1,972
Views
0
CrossRef citations to date
0
Altmetric
Applied & Interdisciplinary Mathematics

Weak convergence of balanced stochastic Runge–Kutta methods for stochastic differential equations

, ORCID Icon & | (Reviewing editor:)
Article: 2163546 | Received 04 Jul 2022, Accepted 23 Dec 2022, Published online: 14 Jan 2023

ABSTRACT

In this paper, weak convergence of balanced stochastic one-step methods and especially balanced stochastic Runge–Kutta (SRK) methods for Itô multidimensional stochastic differential equations is analyzed. Generalizing a corresponding result obtained by H. Schurz for the standard Euler method, it is shown that under certain conditions, balanced one-step methods preserve the weak convergence properties of their underlying methods. As an application, this allows to prove the weak convergence order of the balanced SRK methods presented in earlier work by A. Rathinasamy, P. Nair and D. Ahmadian.

1 Introduction

Determining numerical approximations to solutions of stochastic differential equations is of great interest, as easily evaluable analytical solutions to stochastic differential equations do rarely exist (Kloeden & Platen, Citation1999; Mao, Citation2008; Milstein & Tretyakov, Citation2004). The development of approximation algorithms has therefore been an active field of research during the last decades. Weak convergence of stochastic Runge–Kutta (SRK) methods is studied in literature (Buckwar et al., Citation2010; Debrabant & Rößler, Citation2008; Milstein, Citation1985; Tocino & Vigo-Aguiar, Citation2002). Efficient SRK methods that deliver second-order weak approximations for stochastic differential equations with multiple stochastic integrals have been introduced in Debrabant & Rößler, Citation2009; Rößler (Citation2009). Tang and Xiao (Citation2017, Citation2018, Citation2019) proposed and analyzed new classes of SRK methods to further reduce the computational effort.

Stiff stochastic differential equations are studied, for example, in Abdulle et al. (Citation2013), Higham (Citation2000) and Kloeden and Platen (Citation1999). To easily handle stiff stochastic differential equation systems, Milstein, Platen and Schurz introduced the balanced method, a class of quasi-implicit numerical schemes (Milstein et al., Citation1998). Alcock and Burrage (Citation2006) examined its stability and the choice of parameters. Balanced SRK methods are studied by Amiri and Hosseini (Citation2015). Further, Mardones and Mora (Citation2019) discussed first-order weak balanced schemes to stochastic differential equations. Finally, strong convergence of two classes of second-order balanced SRK methods applied to multidimensional Itô stochastic differential equations has been discussed in Rathinasamy et al. (Citation2020).

The main contribution of the current paper is to prove, generalizing a corresponding statement in Schurz (Citation2005) for the standard Euler method, that under certain conditions, balanced one-step methods preserve the weak convergence properties of their underlying methods, and to apply this result to derive a corresponding compact result for Runge–Kutta methods. As a direct consequence, this proves the weak convergence order of the methods presented in Rathinasamy et al. (Citation2020) and Rathinasamy and Nair (Citation2018).

The paper is organized as follows: In Section 2, some basic results and definitions are gathered. In Section 3, the weak convergence of balanced one-step methods and balanced SRK methods is proven.

2 Preliminaries

Let (Ω,F,P) be a complete probability space with a filtration (Ft)t0 satisfying the usual conditions and I=[t0,T] for some 0t0<T<. Consider the system of Itô stochastic differential equations of the form

(1) dXt=g0(X(t))dt+k=1mgk(X(t))dWk(t),X(t0)=X0,t[t0,T],(1)

where X(t)Rd, g0:RdRd is the drift vector, g=(g1,g2,,gm):RdRd,m is called diffusion matrix, W=(W1,,Wm)T is an m-dimensional Wiener process adapted to the filtration, and X0 is assumed to be Ft0-measurable.

