1,713
Views
3
CrossRef citations to date
0
Altmetric
Original Articles

The COS method for option valuation under the SABR dynamics

&
Pages 444-464 | Received 24 Feb 2016, Accepted 28 Sep 2016, Published online: 28 Feb 2017

ABSTRACT

In this paper, we consider the COS method for pricing European and Bermudan options under the stochastic alpha beta rho (SABR) model. In the COS pricing method, we make use of the characteristic function of the discrete forward process. We observe second-order convergence by using a second-order Taylor scheme in the discretization, or by using Richardson extrapolation in combination with a Euler–Maruyama discretization on the forward process. We also consider backward stochastic differential equations under the SABR model, using the discretized forward process and Fourier-cosine expansion for the occurring expectations. For this purpose, we extend the so-called BCOS method from one to two dimensions.

2010 AMS Subject Classifications:

1. Introduction

Efficient valuation of financial options is an important issue in financial mathematics. In the last few decades, many pricing methods have been introduced, for example, convenient formulas [Citation3,Citation17,Citation19], Monte Carlo methods [Citation5,Citation15,Citation20], finite difference methods [Citation10,Citation18,Citation22], quadrature methods [Citation7] and also Fourier methods [Citation6,Citation12,Citation25,Citation26,Citation28,Citation33]. A variety of option pricing methods is compared in [Citation37], including several Fourier methods. This type of methods, such as the COS method [Citation12,Citation33], is based on the characteristic function (ChF) of the underlying process to determine the option value. However, often no analytic expression for the ChF of the underlying process is available. The authors in [Citation35] use the ChF of the discrete process to price options and to solve backward stochastic differential equations (BSDEs). We extend this so-called BCOS method (BSDE COS method) from one to two dimensions, see also [Citation36], in order to solve BSDEs under the so-called stochastic alpha beta rho (SABR) SDE system. The BCOS method approximates the solutions of BSDEs by iterating backwards in time, so we do not need to simulate the forward process. The theory of BSDEs is described in detail in the literature [Citation4,Citation11,Citation29,Citation30] and relevant applications in mathematical finance are studied in for example [Citation23,Citation34,Citation38]. In this paper, we pay attention especially to the valuation of options with the BCOS method. The basis is formed by using the COS method under SABR dynamics, and we will also explain this. As we will observe, the BCOS method may suffer from the curse of dimensionality, as well as quadrature methods and finite difference methods. Nevertheless, these methods are widely used for two-dimensional models. Therefore, investigation of the two-dimensional BCOS method is of value to acquire a wider diversity in methods.

To obtain rapid convergence, we will employ Richardson extrapolation in combination with the Euler–Maruyama SDE discretization (see also [Citation21]). In particular, we price European and Bermudan options under a two-dimensional model, like the SABR model [Citation17]. Due to the closed-form expression of the implied volatility, the Hagan formula [Citation17], the SABR model is a widely used stochastic local volatility model in the financial world. Many pricing methods for SABR-like models have been presented in the literature, like [Citation2,Citation9,Citation16,Citation18,Citation27], however, many of those methods are mainly accurate for valuing European options with short maturities under the SABR model. In [Citation1] a pricing method is given, which is accurate for options with long maturities, but this method is not applicable for all parameter sets [Citation1]. When we work on the risk neutral measure, we can simply use the COS method, for European and Bermudan options. Under the real-world measure, we encounter a BSDE, where the terminal condition is given by the payoff function.

This paper is organized as follows. In Section 2, we introduce the two-dimensional COS method for pricing European and Bermudan options. We discuss the two-dimensional BCOS method in Section 3. Furthermore, we present several numerical experiments in Section 4. Finally, we conclude in Section 5.

2. The two-dimensional COS method

In this section, we explain the two-dimensional COS method for pricing European and Bermudan options. We assume that the underlying forward stochastic differential equations (FSDEs) can be written in the following general form: (1) dXt1=μ1(Xt1,Xt2)dt+σ1(Xt1,Xt2)dWt1,dXt2=μ2(Xt1,Xt2)dt+ρσ2(Xt1,Xt2)dWt1+1ρ2σ2(Xt1,Xt2)dWt2,(1) where t0, X01=x1, X02=x2, the correlation ρ(1,1), and Wt1 and Wt2 are independent standard Brownian motions on a filtered probability space F. The functions σ1,σ2,μ1,μ2:R2R are assumed to be twice differentiable with respect to x1 and x2. When the corresponding bivariate ChF cannot easily be derived, it is possible to approximate it by the bivariate ChF of the discretized FSDE process. For this discretization, we use the Euler–Maruyama scheme and also the 2.0-weak-Taylor scheme.

We are interested in pricing options under the SABR model [Citation17]. The FSDEs of the SABR model are of the same form as Formula (Equation1), that is, (2) dFt=αt(Ft)βdWt1,F0=f,dαt=ρναtdWt1+1ρ2ναtdWt2,α0=α,(2) where Ft is the forward, for example, the forward swap rate, and αt denotes the volatility. Furthermore, W1 and W2 are independent standard Brownian motions under the forward measure, exponent 0β1, the volatility of the volatility ν0 and the initial forward f and volatility α are non-negative. The authors in [Citation17] proposed a very convenient formula to calculate the Black implied volatility under the SABR model. This formula, also known as the Hagan formula, may however sometimes lead to arbitrage possibilities for low strikes, as one can observe by examining the corresponding PDF [Citation36]. The Hagan formula is not accurate for pricing options with long maturities [Citation1,Citation18,Citation36].

As no analytic expression for the bivariate ChF of the SABR model is available, it may be difficult to calculate accurate reference values for our purposes. That is why we also consider the Heston modelFootnote1 as a reference model for convergence purposes. The ChF of the Heston model is known, which ensures the availability of accurate reference prices for European and Bermudan options. The FSDEs can be written in the form of Formula (Equation1), that is, (3) dXt=12Atdt+AtdWt1,X0=x,dAt=ρνAtdWt1+1ρ2νAtdWt2,A0=a,(3) where Xt denotes the log-asset process and At is the variance. Furthermore, W1 and W2 are independent standard Brownian motions, the volatility of the volatility parameter ν and the initial variance a are both non-negative. FSDEs (Equation3) do not obey the Feller condition, so we have to make sure that we choose the computational domain, see Appendix 2, such that the variance process can attain any non-negative value. The Heston model (Equation3) is an affine model, so an analytic formula for its bivariate ChF is known [Citation36]: (4) φ(Xt+Δt,At+Δt)(u1,u2|Xt,At)=expiu1Xt+1ν2ζtanarctaniu2ν2+iu1ρνζ+Δt2ζiu1ρνAt,(4) where (5) ζ=iu1ν2+u12(ρ21)ν2.(5)

In Section 2.1, we give the Taylor schemes and the necessary formulas to price European options with the COS method. Also, we detail the procedure for valuing Bermudan options in this section. We provide an approach to price multiple European options under the SABR model in one computation in Section 2.2. We describe the use of Richardson extrapolation in Section 2.3. The computational complexity of the proposed method is discussed in Section 2.4.

