Paper 3
Paper 3
Abstract. This paper deals with the numerical solution of the Heston par-
tial differential equation (PDE) that plays an important role in financial op-
tion pricing theory, Heston (1993). A feature of this time-dependent, two-
dimensional convection-diffusion-reaction equation is the presence of a mixed
spatial-derivative term, which stems from the correlation between the two un-
derlying stochastic processes for the asset price and its variance.
Semi-discretization of the Heston PDE, using finite difference schemes on
non-uniform grids, gives rise to large systems of stiff ordinary differential equa-
tions. For the effective numerical solution of these systems, standard implicit
time-stepping methods are often not suitable anymore, and tailored time-
discretization methods are required. In the present paper, we investigate four
splitting schemes of the Alternating Direction Implicit (ADI) type: the Douglas
scheme, the Craig–Sneyd scheme, the Modified Craig–Sneyd scheme, and the
Hundsdorfer–Verwer scheme, each of which contains a free parameter.
ADI schemes were not originally developed to deal with mixed spatial-
derivative terms. Accordingly, we first discuss the adaptation of the above
four ADI schemes to the Heston PDE. Subsequently, we present various nu-
merical examples with realistic data sets from the literature, where we consider
European call options as well as down-and-out barrier options. Combined with
ample theoretical stability results for ADI schemes that have recently been ob-
tained in In ’t Hout & Welfert (2007, 2009) we arrive at three ADI schemes
that all prove to be very effective in the numerical solution of the Heston PDE
with a mixed derivative term. It is expected that these schemes will be useful
also for general two-dimensional convection-diffusion-reaction equations with
mixed derivative terms.
1. Introduction
In the Heston model, values of options are given by a time-dependent partial
differential equation (PDE) that is supplemented with initial and boundary condi-
tions [7, 14, 22, 24]. The Heston PDE constitutes an important two-dimensional
extension to the celebrated, one-dimensional, Black–Scholes PDE. Contrary to the
Received by the editors March 8, 2009 and, in revised form, September 4, 2009.
2000 Mathematics Subject Classification. 65L05, 65L20, 65M06, 65M12, 65M20.
This work is supported financially by the Research Foundation–Flanders, FWO contract no.
G.0125.08.
303
304 K. J. IN ’T HOUT AND S. FOULON
options a closed-form analytical pricing formula has only been obtained [14] in the
literature if the correlation ρ = 0.
Section 4 summarizes our conclusions concerning the four ADI schemes in the
numerical solution of the Heston PDE with a mixed derivative term. Subsequently,
several issues for future research are discussed.
where K > 0 is the given strike price of the option. Further, a boundary condition
at s = 0 holds,
(2.3) u(0, v, t) = 0 (0 ≤ t ≤ T ).
∂u
(2.4) (S, v, t) = e−rf t (0 ≤ t ≤ T ),
∂s
Clearly, the boundary conditions (2.3), (2.5) are of Dirichlet type, whereas (2.4)
is of Neumann type. We take S = 8K and V = 5. This yields a negligible modeling
error with respect to (2.1)–(2.3) on the unbounded domain for a wide range of
parameter values.
1We assume w.l.o.g. that the market price of volatility risk is equal to zero.
306 K. J. IN ’T HOUT AND S. FOULON
3
V
0
0 100 200 300 400 500 600 700 800
S
The FD schemes above are all well-known. Formulas (2.9b), (2.10), (2.11) are
central schemes, whereas (2.9a), (2.9c) are upwind schemes. We note that these
schemes have previously been applied by Kluge [13] in the numerical solution of
the Heston PDE. Through Taylor expansion it can be verified that each of the
formulas (2.9), (2.10), (2.11) has a second-order truncation error, provided that the
function f is sufficiently often continuously differentiable and the meshes in the x-
and y-directions are smooth, cf. (2.7).
The actual FD discretization of the initial-boundary value problem for the Heston
PDE is performed as follows.
In view of the Dirichlet boundary conditions (2.3) and (2.5), the grid in [0, S] ×
[0, V ] is given by
G = {(si , vj ) : 1 ≤ i ≤ m1 , 0 ≤ j ≤ m2 − 1}.
At this grid, each spatial derivative appearing in (2.1) is replaced by its correspond-
ing central FD scheme (2.9b), (2.10), or (2.11) – except in the region v > 1 and at
the boundaries v = 0 and s = S.
In the region v > 1 we apply the upwind scheme (2.9a) for ∂u/∂v whenever
the flow in the v-direction is towards v = V . This is done so as to avoid spurious
oscillations in the FD solution when the volatility-of-variance σ is close to zero.