Throughout this paper, we consider discrete time approximations Yn=Y(tn) with equidistant time discretization Ih={t0,t1,,tN} with t0<t1<<tN=T of the time interval I=[t0,T] with step sizes hn=tn+1tn for n=0,1,,N1 and maximal step size hmax=maxn=0,1,,N1hn.

We consider SRK methods (Burrage & Burrage, Citation1996; Debrabant & Kværnø, Citation2008/09; Rößler, Citation2006, Citation2009)

(2) Yn+1=Yn+k=0mμMˉi=1szik,μ;ngk(Hi(k,μ;n)),Hi(k,μ;n)=Yn+l=0mνMˉj=1sZi,jk,μ;l,ν;ngl(Hj(l,ν;n)).(2)

Here, s is the stage number of the SRK method, Mˉ a  finite  set  of multi-indices (e.g., M¯={{j1},{j1,j2}:0j1,j2m}), and the zik,μ;n and Zi,jk,μ;l,ν;n are some random variables depending on the step size hn that are independent of Ftn, n=0,1,,N1.

Generalizing the ideas in Alcock and Burrage (Citation2006), Amiri and Hosseini (Citation2015), Kahl and Schurz (Citation2006); Mardones and Mora (Citation2019), Milstein et al. (Citation1998), Rathinasamy and Nair (Citation2018) and Schurz (Citation2005), a balanced version of EquationEquation (2) is given as

(3) Yn+1=Yn+k=0mμMˉi=1szik,μ;ngk(Hi(k,μ;n))+Cn(YnYn+1),Hi(k,μ;n)=Yn+l=0mνMˉj=1sZi,jk,μ;l,ν;ngl(Hj(l,ν;n)),(3)

where for an underlying SRK method of weak order p using approximations ΔWnj to the Brownian increments Wtn+1jWtnj the matrices Cn are chosen as

(4) Cn=hnp+12c0(tn,Yn)+j=1m((ΔWnj)2phnp1)cj(tn,Yn),(4)

in which the d×d-matrix valued functions cj, j=0,,m, satisfy the following assumption adapted from (Milstein et al., Citation1998). Here, we denote by Ξ the set of possible values for (ΔWnj)2phnp1. For example, if ΔWnj=Wtn+1jWtnj and p=2, then Ξ=[1,), and if p=1 and ΔWnj=±hn, then Ξ={0}.

Assumption 1 For any sequence of real numbers αi with α0[0,αˉ], αiΞ, i=1,,m and αˉhn for all step sizes hn considered and (t,x)[0,]×Rd, the matrix

M(t,x)=Id+α0p+12c0(t,x)+j=1mαjcj(t,x)

has an inverse fulfilling M(t,x)1M<.

The constant M relates to the individual bounds on the norm of the cj matrices.

3 Weak convergence analysis for balanced Runge–Kutta methods

In this section, for κR, let Cb(κ)l(Rd,Rn) be the set of all l-times continuously differentiable functions f:RdRn with uniformly bounded derivatives up to order l, such that for some constant Kf

max{fn,fn×d,2fn×d×d,}Kf(1+xdκ),

where n denotes the Euclidean vector norm on Rn, n×d the Frobenius norm on Rn,d, and in general

An×d1××dk=i0=1ni1=1d1ik=1dkAi0,i1,,ikforARn,d1,,dk.

To reduce the convergence analysis of balanced Runge–Kutta methods to the one of the underlying methods we will first establish the following theorem that generalizes a corresponding statement in Schurz (Citation2005) for the standard Euler method.