2.1. Pricing options with the COS method

Just as in [Citation36], we define a time-grid tm=mΔt for m=0,1,,M, with fixed time steps Δt=T/M. We write Xm=Xtm, Xmj=Xtmj, j=1,2. The discrete forward process is denoted by XmΔ=XtmΔ and Xmj,Δ=Xtmj,Δ. To determine the values of Xm+1j,Δ, m=0,,M1, and given Xm=x=(x1,x2), we use the Euler–Maruyama and the 2.0-weak-Taylor schemes. We write the discrete schemes for Formula (Equation1) in general form, as follows: (6) Xm+1j,Δ=xj+mj(x)Δt+sj1(x)ΔWm+11+sj2(x)ΔWm+12+κj1,2(x)ΔWm+11ΔWm+12+κj1(x)(ΔWm+11)2+κj2(x)(ΔWm+12)2+ψj(x)Ψm+1,(6) where ΔWm+11 and ΔWm+12 are uncorrelated and normally distributed Wiener increments with mean zero and variance Δt, and Ψm+1 is an independent random variable with probability P(Ψm+1=±Δt)=12.

For the Euler–Maruyama scheme,Footnote2 we have (7) m1(x)=μ1(x),s11(x)=σ1(x),s12(x)=0,κ11,2(x)=0,κ11(x)=0,κ12(x)=0,ψ1(x)=0,m2(x)=μ2(x),s21(x)=ρσ2(x),s22(x)=1ρ2σ2(x),κ21,2(x)=0,κ21(x)=0,κ22(x)=0,ψ2(x)=0.(7)

The functions associated with the 2.0-weak-Taylor scheme are given in Appendix 1.

According to [Citation24], the definitions of the order of strong convergence and the order of weak convergence are:

Definition 2.1.

The discrete process XΔ converges in the strong sense with order γ1(0,] if there exists a finite constant C and a positive constant δ such that (8) E[|XMXMΔ|]C(Δt)γ1,(8) for any time discretization with maximum step size Δt(0,δ).

Definition 2.2.

Discrete process XΔ converges in the weak sense with order γ2(0,] if for any polynomial h there exists a finite constant C and a positive constant δ such that (9) |E[h(XM)]E[h(XMΔ)]|C(Δt)γ2,(9) for any time discretization with maximum step size Δt(0,δ).

For the Euler scheme, the order of strong convergence is γ1=0.5 and the order of weak convergence is γ2=1, while for the 2.0-weak-Taylor scheme γ2=1 and γ10.5. The strong order of convergence of the 2.0-weak-Taylor scheme equals 1 when Equation (Equation1) satisfies the following commutativity conditions [Citation24]: (10) 1ρ2σ2(x)σ1(x)x2=0and1ρ2σ1(x)σ2(x)x1=0x1,x2R2.(10)

As observed in [Citation36], the convergence of the COS method for pricing European options only depends on the order of weak convergence of the discretization scheme, and not on its order of strong convergence, because European options are not path-dependent. In Example 4.2 we observe that even for Bermudan options the order of weak convergence is leading for the rate of convergence. That is why we can use a weak-Taylor scheme for the discretization. Also, this implies that for application of the COS method it does not matter whether the FSDEs in Equation (Equation1) satisfy the commutativity condition (Equation10) or not.

Theorem 2.1.

The analytic expression for the bivariate ChF of Xm+1Δ, as in Equation (Equation6), is given by (11) φXm+1Δ(u1,u2|x)=cosh(ic6Δt)exp(iu1[x1+m1(x)Δt]+iu2[x2+m2(x)Δt])(12ic4Δt)(12ic5Δt)+c32(Δt)2expΔt2c12+c22+[4(c22c42+c12c52)4c1c2c3(c4+c5)+(c12+c22)c32](Δt)21+(2c32+4c42+4c52)(Δt)2+(c324c4c5)2(Δt)4expi(Δt)2c12c4c22c5c1c2c3+(c324c4c5)(c12c5c1c2c3+c22c4)(Δt)21+(2c32+4c42+4c52)(Δt)2+(c324c4c5)2(Δt)4,(11) where XmΔ=x=(x1,x2) and (12) c1=u1s11(x)+u2s21(x),c4=u1κ11(x)+u2κ21(x),c2=u1s12(x)+u2s22(x),c5=u1κ12(x)+u2κ22(x),c3=u1κ11,2(x)+u2κ21,2(x),c6=u1ψ1(x)+u2ψ2(x).(12)

Proof.

The ChF of Xm+1Δ, given XmΔ=x based on Equation (Equation6), is given by (13) φXm+1Δ(u1,u2|x)=E[exp(iu1Xm+11,Δ+iu2Xm+12,Δ)|XmΔ=x]=exp(iu1[x1+m1(x)Δt]+iu2[x2+m2(x)Δt])E[exp(i(c1v+c2w+c3vw+c4v2+c5w2))]E[exp(ic6Ψ)]=cosh(ic6Δt)exp(iu1[x1+m1(x)Δt]+iu2[x2+m2(x)Δt])E[exp(i(c1v+c2w+c3vw+c4v2+c5w2))],(13) where we abbreviated Ψ=Ψm+1, v=ΔWm+11 and w=ΔWm+12.

The ChF of the non-central chi-squared distribution with one degree of freedom and non-centrality parameter λ reads (14) φχ12(λ)(u)=expiλu12iu112iu.(14)

We assume c40 and c50, which gives (15) φXm+1Δ(u1,u2|x)=cosh(ic6Δt)exp(iu1[x1+m1(x)Δt]+iu2[x2+m2(x)Δt])R2expic2w+c5w2+c4v+c1+c3w2c42(c1+c3w)24c4expv2+w22Δtdvdw=cosh(ic6Δt)2πΔtexp(iu1[x1+m1(x)Δt]+iu2[x2+m2(x)Δt])expic2w+c5w2(c1+c3w)24c4expw22Δtφχ12((c1+c3w)2/4c42Δt)(c4Δt)dw=coshic6Δtexpiu1x1+m1(x)Δt+iu2x2+m2(x)Δt2πΔt(12ic4Δt)exp(12+ic4)c12(Δt)21+4c42(Δt)2+μ22σ2exp(wμ)22σ2expic22c1c3c4(Δt)21+4c42(Δt)2w+c5c32c4(Δt)21+4c42(Δt)2w2dw,(15) where (16) σ2=Δt(1+4c42(Δt)2)1+4c42(Δt)2+(Δt)2c32andμ=σ2c1c3Δt1+4c42(Δt)2.(16) We abbreviate (17) c7=c22c1c3c4(Δt)21+4c42(Δt)2andc8=c5c32c4(Δt)21+4c42(Δt)2.(17) So, the integral is given by (18) exp(ic7w+ic8w2)exp(wμ)22σ2dw=σ2πE[exp(ic7[σW+μ]+ic8[σW+μ]2)],(18) where WN(0,1). Rewriting gives us (19) E[exp(i(c7[σW+μ]+c8[σW+μ]2))]=expic7μ+c8μ2(c7+2c8μ)24c8Eexpic8σ2W+c7+2c8μ2c8σ2=expic7μ+c8μ2(c7+2c8μ)24c8φχ12(((c7+2c8μ)/2c8σ)2)(c8σ2)=expic7μ+c8μ2(c7+2c8μ)24c8expi(c7+2c8μ)24c8(12ic8σ2)112iσ2c8.(19) Substituting result (Equation19) in Equation (Equation15) gives the ChF of Xm+1Δ, given XmΔ=x, that is, (20) φXm+1Δ(u1,u2|x)=exp(i(u1[x1+m1(x)Δt]+u2[x2+m2(x)Δt]+c7μ+c8μ2))Δt(12ic4Δt)(12iσ2c8)σcosh(ic6Δt)expc12Δt2(1+4c42(Δt)2)+μ22σ2(c7+2c8μ)2σ22(1+4c82σ4)expic4c12(Δt)21+4c42(Δt)2(c7+2c8μ)2c8σ41+4c82σ4.(20) This results in ChF (Equation11) when we substitute Equations (Equation16) and (Equation17) into Equation (Equation20) and rewrite it. In a similar way, we find that Equation (Equation11) is valid for c4=0 and/or c5=0 as well.