At the outflow boundary v = 0 the derivative ∂u/∂v is approximated using the
upwind scheme (2.9c). All other derivative terms in the v-direction vanish at v = 0,
due to the factor v occurring in (2.1), and hence, these terms do not require further
treatment.
At the boundary s = S the spatial derivatives in the s-direction need to be
considered. First, the Neumann condition (2.4) at s = S implies that the mixed
derivative ∂ 2 u/∂s∂v vanishes there. Next, the derivative ∂u/∂s is directly given by
(2.4). Finally, we approximate ∂ 2 u/∂s2 at s = S = sm1 using the central scheme
(2.10) with the virtual point S + ∆sm1 , where the value at this point is defined by
extrapolation using (2.4).
The FD discretization described above of the initial-boundary value problem
(2.1)–(2.5) for the Heston PDE yields an initial value problem for a large system
of stiff ordinary differential equations (ODEs),
Here A is a given m × m–matrix and b(t) (t ≥ 0), U0 are given m–vectors with m =
m1 m2 . The vector U0 is directly obtained from the initial condition (2.2) and the
vector function b depends on the boundary conditions (2.3)–(2.5). For each given
t > 0, the entries of the solution vector U (t) to (2.12) constitute approximations to
the exact solution values u(s, v, t) of (2.1)–(2.5) at the spatial grid points (s, v) ∈ G,
ordered in a convenient way.
Thus, each step (2.14) requires the solution of a system of linear equations involving
the matrix (I − 12 ∆tA) where I denotes the m×m identity matrix. Since (I − 12 ∆tA)
does not depend on the step index n, one can compute a LU factorization of this
matrix once, beforehand, and next apply it in all steps (2.14) to obtain Un (n ≥ 1).
The Crank–Nicolson scheme can be practical when the number of spatial grid
points m = m1 m2 is moderate. In our application to the two-dimensional Heston
PDE, however, m usually gets large and the Crank–Nicolson scheme becomes inef-
fective. The reason for this is that (I − 21 ∆tA), and hence the matrices in its LU
factorization, possess a bandwidth that is directly proportional to min{m1 , m2 }.
For the numerical solution of the semi-discretized Heston problem (2.12) we shall
consider in this paper splitting schemes of the ADI type. We decompose the matrix
A into three submatrices,
A = A0 + A1 + A2 .
We choose the matrix A0 as the part of A that stems from the FD discretization
of the mixed derivative term in (2.1). Next, in line with the classical ADI idea, we
choose A1 and A2 as the two parts of A that correspond to all spatial derivatives
in the s- and v-directions, respectively. The rd u term in (2.1) is distributed evenly
over A1 , A2 . The FD discretization described in Sect. 2.2 implies that A1 , A2 are
essentially tridiagonal and pentadiagonal, respectively.
Write b(t) from (2.12) as b(t) = b0 (t) + b1 (t) + b2 (t) where the decomposition is
analogous to that of A. Next, define functions Fj (j = 0, 1, 2) by
Currently, the most comprehensive stability results for the Do, CS, MCS and
HV schemes relevant to multi-dimensional PDEs with mixed derivative terms have
been obtained in [9, 10]. We review the main conclusions from loc. cit. relevant to
the situation of our paper.
The Do and CS schemes are both unconditionally stable when applied to 2D
convection-diffusion equations with a mixed derivative term whenever θ ≥ 12 . In
particular, the second-order CS scheme is unconditionally stable.
The MCS and HV schemes are unconditionally stable when applied to 2D pure √
diffusion equations with a mixed derivative term whenever θ ≥ 31 and θ ≥ 1 − 12 2,
respectively. Recall that the MCS and HV schemes are of order two for any given
θ. At this moment, unconditional stability results for the MCS and HV schemes
in the general situation of 2D equations, with convection, are lacking. It was
conjectured [9] however that the HV scheme is unconditionally stable when applied
to 2D√convection-diffusion equations with a mixed derivative term whenever θ ≥
1 1
2 + 6 3.
3. Numerical experiments
3.1. Spatial discretization error. In this section we consider four numerical
examples and assess the actual convergence behavior of the FD discretization (2.12)
of the Heston PDE defined in Sect. 2.2. For any given numbers of mesh points m1 ,
m2 in the s- and v-directions, we define the global spatial discretization error at
time t = T by
e(m1 , m2 ) = max |u(si , vj , T ) − Uk (T )| : 12 K < si < 23 K, 0 < vj < 1 .