Theorem 1 Let Y˜ be a numerical method with one-step representation Y˜[s,y](t) that is weakly consistent of order p, i. e. for fCb(κ)2([t0,T]×Rd,R) there exists a constant KfP depending only on t0,T,f,g0,,gm such that the local weak error fulfills

|E[f(X[u,x](t))f(Y˜[u,x](t))]|KfP(1+xd4κ)(tu)p+1

where X[u,x](t) denotes the solution of stochastic differential Equationequation (1) with initial value X[u,x](u)=x. Let further Y be a numerical method with one-step representation Y[u,y](t), the sequence of approximations Yk defined by Yk+1=Y[tk,Yk](tk+1) and

d[tk,Yk](tk+1)=Y˜[tk,Yk](tk+1)Y[tk,Yk](tk+1)

and assume that for some ρN

E[E[d[tk,Yk](tk+1)|Ftk]d4]4K11+E[Ykd2ρ]4(tk+1tk)p+1,
E[d[tk,Yk](tk+1)d4]K21+E[Ykd2ρ](tk+1tk)p+1.

Assume finally that

maxk=0,,NE[Ykdmax{4κ,2ρ}]K3,maxk=0,,NE[Y˜[tk,Yk](tk+1)d4κ]K4.

Then Y is weakly convergent of order p, i.e. there exists C independent of Ih such that

maxk=0,,N|E[f(X(tk))f(Yk)]|Chmaxp.

Proof The proof follows the one of [22, Theorem 5.2], with a few changes. Suppose that gCb(κ)2(Rd,R) and consider

Mg(k):=E|E[g(Y˜[tk,Yk](tk+1))g(Y[tk,Yk](tk+1))|Ftk]|.

Then we have for some θk,1,θk,2[0,1] that

Mg(k)=E|g(Yk),E[d[tk,Yk](tk+1)|Ftk]d
+E[2g(η2(tk+1))(Yk+1Yk+θk,1d[tk,Yk](tk+1)),d[tk,Yk](tk+1))d|Ftk]|

where η2(tk+1)=Yk+θk,2(Yk+1Yk+θk,1d[tk,Yk](tk+1)) and ,d denotes the Euclidean scalar product in Rd. Applying the Cauchy–Schwarz inequality, we obtain that

Mg(k)=E|g(Yk),E[d[tk,Yk](tk+1)|Ftk]d
+E[θk,12g(η2(tk+1))d[tk,Yk](tk+1),d[tk,Yk](tk+1)d|Ftk]
+E[2g(η2(tk+1))(Yk+1Yk),d[tk,Yk](tk+1)d|Ftk]|
E|g(Yk)dE[d[tk,Yk](tk+1)|Ftk]d
+E[θk,12g(η2(tk+1))d×dd[tk,Yk](tk+1)d2|Ftk]|
E[g(Yk)d2]E[E[d[tk,Yk](tk+1)|Ftk]d2]
+E[2g(η2(tk+1))d×d2]E[d[tk,Yk](tk+1)d4]
[t]E[g(Yk)d4]4E[E[d[tk,Yk](tk+1)|Ftk]d4]4
+E[2g(η2(tk+1))d×d4]4E[d[tk,Yk](tk+1)d4]
23/4Kg1+E[Ykd4κ]4E[E[d[tk,Yk](tk+1)|Ftk]d4]4
+ 23/4Kg1+E[(η2(tk+1))d4κ]4E[d[tk,Yk](tk+1)d4]4
23/4KgK11+E[Ykd4κ]41+E[Ykd2ρ]4(tk+1tk)p+1
+23/4KgK21+3E[Ykd4κ]+3E[Yk+1d4κ]+3E[Y˜[tk,Yk](tk+1)d4κ]4
1+E[Ykd2ρ](tk+1tk)p+1
Cg(tk+1tk)p+1

with Cg:=23/4Kg1+K3K1+K21+6K3+3K44. Let

u(s,x)=E[f(Xs,x(tk+1))].