2.1.1. Pricing European and Bermudan options

We wish to derive the value VC(t0=0,X0) of a European call option with as underlying Xt and exercise date T, where the FSDEs of Xt=(Xt1,Xt2) are given by Equation (Equation1). The payoff of the option at time T is given by VC(T,XT)=g(XT1) for some function g. The value of the option at time t is given by (21) VC(t,Xt)=er(Tt)E[g(XT1)|Xt],(21) where r is the risk-free interest rate and the expectation is taken under the risk-free measure Q. We will denote VC(tm,x) by Vm(x) and the discretized option value by VmΔ(x). For m=M1,,0, we find (22) VmΔ(x)=exp(rΔt)Em[Vm+1Δ(Xm+1Δ)],(22) where XmΔ=x.

A European option can only be exercised at the expiration date T, while a Bermudan option can be exercised at several predetermined dates. There is a minor change of procedure for pricing Bermudan options. Let n be the number of early-exercise dates and let τj denotes the early-exercise dates for j=1,2,...,n, where 0τ1<τ2<<τn=T. We choose the number of time steps M such that each of the early-exercise dates corresponds to a point in our time-grid.Footnote3 We replace Formula (Equation22) by (23) VmΔ(x)=max{g(x1),V~mΔ(x)}for tm=τj,V~mΔ(x)for tmτj,(23) where (24) V~mΔ(x)=exp(rΔt)Em[Vm+1Δ(Xm+1Δ)](24) and XmΔ=x.

In a similar way, we can also price discretely monitored barrier options, see [Citation13,Citation36].

Formula (Equation22), or for a Bermudan option Formula (Equation24), can be recovered recursively, backwards in time, on a two-dimensional grid by means of the COS methodFootnote4 [Citation33], that is, (25) VmΔ(x)k1=0N1k2=0N112φXm+1Δk1πb1a1,k2πb2a2xexpik1πa1b1a1ik2πa2b2a2+φXm+1Δk1πb1a1,k2πb2a2xexpik1πa1b1a1+ik2πa2b2a2Vk1,k2,(25) where indicates that the first term in the series summation is weighted by one-half and (26) Vk1,k2=2b1a12b2a2a2b2a1b1Vm+1Δ(X)cosk1πX1a1b1a1cosk2πX2a2b2a2dX1dX2.(26)

When the above double integral cannot be computed analytically, we compute the function on an x-grid by using the two-dimensional discrete Fourier-cosine transform. For each example in this paper, we evaluate the function VmΔ(x) on a regular grid with N2 grid-points on the domain [a1,b1]×[a2,b2]. In Appendix 2, we explain how to choose these domain boundaries. We proved first-order convergence for the Euler scheme and second-order convergence for the 2.0-weak-Taylor scheme in the error analysis given in [Citation36]. So, we cannot price a European option in only one time step anymore when we work with the ChF of the discrete process.

2.2. Multiple strikes for the SABR model

It is possible to price European options under the SABR model for multiple strikes at once, by using a unique scaling symmetry of the SABR model [Citation31]. The forward value of a European call option under the SABR model with strike value K>0 and time to maturity T is given by (27) VC(t0=0,K,x)=E[(exp(XT1)K)+|x],(27) where Xt1 and Xt2 denote, respectively, the log forward log(Ft) and the log volatility log(αt). In Formula (Equation27), we added the strike K to the notation VC(0,K,x), because we use a general strike K0 in this analysis and it is important to have a clear distinction between those strike values. The corresponding FSDEs, written in the form of Formula (Equation1), are given by (28) dXt1=12exp(2Xt2+2(β1)Xt1)dt+exp(Xt2+(β1)Xt1)dWt1,X01=x1=log(f),dXt2=12ν2dt+ρνdWt1+1ρ2νdWt2,X02=x2=log(α),(28)

To observe the scaling symmetry, we use the following transformations: (29) Xˆt1=logK0K+Xt1andXˆt2=(1β)logK0K+Xt2,(29) and for any K0R>0, it holds that: (30) VC(0,K,x)=E[(exp(XT1)K)+|x]=EKK0exp(XˆT1)K+x=KK0E[(exp(XˆT1)K0)+|x],(30) so this is the scaled price of a call option under the transformed dynamics. Also, it is easy to observe that dXˆt1=dXt1 and dXˆt2=dXt2. This implies that exp(Xˆt1) and exp(Xˆt2) form a SABR process, with the same parameters β, ν and ρ as exp(Xt1) and exp(Xt2), but with different initial values. This observation, together with Equation (Equation30), implies that: (31) VC(0,K,x)=KK0VC0,K0,logK0K+x1,(1β)logK0K+x2.(31) We are thus able to price European options under the SABR model for multiple strikes in one computation by using Formula (Equation31), that is, we price the corresponding European option for one general strike K0 for different initial values. This hardly costs more CPU time, because evaluating VmΔ(x) backwards in time on the x-grid is the time-consuming part of the pricing method as we will also mention in Section 2.4. Such a strike K0 can be, for example, the ATM strike value. For most models such a transformation does not exist.

2.3. Richardson extrapolation

The use of the 2.0-weak-Taylor SDE scheme in combination with the COS method results in second-order convergence, whereas the Euler scheme results in only first-order convergence [Citation36]. Using this second-order scheme is however computationally expensive in two dimensions as we will observe in Example 4.1. The number of terms to be computed grows significantly in the ChF of the 2.0-weak-Taylor scheme (Equation11). This disadvantage of the 2.0-weak-Taylor scheme led us to the use of Richardson extrapolation on the results of the Euler scheme, just as the authors in [Citation21] and many others did, to obtain second-order convergence.

Suppose V0C=VC(0,X0) is the exact option value and vˆ(M1) and vˆ(M2) denote the approximations by the COS method for the Euler scheme to this value, where the number of time steps is respectively given by M1 and M2. We choose M1 and M2 such that M1<M2. The Euler scheme exhibits first-order convergence. When this convergence is smooth and monotone, we obtain the following approximations: (32) vˆ(M1)=V0C+ε1M1+O1M12,(32) (33) vˆ(M2)=V0C+ε1M2+O1M22.(33) Rearranging these equations gives (34) V0C=M2vˆ(M2)M1vˆ(M1)M2M1+O1(M2)2.(34) Formula (Equation34) implies that combining the Euler results for M1 and M2 results in a second-order method if and only if the convergence is smooth and monotoneFootnote5. In this paper, we will always use 2M1=M2.