Here u denotes the exact European call option price function, satisfying the initial-
boundary value problem (2.1)–(2.3) for the Heston PDE on the unbounded domain.
Next, Uk designates the component of the exact solution U to (2.12) that corre-
sponds to the grid point (si , vj ).
Clearly, the global spatial error is defined via a maximum norm. The set of
asset prices s ∈ ( 21 K, 32 K) and variances v ∈ (0, 1) in our definition encompasses
most situations of practical interest. We note that the modeling error, that was
introduced by restricting the domain of the Heston PDE to a bounded set, is
also contained in e(m1 , m2 ). In our experiments this contribution turns out to be
negligible.
For the actual computation √ of the global spatial errors e(m1 , m2 ) we apply the
HV scheme with θ = 12 + 61 3 to (2.12) with the small time step ∆t = T /1000
to obtain a sufficiently accurate approximation of U (T ). Subsequently, we employ
an implementation of Heston’s semi-analytical formula [7] to acquire values of u.
For calculating the single integrals occurring in this formula we use a numerical
quadrature rule. Numerical difficulties one can encounter in the implementation,
due to the presence of multi-valued complex functions, have recently been discussed
in [1]. We adopt the algorithm proposed in loc. cit. where branch cuts in the
complex plane are correctly taken into account.
We perform numerical experiments in the four cases of parameter sets given by
Table 1. Observe that in three of the four cases there is a substantial correlation
factor ρ. Only in case 4 the correlation factor is relatively small.
Case 1 has been taken from Albrecher et. al. [1], where we have chosen T = 1.
Case 2 comes from Bloomberg [3]. A special feature of this parameter set is that
σ is close to zero, which implies that the Heston PDE is convection-dominated in
the v-direction. Values for rd , rf and T were not specified in [3] and have been
ADI FINITE DIFFERENCE SCHEMES FOR OPTION PRICING 313
Table 1. Parameters for the Heston model and European call options.
Case 1 Case 2
150 150
100 100
U U
50 50
0 0
1 1
200 200
0.5 100 0.5 100
V 0 0 S V 0 0 S
Case 3 Case 4
150 150
100 100
U U
50 50
0 0
1 1
200 200
0.5 100 0.5 100
V 0 0 S V 0 0 S
chosen separately. Case 3 has been taken from Schoutens et. al. [21]. Here the
Feller condition is only just met. Finally, case 4 stems from Winkler et. al. [27].
Figure 2 displays the exact option price functions u corresponding to the four
cases of Table 1 on the domain (s, v) ∈ [0, 200] × [0, 1].
Figure 3 subsequently shows for each case from Table 1 the global spatial errors
e(2m2 , m2 ) vs. 1/m2 for m2 = 10, 20, . . . , 100. We have taken m1 = 2m2 as it turns
out that, for efficiency reasons, one can use much less points in the v-direction than
in the s-direction. To determine the numerical order of convergence p of the spatial
discretization, we have fitted in each case a straight line to the outcomes for the
global spatial errors. Accordingly, in the cases 1, 2, 3, 4 we found the respective
314 K. J. IN ’T HOUT AND S. FOULON
Case 1 0
Case 2
0
10 10
−1 −1
Spatial error
Spatial error
10 10
−2 −2
10 10
−3 −3
10 10 −2 −1
−2 −1
10 1/m2 10 10 1/m2 10
0
Case 3 Case 4
0
10 10
−1 −1
Spatial error
Spatial error
10 10
−2 −2
10 10
−3 −3
10 −2 −1
10 −2 −1
10 1/m2 10 10 1/m 10
2
orders p = 1.9, 2.0, 2.1, 2.4. This clearly suggests that the FD discretization of the
Heston PDE described in Sect. 2.2 is convergent of appr. order two.
We remark that for the global spatial errors in relative sense, we obtained in
each case that it is close to 1.0 percent for m2 = 30 and decreases to approximately
0.1 percent for m2 = 100. Here asset-variance pairs (si , vj ) are considered such
that the option value u(si , vj , T ) is always at least 1.
Finally, we repeated the numerical experiments above for uniform grids. Then
the global spatial errors are always found to be much larger, compared to the non-
uniform grid under consideration with the same number of points. Furthermore,
the error behavior is erratic in this case. The latter can be seen to be related to the
relative position of the point s = K to the mesh in the s-direction, and concerns
a known phenomenon, see e.g. Tavella & Randall [24]. This behavior is much less
noticeable for (suitable) non-uniform grids.