For the global weak error

ε0(tk+1):=|E[f(X(tk+1))f(Yk+1)]|

it follows then

ε0(tk+1)=|i=0k1E[u(ti+1,Xti,Yi(ti+1))]E[u(ti+1,Yti,Yi(ti+1))]
+E[f(X[tk,Yk](tk+1))]E[f(Y[tk,Yk](tk+1))]|
i=0k1E|E[u(ti+1,Xti,Yi(ti+1))u(ti+1,Y˜ti,Yi(ti+1))|Fti]|
+i=0k1Mu(i)+E|E[f(X[tk,Yk](tk+1))f(Y˜[tk,Yk](tk+1))|Ftk]|
+Mf(k)
(KfP+KuP)(1+K3)+Cg+Cfi=0k(ti+1ti)p+1
(KfP+KuP)(1+K3)+Cg+Cf(Tt0)hmaxp.

Applying this theorem to a balanced Runge–Kutta method and its underlying method, we obtain

Theorem 2 Let Y be given by the balanced SRK method (3), and assume that its underlying Runge–Kutta method (2) is weakly consistent of order p. Assume further that

a) the coefficients gj of stochastic differential Equationequation (1) fulfill a Lipschitz and linear growth condition, i. e. there exists K,LR such that for all x,yRd

(5) gj(x)gj(y)dLxyd,gj(x)dK(1+xd),(5)

and

b) that the coefficients of the balanced method fulfill that cjCb(κ)0(Rd,Rd,d), j=0,,m, and that there exists Kz such that

(6) |zik,ν;n|Kzhn,|Zi,jk,μ;l,ν;n|Kzhn(6)

and

(7) νMˉi=1sE[(Id+Cn)1zik,ν;n]∥≤Kzhn(7)

for i,j=1,,s as well as μ,νMˉ, k,l=0,,m, and n=0,,N1.

Then Y is weakly convergent of order p.

Proof First, note that EquationEquation (3) implies that

Hi(k,μ;n)d≤∥Ynd+l=0mνMˉj=1s|Zi,jk,μ;l,ν;n|gl(Hj(l,ν;n))d.

Using EquationEquations (5) andEquation(6), we obtain

Hi(k,μ;n)d≤∥Ynd+l=0mνMˉj=1sKzhnK(1+Hj(l,ν;n)d)

which implies

(8) maxk,μ,iHi(k,μ;n)dK5(1+Ynd)(8)

for all hn(0,hmax), where

K5=1+(m+1)|Mˉ|sKzKhmax1(m+1)|Mˉ|sKzKhmax,hmax=1(1+(m+1)|Mˉ|sKzK)2.

Here, |Mˉ| denotes the number of elements in Mˉ.

Further, EquationEquation (3) implies that

(9) Yn+1=Yn+(Id+Cn)1k=0mνMˉi=1szik,ν;ngk(Hi(k,ν;n)).(9)

Let

d[tk,Yk](tk+1)=Y˜[tk,Yk](tk+1)Yk+1

where Y˜[tk,Yk](tk+1) denotes the numerical approximation given by the underlying Runge–Kutta method (EquationEquation (2)) of the balanced method at time point tk+1 when starting at time tk with value Yk. It follows then from EquationEquations (2) and Equation(9) that

d[tk,Yk](tk+1)=(Id(Id+Ck)1)=(Id+Ck)1(Id+CkId)l=0mνMˉi=1szil,ν;kgl(Hi(l,ν;k))
=(Id+Ck)1Ckl=0mνMˉi=1szil,ν;kgl(Hi(l,ν;k)),

implying together with EquationEquation (4), Assumption 1, EquationEquations (6),Equation(5) and Equation(8) that

(10) E[E[d[tk,Yk](tk+1)|Ftk]d4]4K11+E[Ykd4(1+κ)]4(tk+1tk)p+1,E[d[tk,Yk](tk+1)d4]K21+E[Ykd4(1+κ)](tk+1tk)p+1.(10)

Similarly, we obtain from EquationEquation (9) by using EquationEquation (4), Assumption 1, EquationEquations (6),Equation(5) and (Equation8)that

(11) Yn+1YndK6(1+Ynd)hn)(11)

with K6=M(m+1)|Mˉ|sKzK(1+K5), and in addition using EquationEquation (7)that

E(Yn+1Yn|Ftn)d≤∥k=0mνMˉi=1sE[(Id+Cn)1zik,ν;n]gk(Yn)d
(12,13) +k=0mνMˉi=1sE[(Id+Cn)1zik,ν;n[gk(Hi(k,ν;n))gk(Yn)]|Ftn]d(12,13)
(14) K7hn(1+Ynd)(14)

with

K7=(m+1)KzK+(m+1)|Mˉ|sMKzL(m+1)|Mˉ|sKzK(1+K5).

