, are extended Runge–Kutta methods and can be written as monoimplicit symplectic Runge–Kutta methods.
Corollary 1.
-
1.
Extended phase space integrators with symmetric projection methods preserve arbitrary affine symmetries, quadratic integrals, and constant symplectic structures when there are any.
-
2.
Monoimplicit symplectic Runge–Kutta methods of all orders exist.
Proof.
We again write the extended system as
Let ; ; . In these variables the symmetric projection method takes the form
where is determined by the condition that .
We first illustrate the complete construction for the case that is the leapfrog method . Its three substeps are then
Defining the three stages , , and as the -values at which is evaluated, and defining , we have
The condition is
which leads to the constraint
The update equation is or
Thus the parameters for this method are
(For the null space of we take
A direct calculation shows that
and , confirming that the method is quadratic-preserving.)
The extra stage can be explicitly eliminated using the constraint, giving the 3-stage monoimplicit symplectic Runge–Kutta method with tableau
For the general case, let be the composition method with time steps given by parameters , i.e.
Writing out the stages as above, and eliminating the constraint , leads to an -stage monoimplicit Runge–Kutta method with parameters
where
Therefore
where
and thus we have
for all and , which is the condition for a Runge–Kutta method to be symplectic. ∎
4 Discussion
The results here have been established by direct calculation, but the methods appear quite natural and intrinsically defined. One could seek direct geometric proofs, not relying on B-series, of the pseudosymplecticity and pseudosymmetry orders for the midpoint projection method when the extended method is an arbitrary symplectic integrator. Likewise, for the symmetric projection method, the diagonal is a symplectic subspace of the extended symplectic space, suggesting an approach using the symplectic geometry of constraints [11]. The question mentioned earlier, of the best pseudosymplectic method to use in applications, and any potential issues arising from nonsymplecticity, should be examined. Finally, we suggest determining the entire set of monoimplicit symplectic Runge–Kutta methods and their relative merits.
Dedication
A preliminary version of this work was presented at ANODE 2023 in honour of John Butcher’s 90th birthday. It was therefore very pleasing and appropriate to find that, as the work developed, it turned out to involve Runge–Kutta methods and Butcher series so intimately. Happy birthday, John!
References
- [1] Aubry, A. and Chartier, P. (1998). Pseudo-symplectic Runge–Kutta methods. BIT Numerical Mathematics, 38, 439–461.
- [2] Butcher, J. C. (2021). B-series: Algebraic Analysis of Numerical Methods. Springer Series in Computational Mathematics, vol. 55. Berlin: Springer.
- [3] Cash, J. R. (1975). A class of implicit Runge-Kutta methods for the numerical integration of stiff ordinary differential equations. Journal of the ACM (JACM), 22(4), 504-511.
- [4] Celledoni, E., McLachlan, R. I., McLaren, D. I., Owren, B., Quispel, G. R. W., and Wright, W. M. (2009). Energy-preserving Runge–Kutta methods. ESAIM: Mathematical Modelling and Numerical Analysis, 43(4), 645–649.
- [5] Hairer, E., Lubich, C., and Wanner, G. (2006). Geometric Numerical Integration: Structure-Preserving Algorithms for Ordinary Differential Equations (Springer Series in Computational Mathematics, 31) Berlin: Springer.
- [6] Li, D. and Wu, X. (2019). Chaotic motion of neutral and charged particles in a magnetized Ernst-Schwarzschild spacetime. The European Physical Journal Plus, 134, 1–12.
- [7] Luo, J., Wu, X., Huang, G., and Liu, F. (2017). Explicit symplectic-like integrators with midpoint permutations for spinning compact binaries. The Astrophysical Journal, 834(1), 64.
- [8] Jayawardana, B., and Ohsawa, T. (2023). Semiexplicit symplectic integrators for non-separable Hamiltonian systems. Mathematics of Computation, 92(339), 251–281.
- [9] McLachlan, R. I., Modin, K., Munthe-Kaas, H., and Verdier, O. (2016). B-series methods are exactly the affine equivariant methods. Numerische Mathematik, 133, 599–622.
- [10] McLachlan, R. I., Modin, K., Munthe-Kaas, H., and Verdier, O. (2017). Butcher series. A story of rooted trees and numerical methods for for evolution equations, Asia Pacific Mathematics Newsletter 7, No 1, (2017), 1–11.
- [11] McLachlan, R. I., Modin, K., Verdier, O., and Wilkins, M. (2014). Geometric generalisations of Shake and Rattle. Foundations of Computational Mathematics 14 (2)2 339–370.
- [12] McLachlan, R. I., and Perlmutter, M. (2004). Energy drift in reversible time integration. Journal of Physics A: Mathematical and General, 37(45), L593.
- [13] McLachlan, R. I. and Quispel, G. R. W. (2002). Splitting methods. Acta Numerica, 11, 341-434.
- [14] Ohsawa, T. (2023). Preservation of Quadratic Invariants by Semiexplicit Symplectic Integrators for Nonseparable Hamiltonian Systems. SIAM Journal on Numerical Analysis, 61(3), 1293–1315.
- [15] Pihajoki, P. (2015).Explicit methods in extended phase space for inseparable Hamiltonian problems. Celestial Mechanics and Dynamical Astronomy, 121(3):211–231.
- [16] Tao, M. (2016) Explicit high-order symplectic integrators for charged particles in general electromag- netic fields. Journal of Computational Physics, 327:245–251.
- [17] Tao, M. (2016) Explicit symplectic approximation of nonseparable Hamiltonians: Algorithm and long time performance. Physical Review E, 94(4):043303.
- [18] Pan, G., Wu, X., and Liang, E. (2021). Extended phase-space symplectic-like integrators for coherent post-Newtonian Euler-Lagrange equations. Physical Review D, 104(4), 044055.