3.2. Temporal discretization error. In this section we present numerical ex-
periments for the ADI schemes (2.17), (2.18), (2.19), (2.20) which yields important
insight into their actual stability and convergence behavior in the application to
semi-discretized Heston problems (2.12) with non-zero correlation.
We define the global temporal discretization error at time t = T = N ∆t by
eb (N ; m1 , m2 ) = max |Uk (T ) − UN, k | : 12 K < si < 23 K, 0 < vj < 1 ,
ADI FINITE DIFFERENCE SCHEMES FOR OPTION PRICING 315
Case 1 Case 2
1 1
10 10
−1 −1
10 10
Temporal error
Temporal error −3
10
−3
10
−5 −5
10 10
−7 −7
10 10 −3 −2 −1 0
−3 −2 −1 0
10 10 10 10 10 10 10 10
1/N 1/N
1
Case 3 1
Case 4
10 10
−1 −1
10 10
Temporal error
Temporal error
−3 −3
10 10
−5 −5
10 10
−7 −7
10 −3 −2 −1 0
10 −3 −2 −1 0
10 10 10 10 10 10 10 10
1/N 1/N
where the index k is such that Uk (T ) and UN, k correspond to the spatial grid point
(si , vj ). Clearly, the global temporal error is defined for the same (s, v)-domain as
the global spatial error and we also deal again with the maximum norm, cf. Sect. 3.1.
Motivated by the theoretical stability and accuracy results discussed in Sect. 2.3,
we shall consider the Do and CS schemes with θ = 12 and the MCS scheme with
√
θ = 13 . Next, we consider the HV scheme for the two values θ = 1 − 12 2 ≈ 0.293
√
and θ = 12 + 16 3 ≈ 0.789, to which we refer in the following as HV1 and HV2,
respectively. We apply all these ADI schemes in each of the four cases of parameter
sets for European call options in the Heston model listed in Table 1.
In addition, we also apply the Runge–Kutta–Chebyshev (RKC) scheme. This
is an explicit second-order Runge–Kutta scheme which has been constructed such
that its stability region includes a large interval [−β, 0] along the negative real axis.
For a complete discussion of this method see e.g. [12, 23]. The RKC scheme has a
free parameter ε. Since the Heston PDE contains a convective part, we have chosen
ε = 10, cf. [25]. In this case β ≈ 0.34(ν 2 − 1), where ν denotes the number of stages
of the scheme. Accordingly, in a given application to a semi-discretized Heston
316 K. J. IN ’T HOUT AND S. FOULON
1
Case 1 Case 2
1
10 10
−1 −1
10 10
Temporal error
Temporal error
−3 −3
10 10
−5 −5
10 10
−7 −7
10 −3 −2 −1 0
10 −3 −2 −1 0
10 10 10 10 10 10 10 10
1/N 1/N
1
Case 3 1
Case 4
10 10
−1 −1
10 10
Temporal error
Temporal error
−3 −3
10 10
−5 −5
10 10
−7 −7
10 −3 −2 −1 0
10 −3 −2 −1 0
10 10 10 10 10 10 10 10
1/N 1/N
problem (2.12)p with time step ∆t, the number of stages is taken as the smallest
integer ν ≥ 1 + 3∆t r[A] where r[A] denotes the spectral radius of A.
Figures 4, 5 display for m1 = 2m2 = 100 and m1 = 2m2 = 200, respectively,
the results for the global temporal errors eb (N ; m1 , m2 ) vs. 1/N for a range of step
numbers N between N = 1 and N = 1000. We applied the HV2 scheme with
N = 5000 to obtain a reference value for U (T ).
A first observation from Figures 4, 5 is that in the cases 1, 3, 4 the RKC scheme
has global temporal errors that are often comparable to those of the MCS scheme.
For relatively large ∆t, however, the RKC scheme requires a large number of stages
ν since in all cases r[A] ≈ 5.1 ·104 (if m1 = 2m2 = 100) and r[A] ≈ 2.2 ·105 (if m1 =
2m2 = 200). In our experiments, RKC always turned out to be considerably less
efficient than MCS. Note that because the RKC scheme does produce fair temporal
errors for all N in each of the cases 1, 3, 4, this suggests that the eigenvalues of
the corresponding matrices A all lie close to the negative real axis in these three
cases. In case 2, the RKC scheme clearly shows instability for, at least, values
N < 100. We explain this from the corresponding matrices A to have eigenvalues
ADI FINITE DIFFERENCE SCHEMES FOR OPTION PRICING 317
in the left half of the complex plane with substantial imaginary parts. This is
also conceivable, since in case 2 the Heston PDE is convection-dominated in the
v-direction, cf. Sect. 3.1.