Remark 2.1.

Theoretically, we can apply Richardson extrapolation on the Euler or 2.0-weak-Taylor results to achieve a rate of convergence of three or more. In practice, we have to make a trade-off between the order of convergence and the computational costs of the method. To apply Richardson extrapolation, the spatial discretization error should be relatively small with respect to the temporal discretization error. Because of the computational complexity of the method, as we will discuss in Section 2.4, performing more time steps with second-order convergence to achieve a certain accuracy may have lower computational costs than using more grid-points to try to achieve third-order convergence.

2.4. Computational complexity

In this section, we discuss the computational complexity of our algorithm. As explained in Section 2.1, we evaluate the function VmΔ(x) backwards in time on a regular grid with N2 grid-points on the domain [a1,b1]×[a2,b2], for m=M1,,0. At each time step, we use the two-dimensional discrete Fourier-cosine transform to approximate the N2 terms in Formula (Equation26), which is of order O(N2log(N2))=O(N2log(N)). Evaluating the function VmΔ(x) on one grid-point costs O(N2) operations, which is higher than for Lévy processes, because we do not have the benefits of the decomposition into a Hankel and Toeplitz matrix [Citation33]. VmΔ(x) is explicit for each grid-point, so it is possible to evaluate this function for each grid-point parallel at a certain time step. At m=0, we only have to evaluate V0Δ(x) on the initial value x=X0 which is of order O(N2). So, the total number of operations is of the order: (35) O((M1)N4)+O(N2)+O(MN2log(N)).(35) Despite the fact that we can compute many strike values in one computation, these computational costs are a disadvantage of the method.

3. The two-dimensional BCOS method

In Section 2, we considered the COS method for pricing European and Bermudan options under the SABR model. In this section, we generalize a method to solve BSDEs under the SABR model with the COS method on the basis of the ChF of the discrete scheme, the two-dimensional BCOS method. This method is an extension of the one-dimensional BCOS method [Citation34,Citation35].

In Section 3.1, we introduce the 2D BCOS method and we discuss the computational complexity of this method in Section 3.2.

3.1. The method

We wish to derive the value V(t0=0,X0) with as underlying asset process Xt=(Xt1,Xt2) and exercise date T, where the FSDEs of Xt are given by (Equation1). The payoff of the derivative at time T is given by V(T,XT)=g(XT) for some function g. We also assume that we are working in a complete market,Footnote6 that is, a replicating portfolio can be created and the market is frictionless. We make a self-financing portfolio Yt consisting of at1 assets of Xt1, at2 assets of Xt2 and bonds with risk-free return rate r, such that YT=g(XT), and: (36) dYt=r(Ytat1Xt1at2Xt2)dt+at1dXt1+at2dXt2=[rYt+(μ1(Xt)rXt1)at1+(μ2(Xt)rXt2)at2]dt+[σ1(Xt)at1+ρσ2(Xt)at2]dWt1+1ρ2σ2(Xt)at2dWt2,(36) for 0tT. If we set Zt1=σ1(Xt)at1 and Zt2=σ2(Xt)at2, then (Y,Z) solves the BSDE: (37) dYt=f(t,Xt,Yt,Zt)dt+(Zt1+ρZt2)dWt1+1ρ2Zt2dWt2,(37) (38) f(t,x,y,z)= ryμ1(x)rx1σ1(x)z1μ2(x)rx2σ2(x)z2,(38) where YT=g(XT). Yt is a self-financing portfolio, so the value of the derivative is given by V(0,X0)=Y0. The function f:[0,T]×R5R is assumed to be uniformly continuous with respect to x1 and x2 and satisfies a Lipschitz condition in (y,z), with Lipschitz constant Lf and the function g:R2R is assumed to be uniformly continuous with respect to x1 and x2. Details on conditions of the existence and uniqueness of solution (Y,Z) can be found in [Citation29,Citation30].

Just as in Section 2, we define a time-grid of M+1 time points. We also define Λt:=(Xt,Yt,Zt), Λm=Λtm and ΛmΔ=(XmΔ,YmΔ,ZmΔ). Integrating Equation (Equation37) gives us: (39) Y0=g(XT)+0Tf(t,Λt)dt0T(Zt1+ρZt2)dWt11ρ20TZt2dWt2.(39) At time tm, this gives the recursion: (40) Ym=Ym+1+tmtm+1f(t,Λt)dttmtm+1(Zt1+ρZt2)dWt11ρ2tmtm+1Zt2dWt2.(40) We take conditional expectations at both sides of the equation and use numerical integration to approximate the integral, for some θ1[0,1], we find: (41) Ym=Em[Ym+1]+Emtmtm+1f(t,Λt)dt(41) (42) Em[Ym+1]+Em[Δtθ1f(tm,Λm)+Δt(1θ1)f(tm+1,Λm+1)]=Em[Ym+1]+Δtθ1f(tm,Λm)+Δt(1θ1)Em[f(tm+1,Λm+1)].(42) Multiplying Equation (Equation40) with ΔWm+11 gives (43) YmΔWm+11=Ym+1ΔWm+11+tmtm+1f(t,Λt)dtΔWm+11tmtm+1(Zt1+ρZt2)dWt1ΔWm+111ρ2tmtm+1Zt2dWt2ΔWm+11.(43) Again, we take conditional expectations at both sides of the equation and use numerical integration with θ2[0,1], giving: (44) 0Em[Ym+1ΔWm+11]+Em[(Δtθ2f(tm,Λm)+Δt(1θ2)f(tm+1,Λm+1))ΔWm+11]Em[(θ2ΔWm+11(Zm1+ρZm2)+(1θ2)ΔWm+11(Zm+11+ρZm+12))ΔWm+11]1ρ2Em[θ2ΔWm+12(Zm2+Zm+12)ΔWm+11]=Em[Ym+1ΔWm+11]+Δt(1θ2)Em[f(tm+1,Λm+1)ΔWm+11]Δtθ2(Zm1+ρZm2)Δt(1θ2)Em[Zm+11+ρZm+12].(44)

Analogously, we find (45) 0Em[Ym+1ΔWm+12]+Δt(1θ2)Em[f(tm+1,Λm+1)ΔWm+12]Δtθ21ρ2Zm2Δt(1θ2)1ρ2Em[Zm+12].(45)

We use one of the approximation schemes of Section 2 in Formulas (Equation42), (Equation44) and (Equation45), which results, for m=M1,,0, in the following formulas: (46) YmΔ(x)=Em[Ym+1Δ(Xm+1Δ)]+Δtθ1f(tm,ΛmΔ(x))+Δt(1θ1)Em[f(tm+1,Λm+1Δ(Xm+1Δ))],(46) (47) Zm1,Δ(x)=1Δtθ2Em[Ym+1Δ(Xm+1Δ)ΔWm+11]1θ2θ2Em[Zm+11,Δ(Xm+1Δ)+ρZm+12,Δ(Xm+1Δ)]+1θ2θ2Em[f(tm+1,Λm+1Δ(Xm+1Δ))ΔWm+11]ρZm2,Δ(x),(47) (48) Zm2,Δ(x)=1Δtθ21ρ2Em[Ym+1Δ(Xm+1Δ)ΔWm+12]1θ2θ2Em[Zm+12,Δ(Xm+1Δ)]+1θ2θ21ρ2Em[f(tm+1,Λm+1Δ(Xm+1Δ))ΔWm+12],(48) where XmΔ=x. YmΔ(x) is implicit for θ1>0 and can be determined by performing P Picard iterations, starting with initial guess Em[Ym+1Δ(Xm+1Δ)]. We give the approximations of the conditional expectations in Formulas (Equation46)–(Equation48) with the COS method in Appendix 3. The value of the derivative can be approximated by V(0,X0)=Y0Δ(X0).

