Pricing American
Pricing American
1. INTRODUCTION
This article surveys the linear programming (LP) approach to fast valuation of American
options through the use of a special version of the revised simplex method. This special
algorithm makes use of the tridiagonal structure of the finite-difference discretization of
the Black–Scholes partial differential equation, a novel basis factorization, and the nature
of the optimal exercise boundary, to create a pricing algorithm that is essentially linear in
the number of discretization steps in space or time with the other held fixed. The method
is applicable to a variable coefficient version of the Black–Scholes equation which allows
the treatment and rapid solution of path-dependent exotic options taking account of local
volatility. One emphasis of the paper is a complete overview of the theory underlying
the LP approach to these difficult problems. Another is a representative survey of the
numerical results obtained to date.
The paper is structured as follows. In Section 2 we review the formulation of the
American put option valuation problem presented in Dempster and Hutton (1997a, 1999)
and Dempster, Hutton, and Richards (1998). The problem is a classical optimal stopping
problem that may be formulated as a free-boundary problem by considering the domain
partition of optimal stopping. Removing any explicit reference to the free boundary,
the option value may be seen to be the unique solution of an order complementarity
problem (Borwein and Dempster 1989) by considering its equivalent formulation as a
variational inequality and utilizing standard results for coercive operators. Finally, the
value is the solution of an abstract linear program (for which a simpler proof than previ-
ously is given) which can be solved with standard LP techniques upon suitable domain
Address correspondence to the author at the Centre for Financial Research, Judge Institute of Manage-
ment Studies, University of Cambridge, Cambridge CB2 1AG, United Kingdom (01223) 339700; e-mail:
m.dempster@jims.cam.ac.uk & d.richards@jims.cam.ac.uk; www-cfr.jims.cam.ac.uk.
© 2000 Blackwell Publishers, 350 Main St., Malden, MA 02148, USA, and 108 Cowley Road, Oxford,
OX4 1JF, UK.
157
158 m. a. h. dempster and d. g. richards
truncation and discretization. In Section 2 basic numerical methods and our variable
coefficient tridiagonal simplex algorithm, together with the degenerate PDE approach
(Wilmott, Howison, and Dewynne 1993; Barraquand and Pudet 1996) to valuing market-
traded, discretely sampled exotic options, are also reviewed.
Section 3 discusses both theory and basic numerical methods for pricing exotic options
with the local volatility surface implied by market values of European options. This is
an area pioneered in Rubinstein (1994) and Dupire (1997) and, although some new
theoretical proofs are given in Dempster and Richards (1999) and reliable numerical
methods are developed in this paper, its treatment here is mainly seen as a vehicle to
demonstrate the generality and efficiency of the LP valuation algorithm in Section 2.
In Section 4 results for FTSE 100 exotic American index options—fixed-strike Asian
puts—are presented to substantiate these claims. First, representative numerical results
for vanilla and constant volatility path-dependent American options are presented. Then
vanilla European and American options fitting the smile are studied in order to evaluate
potential pricing errors in fitting the local volatility surface used to finally price American
exotics. Conclusions are drawn in Section 5 and some directions of current and future
work are indicated.