With EquationEquations (11) andEquation(12), it follows from [15, Lemma 9.1] that for each rN, there exists a constant K2r such that for all N

(15) maxk=0,,NE[Ykd2r]K2,r,maxk=0,,NE[Y˜[tk,Yk](tk+1)d2r]K2,r.(15)

With EquationEquations (10) and (Equation13), the assertion follows now by Theorem 1.

As a direct consequence, Theorem 2 yields the weak convergence order of the balanced versions of W2Ito1, SRI1, W2Ito2 and SRK4 discussed in Rathinasamy et al. (Citation2020) and Rathinasamy and Nair (Citation2018).

Corollary 1 For sufficiently smooth drift and diffusion, the balanced versions of W2Ito1, SRI1, W2Ito2 and SRK4 are of weak convergence order 2.

Acknowledgements

The authors are grateful to two anonymous referees for their helpful hints which improved the presentation of the material.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been corrected with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

The authors received no direct funding for this research.

References

  • Abdulle, A., Vilmart, G., & Zygalakis, K. C. (2013). Mean-square A-stable diagonally drift-implicit integrators of weak second order for stiff Itô stochastic differential equations. BIT, Vol. 53(4), 827–6. http://dx.doi.org/10.1007/s10543-013-0430-8
  • Alcock, J., & Burrage, K. (2006). A note on the balanced method. BIT Numerical Mathematics, 46(4), 689–710. https://doi.org/10.1007/s10543-006-0098-4
  • Amiri, S., & Hosseini, S. M. (2015). A class of balanced stochastic Runge-Kutta methods for stiff SDE systems. Numerical Algorithms, 69(3), 531–552. https://doi.org/10.1007/s11075-014-9911-3
  • Buckwar, E., Rößler, A., & Winkler, R. (2010). Stochastic Runge-Kutta methods for Itô SODEs with small noise. SIAM J. Sci. Comput, 32(4), 1789–1808. https://doi.org/10.1137/090763275
  • Burrage, K., & Burrage, P. M. (1996). High strong order explicit Runge-Kutta methods for stochastic ordinary differential equations. Applied Numerical Mathematics 22: 81–101. Special issue celebrating the centenary of Runge-Kutta methods https://doi.org/10.1016/S0168-9274-96-00027-X 1–3.
  • Debrabant, K., & Kværnø, A. (2008/09). B-series analysis of stochastic Runge-Kutta methods that use an iterative scheme to compute their internal stage values. SIAM J. Numer. Anal, Vol. 47(1), 181–203. https://doi.org/10.1137/070704307
  • Debrabant, K., & Rößler, A. (2008). Continuous weak approximation for stochastic differential equations. Journal of Applied Mathematics, Vol. 214(1), 259–273. https://doi.org/10.1016/j.cam.2007.02.040
  • Debrabant, K., & Rößler, A. (2009). Families of efficient second order Runge-Kutta methods for the weak approximation of Itô stochastic differential equations. Applied Numerical Mathematics, 59(3–4), 582–594. https://doi.org/10.1016/j.apnum.2008.03.012
  • Higham, D. J. (2000). Mean-square and asymptotic stability of the stochastic theta method. SIAM Journal on Numerical Analysis, 38(3), 753–769 (electronic). https://doi.org/10.1137/S003614299834736X
  • Kahl, C., & Schurz, H. (2006). Balanced Milstein methods for ordinary SDEs. Monte Carlo Methods and Applications, 12(2), 143–170. https://doi.org/10.1163/156939606777488842
  • Kloeden, P. E., & Platen, E. (1999). Numerical solution of stochastic differential equations. In Applications of Mathematics (2 edn), 21 Springer-Verlag. https://doi.org/10.1007/978-3-662-12616-5