We expect to obtain first-order convergence for the Euler scheme, independent of the choice for θ1 and θ2. This means that we can avoid the use of Picard iterations for the Euler scheme by taking θ1=0. Second-order convergence can be achieved when we apply Richardson extrapolation on the results of the Euler scheme, in the way we explained in Section 2.3. This is not only possible for the Y -process, but also for both Z-processes. We obtain second-order convergence if and only if the Euler results are converging smoothly and monotonically and this holds when θ2=1 [Citation21]. Also, we expect to obtain second-order convergence for the 2.0-weak-Taylor scheme, when we choose the second-order mid-point method as the numerical integration method in Formulas (Equation42), (Equation44) and (Equation45), so θ1=θ2=12.

3.2. Computational complexity

In this section, we discuss the computational complexity of the BCOS method. This method is more expensive than the COS method of Section 2. These extra costs are caused by the two additional processes Zt1 and Zt2 and also by the use of Picard iterations in the implicit formula (Equation46). We evaluate the functions YmΔ(x), Zm1,Δ(x) and Zm2,Δ(x) backwards in time on a regular grid with N2 grid-points on the domain [a1,b1]×[a2,b2], for m=M1,,0. At each time step, we use the two-dimensional discrete Fourier-cosine transform to approximate the N2 cosine coefficients of each process. Also, we do P Picard iterations at each time step. At m=0, the costs of these iterations are of order O(P), and for m=M1,,1 these costs are of order O(PN2) for each time step. So, the total number of operations is of the order: (49) O(3(M1)N4)+O(3N2)+O(3MN2log(N))+O(P((M1)N2+1)).(49)

4. Numerical examples

In this section, we present some numerical experiments. We price European and Bermudan options in the examples of Section 4.1 and several BSDE examples are given in Section 4.2. We used MATLAB R2015a for the coding and the computations are performed on Intel(R) Core(TM) i5-2400 3.10 GHz processor with 16 GB RAM.

4.1. European and Bermudan examples

In the first example, we price multiple European call options under the uncorrelated SABR model. We consider the correlated Heston model in the second example, where we value not only call options, but also a Bermudan put option.

Example 4.1.

For the SABR dynamics (Equation28), we consider initial log forward x1=log(2), initial log volatility x2=log(0.35), exponent β=0.8, volatility of volatility parameter ν=0.4, correlation ρ=0, strikes K=2.55,2.7,2.85, time to expiration T=1 and number of grid-points and cosine coefficients in each dimension N=27. Figure  shows the mean of the absolute error in the Black implied volatility in basis points (BPS) for the Euler scheme (with and without Richardson extrapolation) and the 2.0-weak-Taylor scheme with the COS method, where we use Formula (Equation31).Footnote7 We determined our reference values by using the method given in [Citation1], which gives for each option a reference Black implied volatility with an error of order O(101) in BPS, for ρ=0.

Figure 1. SABR example.

Figure 1. SABR example.

As expected, the 2.0-weak-Taylor scheme shows faster convergence than the Euler scheme, but not in terms of CPU time as we show in Table . This is the main disadvantage of using the 2.0-weak-Taylor scheme, and the reason why we use Richardson extrapolation on the Euler scheme which also results in a second-order method. With Richardson extrapolation, we only need two additional time steps to observe an accuracy of less than 1 BPS, compared to 2.0-weak-Taylor scheme, as is shown in Figure  . The use of Richardson extrapolation leads to a significant reduction in CPU time.

Table 1. Number of time steps and CPU time needed to obtain an absolute error of 1 or 0.5 basis points in the Black implied volatility for the Euler (with and without Richardson extrapolation) and 2.0-weak-Taylor schemes.

Example 4.2.

In this example, we consider the Heston model with different values for the correlation ρ. We wish to derive the value of a European call option with initial log-asset price x=log(2), initial variance a=0.2, vol–vol ν=0.3, correlations ρ=0.8,0.4,0,0.4,0.8, strike K=1.9 and time to expiration T=0.1. The reference value is obtained by using the COS method, with the ChF of the continuous process, which is given in Formula (Equation4), and with 214 cosine coefficients. The convergence of the COS method with the Euler scheme is shown in Figure (a) and we observe for all values ρ first-order convergence.

Figure 2. The Heston model with different correlations: (a) the Euler–Maruyama scheme and (b) Richardson extrapolation on the Euler results.

Figure 2. The Heston model with different correlations: (a) the Euler–Maruyama scheme and (b) Richardson extrapolation on the Euler results.

Also, we apply Richardson extrapolation on these Euler scheme results. The results are given in Figure (b) and we observe second-order convergence for all ρ values, confirming the accuracy of Richardson extrapolation.

We continue by considering early-exercise dates for this problem. We take correlation ρ=0.2 and we wish to derive the value of a Bermudan put option under the Heston model with strike price K=1.9 and 5 early-exercise dates: T=0.02, 0.04, 0.06, 0.08, 0.1. The convergence results are given in Figure . We observe, as expected, first-order convergence for the Euler scheme and second-order convergence with Richardson extrapolation on the Euler results. Pricing Bermudan options under the Heston model with the COS method is also considered in [Citation14,Citation33], and we used the method given in this last paper to obtain the reference values. The difference with our method is that we employed the ChF of the discretized process.

Figure 3. Heston Bermudan put with 5 early-exercise dates.

Figure 3. Heston Bermudan put with 5 early-exercise dates.

4.2. BSDE examples

In the first experiment, we approximate the value of a geometric basket call option and in the second example we consider a spread option. We price a European call option under the SABR model in the last experiment. All forward processes are under the P-measure and we choose θ1=0 and θ2=1 for the Euler scheme in each example, because for these values we obtain smooth, monotonic convergence for the Euler scheme, which is a necessary condition for applying Richardson extrapolation, see also [Citation21]. For the 2.0-weak-Taylor scheme, we take θ1=θ2=12, because a second-order numerical integration method is required to achieve second-order convergence.

Example 4.3.