On the continuation region, the value function satisfies the Black–Scholes parabolic
partial differential equation (PDE) LBS v + (∂v/∂t) = 0 for (x, t) ∈ R+ × [0, T ], where
the elliptic operator LBS := 21 σ 2 x 2 (∂ 2 /∂x 2 )+rx(∂/∂x)−r, since the discounted stopped
price process of the option is a martingale, while as soon as the process crosses into S,
v = ψ, and to preclude arbitrage LBS v + (∂v/∂t) ≤ 0. Hence we have
∂v
(2.1) −LBS v − ∧ (v − ψ) = 0
∂t
on the whole domain R+ × [0, T ), where ∧ denotes the pointwise minimum of two
functions. We now have a free-boundary formulation where v(x, t) = ψ(x, t) for (x, t)
on the optimal stopping or exercise boundary. We can remove any reference to the
optimal stopping boundary by formulating the problem in terms of (2.1) as a linear
order complementarity problem (OCP) using the log-transformed stock price variable
ξ := ln x, with respect to which the Black–Scholes operator is given by −Lv − (∂v/∂t)
with constant coefficient elliptic part L := 21 σ 2 (∂ 2 /∂ξ 2 ) + r − 21 σ 2 (∂/∂ξ ) − r and v
is now the option value as a function of ξ . The various inequalities carry through the
domain transformation and the new payoff function is given by ψ(ξ ) := K − eξ . It
will also be convenient to reverse time—to remaining time to maturity, for which we
will again use the symbol t for simplicity—so that the payoff function ψ becomes an
initial condition for the Black–Scholes PDE. The American put value function is then
the unique solution to
v(·, 0) = ψ
v ≥ ψ
(2.2) (OCP)
−Lv + ∂v
∂t ≥ 0
−Lv + ∂v ∂t ∧ (v − ψ) = 0 a.e. in R × [0, T ]
posed in a suitable vector lattice Hilbert space, which is a Hilbert space H with inner
product h·, ·i and partial order defined by a positive cone P such that for any points x
and y the maximum x ∨ y and the minimum x ∧ y exist in the given order (Cryer and
Dempster 1980; Borwein and Dempster 1989). Dempster and Hutton (1999) (see also
Jaillet, Lamberton, and Lapeyre (1990)) use another equivalent formulation of the value
function problem as a variational inequality (VI) to show the uniqueness of the solution
to (OCP) if the differential operator −L is coercive; that is, ∃α ∈ R+ s.t. hu, −Lui ≥
αkuk2 ∀u ∈ H . They show that the value function, as the unique solution to (OCP),
can be expressed as the unique solution of an abstract linear program given by
(2.3) (LP) inf hu, ci s.t. u ∈ F for any c > 0 a.e. on R × [0, T ],
u
where
∂u
(2.4) F := u ∈ H : u(·, 0) = ψ, u ≥ ψ, −Lu + ≥0 ⊂P
∂t
since the linear operator −L on the Hilbert space H is of type-Z; that is, v ∧ y = 0 (⇒
hv, yi = 0) ⇒ hv, −Lyi ≤ 0 ∀v, y ∈ H . This basic result giving equivalence between
160 m. a. h. dempster and d. g. richards
(OCP) and (LP) is an extension to the parabolic case of a result of Cryer and Dempster
(1980) for elliptic partial differential operators. To see this we shall need some notation.
Define the Sobolev space W m, p, µ (R2 ) as the space of functions u ∈ Lp (R2 , e−µ|x| dx)
whose weak derivatives of order not exceeding m ∈ N exist and are also in
Lp (R2 , e−µ|x| dx), for p ∈ [0, ∞] and µ ∈ (0, ∞). (Here | · | denotes the L1 norm
on R2 and dx denotes the Lebesgue measure on R2 , and it should be noted that the
extension of the results in the sequel to Rn+1 , for arbitrary n ∈ N, is completely straight-
forward.) Consider the Hilbert space H 1 (R2 ) := W 1, 2, µ (R2 ), for some fixed µ > 0, of
square integrable functions with square integrable derivatives defined on R2 . The Hilbert
space H 1 (R2 ) has as Banach dual the Sobolev space H −1 (R2 ) := W −1, 2, µ (R2 ), also a
Hilbert space of Radon measures, with which it may be identified. Consider the pairing
h·, ·i : H 1 × H −1 → R between dual spaces given by
Z
(2.5) hu, vi := u(ξ, t)v(ξ, t)e−µ(|ξ |+|t|) dξ dt,
R2
where we may interpret v ∈ H −1 as the density function of the Radon measure ele-
ment of the dual space H −1 of H 1 with respect to e−µ(|ξ |+|t|) dξ dt. Alternatively, we
may consider h·, ·i given by (2.5) as an inner product on the Hilbert space H 0 (R2 ) :=
L2 (R2 , e−µ|x| dx) by virtue of the canonical injections H 1 ,→ H 0 ,→ H −1 , see Baioc-
chi and Capelo (1984, p. 79). In this setting the partial differential operator L may be
interpreted either as a map H 1 → H −1 or as an operator on H 1 . Consider also the
bilinear form a(·, ·) : H 1 × H 1 → R given by
Z
σ2 σ2 σ2 ξ
(2.6) a(u, v) := uξ vξ − r− +µ uξ v + ruv
R2 2 2 2 |ξ |
× e−µ(|ξ |+|t|) dξ dt.