For the Do, CS, MCS and HV2 schemes, the Figures 4, 5 clearly reveal that in
all cases 1–4 the global temporal errors always stay below a moderate bound and
decrease monotonically with N . This is a very favorable result and indicates an
unconditionally stable behavior of these ADI schemes in all cases. Note that it is a
non-trivial result, as it does not directly follow e.g. from the von Neumann analysis
discussed in Sect. 2.3.
For the HV1 scheme, case 2 shows a peak in the global temporal errors which is
higher if m1 = 2m2 = 200 than if m1 = 2m2 = 100. We conjecture that the HV1
scheme is just conditionally stable, under a CFL condition. We mention in passing
that the same is found for the HV scheme with value θ = 0.3, cf. [8].
Subsequent analysis of the results in Figures 4, 5 yields for the MCS and HV2
schemes in each case 1–4 a convergence order equal to 2.0. Moreover, the conver-
gence behavior is of the form C(∆t)2 (0 < ∆t ≤ τ ) with constants C, τ > 0 that
are only weakly dependent on the number of spatial grid points m. Hence, the MCS
and HV2 schemes show a stiff order of convergence equal to two. This clearly agrees
with their orders of consistency, cf. Sect. 2.3. Remark that this is not obvious, as
the order of consistency is a priori only relevant to fixed, non-stiff ODEs.
For the HV1 scheme, we obtain a stiff order of convergence equal to two in the
cases 1, 3, 4.
The Do and CS schemes exhibit in all cases an undesirable convergence behavior,
with temporal errors that are relatively large for modest time steps ∆t. This
atypical behavior becomes more pronounced when m1 , m2 get larger. Though
the results are not included in the figures, we remark that the (time-consuming)
Crank–Nicolson scheme shows a similar behavior. The cause for this phenomenon
is related to the fact that at s = K the payoff function (2.2) is non-smooth and the
Do, CS and Crank–Nicolson schemes do not sufficiently damp local (high-frequency)
errors incited by this. A remedy for this situation is to first apply, at t = 0, two
backward Euler steps with step size ∆t/2, and then to proceed onwards from t = ∆t
with the time-stepping schemes under consideration, cf. Rannacher [20] and also
e.g. [6, 12, 18]. By adopting this damping procedure, we recover in each case 1–4 a
stiff order of convergence equal to one for the Do scheme and equal to two for the
CS scheme.
Concerning implementation, we mention that all codes have been written in
Matlab, version 7.8, where all matrices have been defined as sparse. As an indication
for the computing times, our implementation of the CS, MCS, HV schemes each
took per time step about 0.005 cpu-sec (if m1 = 2m2 = 100) and 0.02 cpu-sec
(if m1 = 2m2 = 200) on an Intel Duo Core T5500 1.6 GHz processor with 1 GB
memory. For the Do scheme these times are appr. halved. Note that the times are
directly proportional to the number of spatial grid points m.
u(B, v, t) = 0 (0 ≤ t ≤ T )
318 K. J. IN ’T HOUT AND S. FOULON
and
u(s, V, t) = (s − B)e−rf t (0 ≤ t ≤ T ).
performance of ADI schemes relevant to other exotic options in the Heston model
is of much interest. The use of variable time steps, especially near the initial
time t = 0, is likely to increase the efficiency, cf. e.g. [13]. Finally, ADI schemes
can be attractive in the numerical solution of other multi-dimensional PDEs from
finance with mixed derivative terms, e.g. the three-dimensional hybrid Heston–
Hull–White model. The extension of the ADI schemes (2.17)–(2.20) to such PDEs
is straightforward. Positive results on unconditional stability of these schemes in
arbitrary spatial dimensions, for pure diffusion equations with mixed derivative
terms, have recently been proved in [10].
Acknowledgments
The first author wishes to thank Willem Hundsdorfer for suggesting to consider
the RKC scheme and for providing a Matlab implementation of this scheme. He is
also grateful to Marc Spijker for his comments, which have enhanced the presenta-
tion of this paper. Further, the authors would like to thank an anonymous referee
for constructive remarks.
References
[1] H. Albrecher, P. Mayer, W. Schoutens & J. Tistaert, The little Heston trap, Wilmott Mag.,
January 2007, 83–92.
[2] J. Andreasen, Back to the future, Risk 18 (2005) 104–109.