We determine the value of a geometric basket call option, where both assets follow a geometric Brownian motion (GBM): (50) dSt1=μ1St1dt+σ1St1dWt1,S01=s1,dSt2=μ2St2dt+ρσ2St2dWt1+1ρ2σ2St2dWt2,S02=s2,(50) and the payoff function is g(ST1,ST2)=max(ST1ST2K,0). For the experiment we consider the following parameter values: s1=0.90,s2=1.10,μ1=0.1,μ2=0.1,σ1=0.2,σ2=0.3,ρ=0.25,r=0.04,K=1,T=0.1.

We evaluate the BSDE for pricing and hedging this basket option, such as in Formula (Equation37), that is, (51) dYt=rYt+μ1rσ1Zt1+μ2rσ2Zt2dt+(Zt1+ρZt2)dWt1+1ρ2Zt2dWt2.(51) The reference value for this option is given by the Black–Scholes price of a European call option with initial stock value S0=s1s2=0.91.1, constant dividend yield q=14((σ1)2+(σ2)22σ2)=0.0125, volatility σ=14((σ1)2+(σ2)2+2ρσ1σ2)=0.2 and the same risk-free interest rate, time of maturity and strike price. This gives the reference values Y0=0.023982340, Z01=0.049514739 and Z02=0.074272109. We observe first-order convergence of the BCOS method with the Euler scheme in Figure .

Figure 4. Geometric basket call option with the Euler scheme.

Figure 4. Geometric basket call option with the Euler scheme.

Example 4.4.

In this example we consider a spread option. The payoff function of such an option is given by g(ST1,ST2)=max(ST1ST2K,0). We assume that the underlying assets both follow a constant elasticity of variance (CEV) process: (52) dSt1=μ1St1dt+α1(St1)βdWt1,S01=s1,dSt2=μ2St2dt+ρα2(St2)γdWt1+1ρ2α2(St2)γdWt2,S02=s2,(52) For this example, we consider the following parameter values: s1=2.1,s2=1.9,μ1=0.1,μ2=0.1,α1=0.2,α2=0.3,β=0.4,γ=0.6,ρ=0.2,r=0.04,K=0.2,T=0.1. The BSDE for pricing and hedging this spread option is given by (53) dYt=rYt+μ1rα1(St1)1βZt1+μ2rα2(St2)1γZt2dt+(Zt1+ρZt2)dWt1+1ρ2Zt2dWt2.(53) There is no analytic formula available for calculating the reference values. We expect to obtain similar satisfactory results for this example as for the example with GBMs in Figure . Therefore, we use the BCOS method with very large numbers for M and N to get the reference values Y0=0.0710487, Z01=0.137419 and Z02=0.216524. To double-check our results, we used Monte Carlo to obtain, respectively, the following 95%-confidence intervals for these reference values: [0.0698,0.0723], [0.13696,0.13781] and [0.21719,0.21573]. The convergence of the BCOS method is shown in Figure , as expected we obtain first-order convergence with the Euler scheme.

Figure 5. Spread option with the Euler scheme and CEV FSDEs (Equation52).

Figure 5. Spread option with the Euler scheme and CEV FSDEs (Equation52(47) Zm1,Δ(x)=1Δtθ2Em[Ym+1Δ(Xm+1Δ)ΔWm+11]−1−θ2θ2Em[Zm+11,Δ(Xm+1Δ)+ρZm+12,Δ(Xm+1Δ)]+1−θ2θ2Em[f(tm+1,Λm+1Δ(Xm+1Δ))ΔWm+11]−ρZm2,Δ(x),(47) ).

Example 4.5.

For the SABR model, we use very large numbers for M and N to get reference value V0, Z01 and Z02, like we did in Example 4.4. In this experiment we have the following SABR FSDEs for the log-asset price Xt1 and the log volatility Xt2 under the P-measure: (54) dXt1=[μ12exp(2Xt2+2(β1)Xt1)]dt+exp(Xt2+(β1)Xt1)dWt1,X01=x1,dXt2=12ν2dt+ρνdWt1+1ρ2νdWt2,X02=x2,(54) We make a self-financing portfolio consisting of assets exp(Xt1), assets depending on Xt2 (for example a volatility swap), and bonds and bonds.Footnote8 By using Formula (Equation36), we find (55) dYt=rYt+μrexp(Xt2+(β1)Xt1)Zt1dt+(Zt1+ρZt2)dWt1+Zt2dWt2.(55) For the experiment we take the following parameter values: x1=log(2),x2=log(0.4),μ=0.2,β=0.8,ν=0.4,ρ=0.25,r=0.04,K=1.9,T=0.1.

The convergence for Y0, Z01 and Z02 are shown in Figure , where we used the following reference values for the European call option: Y0=0.149767, Z01=0.489597 and Z02=0.030305. We observe first-order convergence for the Euler scheme and second-order convergence when we use Richardson extrapolation on the Euler results. We also achieve second-order convergence with the 2.0-weak-Taylor scheme, but just as in Example 4.1, the corresponding computational costs are high. By comparing the CPU times in Tables  and , we observe that one time step with the BCOS method is around three times more expensive than one time step with the COS method. As explained in Section 3.2, these extra costs are caused by the two additional processes Zt1 and Zt2. Moreover, the 2.0-weak-Taylor scheme is also more expensive because of the use of P=10 Picard iterations per time step in the implicit Formula (Equation46) and the additional expectations in Formulas (Equation47) and (Equation48) caused by θ2=0.5.

Figure 6. SABR example.

Figure 6. SABR example.

Table 2. Number of time steps and CPU time needed to obtain a certain accuracy in option price for the Euler (with and without Richardson extrapolation) and 2.0-weak-Taylor schemes.

5. Conclusion

In this paper, we considered the COS method for pricing European and Bermudan options and the BCOS method for solving backwards stochastic differential equations. We made use of the bivariate ChF of the discretized FSDEs in this technique, where we use the Euler-Maruyama or the 2.0-weak-Taylor scheme for the discretization. The use of these schemes in combination with the COS method results, respectively, in first-order and second-order convergence. Second-order convergence can also be observed by using Richardson extrapolation in combination with a Euler–Maruyama discretization on the forward process, which provides a significant reduction in computational costs compared to the 2.0-weak-Taylor scheme.

Disclosure statement

No potential conflict of interest was reported by the authors.

Notes

1 We omit the mean reverting term for the Heston variance.

2 From now on, we will refer to this scheme as the Euler scheme.

3 If desired, it is possible to choose Δt non-constant.

4 We left out the discount term exp(rΔt) for convenience.

5 For the COS method, we observe smooth, monotonic convergence for the Euler scheme, this behaviour is only observed when θ2=1 for the BCOS method

6 Sometimes additional assumptions are required to complete a market [Citation8,Citation32], such as the tradability of volatility swaps in Example 4.5.

7 In this paper, we assume a deterministic risk-free interest rate, which implies that the risk-free measure and the forward measure coincide.