It can be shown that a given by (2.6), and hence −L, is coercive (see Jaillet et al.
(1990, p. 267), whose spaces L2 ([0, T ], Vµ ) and L2 ([0, T ], Hµ ) may be considered
restrictions respectively of our spaces H 1 and H 0 ). Then the Lions–Stampacchia theorem
(Baiocchi and Capelo 1984) implies that the solution to (VI) is unique. The formulation
(VI) is a type of classical physical problem, termed the (Stefan) obstacle problem, where
the payoff function ψ is the obstacle below which the solution cannot fall.
Next define, for a closed subset F ⊆ P ⊆ H of a vector lattice Hilbert space H with
positive cone P := {u ∈ H : u ≥ 0}, the least element problem
The least element v is denoted by LE(F ). Note that if it exists, the least element is
always unique since, if v1 and v2 are least elements of F , then v1 ≤ v2 and v2 ≤ v1 , so
from the vector lattice property v1 = v2 .
Proof. We first prove the equivalence between (OCP) and (LE), after making the
trivial domain extensions of the problem functions given above to set them in H 1 . Let
L denote the Laplace transform operator with respect to the measure e−µ|t| , so that, for
(ξ, λ) ∈ R2 , the Laplace transform û ∈ H 1 of a function v ∈ H 1 is defined by
Z ∞
(2.8) û(ξ, λ) := Lu(ξ, ·)(λ) := e−|λ|t u(ξ, t)e−µt dt.
0
As noted above, we have extended the temporal domain of our value functions v to [0, ∞)
as constant on (T , ∞), so that this generalized Laplace transform is well defined. L is
a linear operator and T is temporally homogeneous (i.e., it has time-independent coef-
ficients), and therefore commutes with the Laplace operator, so that taking the Laplace
transform of the operator T + ∂t∂ gives T L+L ∂t∂ . The Laplace transform of the first-order
time derivative is given by
Z ∞
∂v ∂v
(2.9) L (ξ, λ) := e−|λ|t (ξ, t)e−µt dt
∂t 0 ∂t
and v(ξ, 0) is given by the initial condition v(·, 0) = ψ (in backwards time).
162 m. a. h. dempster and d. g. richards
Now, note that the Laplace transform is positivity-preserving in the sense that v ≥ 0 ⇒
v̂ ≥ 0 a.e. on R2 . Then, writing the initial condition, constant in λ, as q̂(·, λ) :≡ −v(·, 0)
to agree with the notation of Borwein and Dempster (1989), (OCP) is equivalent to the
transformed order complementarity problem (OCP), \ also posed in H 1 , given by
v̂ ≥ ψ̂
\
(OCP) T v̂ + q̂ ≥ 0
(T v̂ + q̂) ∧ (v̂ − ψ̂) = 0 a.e. on R2 ,
It follows that v̂ is the unique solution to the least element problem (LE)d defined by
LE(F̂ ), where F̂ is defined by F̂ := {v̂ : v̂ ≥ ψ̂, T v̂ + q̂ ≥ 0}. Applying the inverse
Laplace transform L−1 to v̂ shows that v = L−1 v̂ solves both (LE), given by LE(F ),
and (OCP), as required. Indeed suppose the contrary: that there exists u ∈ F such that
u ≤ v, u 6= v. Then it follows since L is positivity preserving that û ∈ F̂ and û ≤ v̂,
û 6= v̂, a contradiction to v̂ = LE(F̂ ).