  • Mao, X. (2008). Stochastic differential equations and applications (second edn). Horwood Publishing Limited. https://doi.org/10.1533/9780857099402
  • Mardones, H., & Mora, C. (2019). First-order weak balanced schemes for stochastic differential equations. In Methodol. Comput. Appl. Probab. https://doi.org/10.1007/s11009-019-09733-5
  • Milstein, G. N. (1985). Weak approximation of solutions of systems of stochastic differential equations. Theory of Probability & Its Applications, 30(4), 750–766. https://doi.org/10.1137/1130095
  • Milstein, G. N. (1995). Numerical integration of stochastic differential equations. In Mathematics and its Applications. Vol. 313. Kluwer Academic Publishers Group:. Translated and revised from the 1988 Russian original. https://doi.org/10.1007/978-94-015-8455-5
  • Milstein, G. N., Platen, E., & Schurz, H. (1998). Balanced implicit methods for stiff stochastic systems. SIAM Journal on Numerical Analysis, 35(3), 1010–1019 (electronic). https://doi.org/10.1137/S0036142994273525
  • Milstein, G. N., & Tretyakov, M. V. (2004). Stochastic numerics for mathematical physics. Scientific Computation. Springer. ixx. https://doi.org/10.1007/978-3-662-10063-9
  • Rathinasamy, A., Ahmadian, D., & Nair, P. (2020). Second-order balanced stochastic Runge-Kutta methods with multi-dimensional studies. J. Comput. Appl. Math, Vol. 377(112890), 23. https://doi.org/10.1016/j.cam.2020.112890
  • Rathinasamy, A., & Nair, P. (2018). Asymptotic mean-square stability of weak second-order balanced stochastic Runge-Kutta methods for multi-dimensional Itô stochastic differential systems. Applied Mathematics and Computing, 332, 276–303. https://doi.org/10.1016/j.amc.2018.03.065
  • Rößler, A. (2006). Rooted tree analysis for order conditions of stochastic Runge–Kutta methods for the weak approximation of stochastic differential equations. Stochastic Analysis and Applications, 24(1), 97–134. https://doi.org/10.1080/07362990500397699
  • Rößler, A. (2009). Second order Runge–Kutta methods for Itô stochastic differential equations. SIAM J. Numer. Anal, Vol. 47(3), 1713–1738 (electronic). https://doi.org/10.1137/060673308
  • Schurz, H. (2005). Convergence and stability of balanced implicit methods for systems of SDEs. International Journal of Numerical Analysis & Modeling, 2(2), 197–220.
  • Tang, X., & Xiao, A. (2017). Efficient weak second-order stochastic Runge-Kutta methods for Itô stochastic differential equations. BIT Numerical Mathematics, 57(1), 241–260. https://doi.org/10.1007/s10543-016-0618-9
  • Tang, X., & Xiao, A. (2018). Efficient stochastic Runge-Kutta methods for stochastic differential equations with small noises. Advances in Applied Mathematics and Mechanics, 10(4), 845–878. https://doi.org/10.4208/aamm.oa-2017-0181
  • Tang, X., & Xiao, A. (2019). New explicit stabilized stochastic Runge-Kutta methods with weak second order for stiff Itô stochastic differential equations. Numerical Algorithms, 82(2), 593–604. https://doi.org/10.1007/s11075-018-0615-y
  • Tocino, Á., & Vigo-Aguiar, J. (2002). Weak second order conditions for stochastic Runge–Kutta methods. SIAM Journal of Scientific Computing, Vol. 24(2), 507–523 (electronic). https://doi.org/10.1137/S1064827501387814