8 Note that we transform Xt1 and Xt2, before applying the theory of Section 3.1.

References

  • A. Antonov, M. Konikov, and M. Spector, SABR spreads its wings, Risk (2013), pp. 58–63.
  • P. Balland and Q. Tran, SABR goes normal, Risk (2013), pp. 76–81.
  • F. Black and M. Scholes, The pricing of options and corporate liabilities, J. Polit. Econ. 81(3) (1973), pp. 637–654.
  • B. Bouchard and N. Touzi, Discrete-time approximation and Monte-Carlo simulation of backward stochastic differential equations, Stochastic Processes Appl. 11(2) (2004), pp. 175–206.
  • P.P. Boyle, Options: A Monte Carlo approach, J. Financ. Econ. 4 (1977), pp. 323–338.
  • P. Carr and D.B. Madan, Option valuation using the fast Fourier transform, J. Comput. Finance 2 (1999), pp. 61–73.
  • D. Chen, H.J. Harkonen, and D.P. Newton, Advancing the universality of quadrature methods to any underlying process for option pricing, J. Finan. Econ. 114(3) (2014), pp. 600–612.
  • M.H.A. Davis, Complete-market models of stochastic volatility, Proc. Roy. Soc. A 460(2041) (2004), pp. 11–26.
  • P. Doust, No-arbitrage SABR, J. Comput. Finance 15(3) (2012), pp. 3–31.
  • D.J. Duffy, Finite Difference Methods in Financial Engineering: A Partial Differential Equation Approach, Wiley, Chichester, 2006.
  • N. El Karoui, S. Peng, and M.C. Quenez, Backward stochastic differential equations in finance, Math. Finance 7(1) (1997), pp. 1–71.
  • F. Fang and C.W. Oosterlee, A novel pricing method for European option based on Fourier-cosine series expansions, SIAM J. Sci. Comput. 31(2) (2008), pp. 826–848.
  • F. Fang and C.W. Oosterlee, Pricing early-exercise and discrete barrier options by Fourier-cosine series expansions, Numer. Math. 114(1) (2009), pp. 27–62.
  • F. Fang and C.W. Oosterlee, A Fourier-based valuation method for Bermudan and barrier options under Heston's model, SIAM J. Finan. Math. 2 (2011), pp. 439–463.
  • P. Glasserman, Monte Carlo Methods in Financial Engineering, Springer-Verlag, New York, 2003.
  • L.A. Grzelak and C.W. Oosterlee, From arbitrage to arbitrage-free implied volatilities, J. Comput. Finance 20 (2016), pp. 1–19.
  • P.S. Hagan, D. Kumar, A.S. Lesniewski, and D.E. Woodward, Managing smile risk, Wilmott Mag. (2002), pp. 84–108.
  • P.S. Hagan, D. Kumar, A.S. Lesniewski, and D.E. Woodward, Arbitrage-free SABR, Wilmott Mag. (2014), pp. 60–75.
  • S.L. Heston, A closed-form solution for options with stochastic volatility with applications to bond and currency options, Rev. Finan. Stud. 6(2) (1993), pp. 327–343.
  • D.J. Higham, An introduction to multilevel Monte Carlo for option valuation, Int. J. Comput. Math. 92(12) (2015), pp. 2347–2360.
  • T.P. Huijskens, M.J. Ruijter, and C.W. Oosterlee, Efficient numerical Fourier methods for coupled forward–backward SDEs, J. Comput. Appl. Math. 296 (2016), pp. 593–612.
  • K.J. in 't Hout and S. Foulon, ADI finite difference schemes for option pricing in the Heston model with correlation, Int. J. Numer. Anal. Modeling 7(2) (2010), pp. 303–320.
  • A. Khedher and M. Vanmaele, Discretisation of FBSDEs driven by cádlág martingales, J. Math. Anal. Appl. 435(1) (2016), pp. 508–531.
  • P.E. Kloeden and E. Platen, Numerical Solution of Stochastic Differential Equations, Springer-Verlag, Berlin, 1992.
  • A.L. Lewis, A simple option formula for general jump-diffusion and other exponential Lévy processes, SSRN Paper, 2001.
  • A. Lipton, Mathematical Methods for Foreign Exchange: A Financial Engineer's Approach, World Scientific, Singapore, 2001.
  • J. Obłój, Fine-Tune Your Smile: Correction to Hagan et al, Wilmott Mag. (2008), pp. 102–104.
  • L. Ortiz-Gracia and C.W. Oosterlee, Robust pricing of European options with wavelets and the characteristic function, SIAM J. Sci. Comput. 35(5) (2013), pp. B1055–B1084.
  • E. Pardoux and S.G. Peng, Adapted solution of a backward stochastic differential equation, Syst. Control Lett. 14(1) (1990), pp. 55–61.
  • E. Pardoux and S.G. Peng, Backward stochastic differential equations and quasilinear parabolic partial differential equations, in Stochastic Partial Differential Equations and Their Applications, B.L. Rozovskii and R.B. Sowers, eds., Lecture Notes in Control and Information Sciences, Vol. 176, Springer, Berlin, 1992, pp. 200–217.
  • H. Park, Efficient valuation method for the SABR model, SSRN Paper, 2013.
  • C.W. Potter, Complete stochastic volatility models with variance swaps, Tech. Rep., 2004.
  • M.J. Ruijter and C.W. Oosterlee, Two-dimensional Fourier cosine series expansion method for pricing financial options, SIAM J. Sci. Comput. 34(5) (2012), pp. B642–B671.
  • M.J. Ruijter and C.W. Oosterlee, A Fourier-cosine method for an efficient computation of solutions to BSDEs, SIAM J. Sci. Comput. 37(2) (2015), pp. A859–A889.
  • M.J. Ruijter and C.W. Oosterlee, Numerical Fourier method and second-order Taylor scheme for backward SDEs in finance, Appl. Numer. Math. 103 (2016), pp. 1–26.
  • Z. van der Have, Arbitrage-free methods to price European options under the SABR model, Master's thesis, TU Delft, 2015.
  • L. von Sydow, L.J. Höök, E. Larsson, E. Lindström, S. Milovanović, J. Persson, V. Shcherbakov, Y. Shpolyanskiy, S. Sirén, J. Toivanen, J. Waldén, M. Wiktorsson, M.B. Giles, J. Levesley, J. Li, C.W. Oosterlee, M.J. Ruijter, A. Toropov, and Y. Zhao, BENCHOP: The BENCHmarking project in Option Pricing, Int. J. Comput. Math. 92(12) (2015), pp. 2361–2379.
  • W. Zhao, L. Chen, and S. Peng, A new kind of accurate numerical method for backward stochastic differential equations, SIAM J. Sci. Comput. 28(4) (2006), pp. 1563–1581.