With this least element result, the LP equivalence is immediate: v is the least element
of F ⇐⇒ v ≤ u for all u ∈ F , and so hc, vi ≤ hc, ui for all u ∈ F and any vector
c > 0. Therefore u minimizes hc, vi over all v in F and is thus the solution to the abstract
linear program (LP). Restricting to the original problem domain yields the result. 2
It should be noted that the above proof depends on time running “backwards” since
otherwise we cannot substitute ψ for v(·, 0) in (2.9). The least element result tells us
that the linear constraint set lies within the positive cone translated so that its apex lies
at v. We can pick out the least element of the constraint set by minimizing hc, ui over
the set u ∈ F , where c > 0; specifically in R2 by minimizing the intercept of negatively
sloped lines defined by c0 u with normal c > 0 intersecting F .
Theorem 2.2 gives equivalence between (VI), (OCP), (LE), and (LP) for the American
put, since −L is coercive type Z (see Jaillet et al. 1990). It should be stressed that the
result is very general and applies to virtually any parabolic partial differential operator
with a temporally homogeneous coercive type Z elliptic part, and virtually any payoff
function. For example, it may be applied to the Black–Scholes operator −LBS directly
without prior logarithmic transformation of the space variable. A more delicate argument
involving step function coefficient approximation and a suitable passage to the limit can
be used to establish the results of Theorem 2.2 for time-dependent coefficient operators
(the details will appear elsewhere) which are required in this paper to handle the time-
varying nature of local volatility coefficients. Theorem 2.2 also suggests a simple way
to solve the equivalent problems numerically: by a suitable discretization the infinite-
dimensional abstract linear program (LP) reduces to an ordinary linear program with
solutions in Rn . Thus we next discretize the problem and consider efficient LP numerical
solution methods.
pricing american options fitting the smile 163
2.1. Computation
We shall approximate the value function by a function that is piecewise constant
on rectangular intervals between points in a regular lattice. Approximating the partial
derivatives by standard Crank–Nicolson finite differences (see, e.g., Wilmott et al. 1993)
we obtain a discrete form of (OCP) that can be rewritten in matrix form upon collapsing
the space index. A matrix is type-Z if it has nonnegative off-diagonal elements (see,
e.g., Borwein and Dempster 1989), which in the case of the matrix derived from the
discretized negative Black–Scholes operator occurs when |r − σ 2 /2| ≤ σ 2 /1ξ and can
be satisfied by adjusting the number of space steps I in the discretization (Dempster
and Hutton 1999). From this condition it can also be shown that this matrix is coercive
(Jaillet et al. 1990; Hutton 1995).
The LP formulation can be solved in backwards time either directly or iteratively
and the interested reader can find comparisons of solution methods in Hutton (1995) and
Dempster and Hutton (1999). Here we solve the discretized (LP) using time-stepping and
a simplified revised simplex algorithm that takes advantage of the tridiagonal structure
of the constraint matrix formed from standard Crank–Nicolson finite difference approx-
imations to produce a fast accurate direct solution method detailed in Dempster et al.
(1998) and Richards (1999). This procedure is suitable for any standard constant param-
eter Black–Scholes type formulation, but also yields significant computational savings
for valuation problems with volatility and drift parameters which are functions of time. It
incorporates a technique for the solution of problems with nonconstant constraint matrix
coefficients such as those involving the untransformed Black–Scholes PDE, which has
coefficients given by functions of the underlying asset price, or for exotic option pric-
ing problems, where the coefficients vary with the third variable representing the path-
dependency. In Dempster et al. (1998) results are presented for this updating procedure
which show that even for a general constraint matrix the procedure outperforms standard
commercial LP solvers by orders of magnitude.