[3] Bloomberg Quant. Finan. Devel. Group, Barrier options pricing under the Heston model,
2005.
[4] I. J. D. Craig & A. D. Sneyd, An alternating-direction implicit scheme for parabolic equations
with mixed derivatives, Comp. Math. Appl. 16 (1988) 341–350.
[5] J. Douglas & H. H. Rachford, On the numerical solution of heat conduction problems in two
and three space variables, Trans. Amer. Math. Soc. 82 (1956) 421–439.
[6] M. B. Giles & R. Carter, Convergence analysis of Crank–Nicolson and Rannacher time-
marching, J. Comp. Finan. 9 (2006) 89–112.
[7] S. L. Heston, A closed-form solution for options with stochastic volatility with applications
to bond and currency options, Rev. Finan. Stud. 6 (1993) 327–343.
[8] K. J. in ’t Hout, ADI schemes in the numerical solution of the Heston PDE, in: Numerical
Analysis and Applied Mathematics, eds. T. E. Simos et. al., AIP Conf. Proc. 936 (2007)
10–14.
[9] K. J. in ’t Hout & B. D. Welfert, Stability of ADI schemes applied to convection-diffusion
equations with mixed derivative terms, Appl. Num. Math. 57 (2007) 19–35.
[10] K. J. in ’t Hout & B. D. Welfert, Unconditional stability of second-order ADI schemes
applied to multi-dimensional diffusion equations with mixed derivative terms, Appl. Num.
Math. 59 (2009) 677–692.
[11] W. Hundsdorfer, Accuracy and stability of splitting with Stabilizing Corrections, Appl. Num.
Math. 42 (2002) 213–233.
[12] W. Hundsdorfer & J. G. Verwer, Numerical Solution of Time-Dependent Advection-Diffusion-
Reaction Equations, Springer, Berlin, 2003.
[13] T. Kluge, Pricing derivatives in stochastic volatility models using the finite difference method,
Dipl. thesis, TU Chemnitz, 2002.
[14] A. Lipton, Mathematical Methods for Foreign Exchange, World Scientific, Singapore, 2001.
[15] S. McKee & A. R. Mitchell, Alternating direction methods for parabolic equations in two
space dimensions with a mixed derivative, Computer J. 13 (1970) 81–86.
[16] S. McKee, D. P. Wall & S. K. Wilson, An alternating direction implicit scheme for parabolic
equations with mixed derivative and convective terms, J. Comp. Phys. 126 (1996) 64–76.
[17] D. W. Peaceman & H. H. Rachford, The numerical solution of parabolic and elliptic differ-
ential equations, J. Soc. Ind. Appl. Math. 3 (1955) 28–41.
[18] D. M. Pooley, K. R. Vetzal & P. A. Forsyth, Convergence remedies for non-smooth payoffs
in option pricing, J. Comp. Finan. 6 (2003) 25–40.
[19] C. Randall, PDE Techniques for Pricing Derivatives with Exotic Path Dependencies or Exotic
Processes, Lecture notes, Workshop CANdiensten, Amsterdam, 2002.
320 K. J. IN ’T HOUT AND S. FOULON
[20] R. Rannacher, Finite element solution of diffusion problems with irregular data, Numer.
Math. 43 (1984) 309–327.
[21] W. Schoutens, E. Simons & J. Tistaert, A perfect calibration ! Now what ?, Wilmott Mag.,
March 2004, 66–78.
[22] S. E. Shreve, Stochastic Calculus for Finance II, Springer, New York, 2004.
[23] B. P. Sommeijer, L. F. Shampine & J. G. Verwer, RKC: An explicit solver for parabolic
PDEs, J. Comp. Appl. Math. 88 (1997) 315–326.
[24] D. Tavella & C. Randall, Pricing Financial Instruments, Wiley, New York, 2000.
[25] J. G. Verwer, B. P. Sommeijer & W. Hundsdorfer, RKC time-stepping for advection-diffusion-
reaction problems, J. Comp. Phys. 201 (2004) 61–79.
[26] J. G. Verwer, E. J. Spee, J. G. Blom & W. Hundsdorfer, A second-order Rosenbrock method
applied to photochemical dispersion problems, SIAM J. Sci. Comp. 20 (1999) 1456–1480.
[27] G. Winkler, T. Apel & U. Wystup, Valuation of options in Heston’s stochastic volatility
model using finite element methods, in: Foreign Exchange Risk, eds. J. Hakala & U. Wystup,
Risk Publ., 2002.