Appendices

Appendix 1

In this appendix, we give the functions for Formula (Equation6) associated with the 2.0-weak-Taylor scheme. (A.1a) m1(x)=μ1(x)12σ1(x)σ1(x)x1+ρσ2(x)σ1(x)x2+12μ1(x)μ1(x)x1+μ2(x)μ1(x)x2+12σ12(x)2μ1(x)x12+ρσ1(x)σ2(x)2μ1(x)x1x2+12σ22(x)2μ1(x)(x2)2Δt,(A.1a) (A.1b) s11(x)=σ1(x)+12σ1(x)μ1(x)x1+ρσ2(x)μ1(x)x2+μ1(x)σ1(x)x1+μ2(x)σ1(x)x2+12σ12(x)2σ1(x)x12+ρσ1(x)σ2(x)2σ1(x)x1x2+12σ22(x)2σ1(x)(x2)2Δt,(A.1b) (A.1c) s12(x)=1ρ22σ2(x)μ1(x)x2Δt,    κ11,2(x)=1ρ22σ2(x)σ1(x)x2,(A.1c) (A.1d) κ11(x)=12σ1(x)σ1(x)x1+ρσ2(x)σ1(x)x2,κ12(x)=0,(A.1d) (A.1e) ψ1(x)=1ρ22σ2(x)σ1(x)x2,(A.1e) (A.1f) m2(x)=μ2(x)12ρσ1(x)σ2(x)x1+σ2(x)σ2(x)x2+12μ1(x)μ2(x)x1+μ2(x)μ2(x)x2+12σ12(x)2μ2(x)x12+ρσ1(x)σ2(x)2μ2(x)x1x2+12σ22(x)2μ2(x)(x2)2Δt,(A.1f) (A.1g) s21(x)=ρσ2(x)+12σ1(x)μ2(x)x1Δt+ρ2σ2(x)μ2(x)x2+μ1(x)σ2(x)x1+μ2(x)σ2(x)x2+12σ12(x)2σ2(x)x12+ρσ1(x)σ2(x)2σ2(x)x1x2+12σ22(x)2σ2(x)(x2)2Δt,(A.1g) (A.1h) s22(x)=1ρ2σ2(x)+1ρ22σ2(x)μ2(x)x2+μ1(x)σ2(x)x1+μ2(x)σ2(x)x2+12σ12(x)2σ2(x)x12+ρσ1(x)σ2(x)2σ2(x)x1x2+12σ22(x)2σ2(x)(x2)2Δt,(A.1h) (A.1i) κ21,2(x)=1ρ22σ1(x)σ2(x)x1+2ρσ2(x)σ2(x)x2,(A.1i) (A.1j) κ21(x)=ρ2σ1(x)σ2(x)x1+ρσ2(x)σ2(x)x2,(A.1j) (A.1k) κ22(x)=1ρ22σ2(x)σ2(x)x2,ψ2(x)=1ρ22σ1(x)σ2(x)x1.(A.1k)

Appendix 2

In this appendix, we provide some information on how to choose the computational domain [a1,b1]×[a2,b2]. The FSDEs are given by Formula (Equation1), that is, (A.2) dXt1=μ1(Xt)dt+σ1(Xt)dWt1,dXt2=μ2(Xt)dt+ρσ2(Xt)dWt1+1ρ2σ2(Xt)dWt2,(A.2) where X01=x1 and X02=x2.

When the first, second and fourth cumulants of XT1 and XT2 given (x1,x2) are known we advise to choose the domain [a1,b1]×[a2,b2] as presented in [Citation33], so for i=1,2: (A.3) [ai,bi]=[xi+ci,1Lci,2+ci,4, xi+ci,1+Lci,2+ci,4],(A.3) where L=10 and ci,j denotes the j-th cumulant of XTi given (x1,x2). When the first cumulant is relatively large, we suggest to use a time-dependent domain, that is, (A.4) [ai(t),bi(t)]=[xi+ci,1(t)Lci,2(t)+ci,4(t), xi+ci,1(t)+Lci,2(t)+ci,4(t)],(A.4) where ci,j(t) denotes the j-th cumulant of Xti given (x1,x2). This time dependency is important, because otherwise, for example xi may be outside the interval [ai,bi]. The (B)COS method only needs a minor adaptation regarding the time-dependent domain: ai and bi have to be replaced by ai(tm) and bi(tm), respectively, in Formulas (Equation25), (Equation26) and in each formula in Appendix 3.

When the cumulants of XT1 and/or XT2 given (x1,x2) are unknown, we choose the domain as in [Citation36]: (A.5) [ai,bi]=[ci,1Lci,2,ci,1+Lci,2],(A.5) where L=10, ci,1=xi+μi(x1,x2)T and ci,2=σi(x1,x2)T. When desired, this domain can also be time dependent. The interval [ai,bi] can be adjusted such that it satisfies specific constraints of Xti, for example, non-negativity.

We used Equation (A.5) for all experiments for the SABR model, but not for problems in which a drift function is large compared to the associated volatility, for example, (A.6) dXt=αtdt,X0=x,dα0=γαtdWt,σ0=α,(A.6) where W is a standard Brownian motion and the vol–vol parameter γ0.

If we follow Equation (A.5), we would obtain (A.7) [a1(t),b1(t)]=x+αt.(A.7) This is incorrect, because Xt is not deterministic.

A better domain [a1(t),b1(t)]×[a2(t),b2(t)] in this case would be based on Formula (A.5), as follows: (A.8) [a2(t),b2(t)]=[max{0,αLγαt},α+Lγαt],(A.8) so that we find (A.9) Xtx+0ta2(t)dt=a1(t),(A.9) and (A.10) Xtx+0tb2(t)dt=x+αt+23Lγαt3/2=b1(t).(A.10)

Appendix 3

In this appendix, we give approximations of Em[h(tm+1,Xm+1Δ)], Em[h(tm+1,Xm+1Δ)ΔWm+11] and Em[h(tm+1,Xm+1Δ)ΔWm+12] with the COS method for a general sufficiently smooth function h(t,x). We denote (A.11) Hk1,k2(tm+1)=2b1a12b2a2a2b2a1b1h(tm+1,x)cosk1πx1a1b1a1cosk2πx2a2b2a2dx1dx2.(A.11) Using the two-dimensional COS method, we find (A.12) Em[h(tm+1,Xm+1Δ)]k1=0N11k2=0N2112φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1ik2πa2b2a2+φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1+ik2πa2b2a2Hk1,k2(tm+1).(A.12) Also, we have (A.13) Em[h(tm+1,Xm+1Δ)ΔWm+11]k1=0N11k2=0N2112(id1+Δtd2+d3+(Δt)22d1+d4+(Δt)2)φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1ik2πa2b2a2+(id1Δtd2d3(Δt)22d1d4(Δt)2)φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1+ik2πa2b2d2Hk1,k2(tm+1),(A.13) and (A.14) Em[h(tm+1,Xm+1Δ)ΔWm+12]k1=0N11k2=0N2112(id2+Δtd1+d3+(Δt)22d2+d5+(Δt)2)φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1ik2πa2b2a2+(id2Δtd1d3(Δt)22d2d5(Δt)2)φXm+1Δk1πb1a1,k2πb2a2XmΔ=xexpik1πa1b1a1+ik2πa2b2a2Hk1,k2(tm+1),(A.14) where we abbreviated: (A.15) d1±=k1πb1a1s11(x)±k2πb2a2s21(x),d4±=k1πb1a1κ11(x)±k2πb2a2κ21(x),d2±=k1πb1a1s12(x)±k2πb2a2s22(x),d5±=±k2πb2a2κ22(x),d3±=k1πb1a1κ11,2(x)±k2πb2a2κ21,2(x).(A.15) The approximations (A.13) and (A.14) have, besides the error introduced by the COS method, an error of order O((Δt)3) [Citation36].