To understand the exact computational savings of these simplex methods, first consider
the complexity of the vanilla American put option valuation problem after transformation
to the constant-coefficient Black–Scholes operator. At each time-step the maximum num-
ber of real variables which can enter the simplex basis is O(I ) and hence we have O(I )
iterations at each time step, where I is the number of points in the spatial discretization.
In fact, after the first few time steps—where the exercise boundary has greatest curvature
away from ln K (see Figure 4.3 later)—at most one new basic variable enters at each
time step. Far from maturity, calculations for several time steps may even utilize the
same basis. Each iteration requires O(n) operations to solve, where n ≤ I , giving O(I )
operations at each time step. Hence the space complexity of the algorithm is linear and
the total operation count is O(MI ), where M is the number of time steps.
For the basis factorization updating technique required by each simplex iteration (space
step) the calculations result in a similar complexity, but can be performed in three floating
point operations, although extra computation time is needed for the dynamic allocation
of the upper-lower (UL) factorization. Results for the constant coefficient method and
for the nonconstant coefficient updating technique are reported in Dempster et al. (1998),
along with results for a complete calculation of the full LU factorization at each iteration
to highlight the overheads of using general commercial solvers.
as linear programs through state augmentation, particularly for American discretely sam-
pled lookback and Asian options. Exotic option values are dependent on the underlying
stock price, (forward) time, and an additional “independent” variable that encapsulates
the required path information.
We now outline the formulation of a generic American exotic option in a discretely
sampled setting using the unifying framework of Wilmott et al. (1993). Denote by
V (S, M, t) the value function of the option with V : R+ × R+ × [0, T ] → R, where S
denotes the asset price and M denotes the value of a path-dependent variable, such as the
average in the case of Asian options, or the maximum/minimum for lookback options.
Assuming that the asset price is sampled on N occasions during the life of the option
with maturity T , denote by Mn the observed value of the augmented variable at the
sampling date tn , n = 0, . . . , N − 1, so that sampling begins at time 0 and t0 = 0. For
completeness, define tN := T . The variable Mn is a constant value throughout the period
[tn , tn+1 ), since no sampling takes place until time tn+1 . Effectively Mn is a parameter in
the formulation during this period and any randomness in the model is due to the asset
price process. The Black–Scholes PDE will thus be satisfied within the period with jump
conditions applied at sampling dates (see Wilmott et al. 1993 for more details). Across
a general sampling date tn the augmented variable is updated from a value Mn−1 just
prior to the date to a value Mn at the sample date. No-arbitrage arguments lead to the
jump condition V S, Mn−1 , tn− = V (S, Mn , tn ) n = 1, . . . , N − 1, where tn− is a time
immediately before the sampling date tn .
In the time interval [tn , tn+1 ) the European value V satisfies the augmented Black–
Scholes PDE defined by LBS + f (S, t)∂V /∂M + 2V /2t, where f (S, t) is a function
specified for the option of interest. We consider the final period [tN −1 , T ] and use
a dynamic programming algorithm to determine values for earlier periods. As in the
vanilla American put case treated above, the American exotic valuation domain in R3
can be partitioned into a continuation region CN and a stopping region SN and we can
establish the existence of an optimal exercise boundary (Dempster and Richards 1999).
To complete the formulation of the discretely sampled exotic value in the final period
we require a terminal condition V (S, MN−1 , T ) = ψ(S, MN −1 ) for all S, MN −1 ∈ R+
and boundary conditions in S at S = 0 and as S → ∞ which are option dependent
and are discussed in more detail in the cited paper. If again we log-transform the prim-
itive variables (ξ := ln S, ζN−1 := ln MN−1 ) and formulate the valuation problem
with fixed ζN−1 as an OCP with respect to the transformed operator L, we may define
a new partition with regions CN and SN . Thus the American exotic valuation prob-
lem in the final period may be formulated in terms of the transformed value function
V := V eξ , eζN −1 , t as the unique solution of the order complementarity problem of
the form (2.2) over R2 × (tN−1 , T ] involving payoff ψ(ξ, ζN −1 ) := max eζN −1 − eξ , 0
with V also denoting the option value as a function of ξ and ζN −1 . This puts us in a
framework equivalent to that of the vanilla American put of Section 2 but with the addi-
tional parameter ζN−1 , and hence we have equivalence to an abstract LP for each value of
the augmented variable ζN−1 ∈ (−∞, ∞). This problem must be solved for all possible
values of the parameter ζN−1 . Applying the jump conditions at tN −1 to obtain the termi-
−
nal value V (S, MN−2 , tN−1 ), the argument may be repeated for the period [tN −2 , tN −1 ]
and, by backwards recursion, eventually for the period [0, t1 ].
of equity options has exhibited variation both with the strike price—the volatility smile
(or skew)—and the option’s maturity—the volatility term structure. These effects high-
light the market’s deviation from the Black–Scholes assumption that the future asset
price has a constant variance lognormal conditional probability density.
Several approaches have been suggested in the literature to model this behavior. One
approach is to treat the volatility as a second stochastic factor with the aim of spec-
ifying the time-varying volatility from the model (Hull and White 1987; Clarke and
Parrott 1998). However this approach is difficult to fit to market data, is not arbitrage
free, and introduces an additional dimension to the pricing problem. An alternative one-
factor approach allows the volatility σ := σ (S, t) to be a variable that is both state and
time dependent. By starting from the market data and finding the local volatilities that
are consistent with the market—commonly termed an inverse problem (Andersen and
Brotherton-Ratcliffe 1997; Lagnado and Osher 1997; Bouchouev and Isakov 1999)—
this model can be made to price the market nearly exactly. The most popular structures
on which this local volatility is determined are binomial or trinomial trees, which allow
specification of nodal transitional probabilities to fit the smile (Rubinstein 1994; Der-
man and Kani 1998). All methods used to fit market data are prone, however, to the
same instabilities. Generally the data can imply unreasonable (e.g., negative or large)
values of the local volatility, which may create negative transitional probabilities and
allow arbitrage possibilities in the model. The general inverse problem is ill-posed since
the number of volatility parameters to be found far outnumbers the limited number of
available option prices in the market. Therefore, it is often assumed that a continuum of
European call option prices C(K, T ) are available for all strikes and maturities. Alterna-
tively, recent papers have implemented regularization methods (Jackwerth and Rubinstein
1996; Bodurtha 1997; Lagnado and Osher 1997; Coleman, Li, and Verma 1999) to make
the inverse problem stable.
All approaches to fitting the smile suffer from the consequences of inconsistent data
and will not price correctly all options—particularly those far out of the money—in the
face of these data problems. The modeling approach used in this paper is not claimed
to be the most accurate or efficient, but it is quick and highlights the versatility of
the LP pricing method in the face of a degenerate ill-posed problem with nonconstant
coefficients.
density function (pdf) p(s, T |S, t) of the underlying asset S having value s at time T
given that the asset price at time t is S as
Z ∞
(3.11) C(S, t; K, T ) = P (t, T ) p(s, T |S, t)(s − K)+ ds,
0
RT
where t ≤ T and the discount factor P (t, T ) := exp − t r(u) du .
Breeden and Litzenberger (1978) showed that the European call option price and this
conditional probability density were related as p(K, T |S, t) = P (t, T )−1 ∂ 2 C(S, t; K, T )/
∂K 2 by differentiation of (3.11). The function p(K, T |S, t) is also the Green’s function
(or fundamental solution) of the Black–Scholes PDE. Thus it satisfies this PDE with
terminal condition (t = T ) p(K, T |S, t) = δ(S − K), where δ(·) is the Dirac delta func-
tion. Since C is assumed known, this density function can be found from the idealized
market data. The function p(K, T |S, t) can also be shown to satisfy the Fokker–Planck
(or forward Kolmogorov) PDE utilizing the following theorem (see, e.g., Jackwerth and
Rubinstein 1996).
Theorem 3.1. The conditional pdf p(y, τ |x, t) of a general stochastic process X(t)
where t ≥ 0 given by dX(t) = µ(X, t) dt + σ (X, t) dW(t) satisfies the Fokker–Planck
or forward Kolmogorov equation
∂p(y, τ |x, t) ∂ (µ(y, τ )p(y, τ |x, t)) 1 ∂ 2 σ 2 (y, τ )p(y, τ |x, t)
(3.12) + − = 0,
∂τ ∂y 2 ∂y 2
for fixed (x, t) ∈ R × R+ with initial condition p(y, t|x, t) = δ(x − y).
Corollary 3.2. The transitional probability density function p(S 0 , T |S, t) of the
stock price process dS/S = r(t) dt + σ (S, t) dW satisfies the PDE
Corollary 3.3. Given that the underlying asset price process in Corollary 3.2 the
price of a European call option C(S, t; K, T ) solves the partial differential equation
∂C σ (K, T )2 K 2 ∂ 2 C ∂C
(3.14) = − r(T )K
∂T 2 ∂K 2 ∂K
with boundary condition C(S, t; S, t) = 0 (see also Derman and Kani 1998).
3.2. Computation
The local volatility function σ (S, t) can be fully determined from (3.14) since all other
terms in the equation can be found from market data. However, since the inverse problem
is ill-posed there may not be a unique local volatility functional that fits the market data.
To obtain an approximately arbitrage-free pricing algorithm the implied volatility (or,
alternatively, call price) data available from the market must be fitted exactly. To this
pricing american options fitting the smile 167
end we apply cubic splines to approximate the implied volatility surface. The ease with
which approximations to first and second derivatives can be obtained from a spline fit
to market data is a major advantage in reducing the computational complexity of our
method since the construction of a length N cubic spline is the solution of a tridiagonal
system and hence an O(N) calculation. The complexity of the problem can be further
reduced by precomputing derivative information, but at the expense of additional memory
requirements.
We outline now the approach used to obtain a consistent local volatility surface that
fits the market-implied volatility data for European call options. This will be used in the
linear programming approach of Section 2 to price American exotic options consistent
with the volatility smile. Since we consider the implied volatility nonconstant in the
underlying diffusion, assume that the quoted European call prices are Black–Scholes
prices with implied volatilities as an additional parameter and define the call prices in
terms of the implied volatility ν(K, T ) for the call option of strike K and maturity T as
C(K, T ) := CBS (S, t; K, T , ν(K, T )), where CBS is the Black–Scholes European call
price. Using this formulation we can write the local volatility in terms of derivatives of
the Black–Scholes implied volatility as follows.1
Theorem 3.4. Let the asset price S follow the diffusion process in Corollary 3.2.
Then the local volatility function σ (S, t) consistent with the arbitrage-free European
call prices is given uniquely, in the absence of dividends, by
−1
∂C ∂C ∂ 2C
(3.15) σ 2 (K, T ) = 2 + r(T )K K2
∂T ∂K ∂K 2
with S = K. In terms of the implied volatility function ν(K, T ) this can be written as
(3.16)
∂ν 1ν ∂ν
2 γ ∂T + 2γ + r(T )K ∂K
σ 2 (K, T ) = ,
1 ∂ν 1 ∂ν 1 ∂2ν γ ∂ν
K2 ν ∂K ((T − t) − d1 ) − Kγ −d1 ∂K γ− K + ∂K 2
γ + K ∂K
where
ln(S/K) + r + 21 ν(K, T )2 γ 2 √
d1 := and γ = T − t.
ν(K, T )γ
We will use (3.16) to obtain the local volatilities. For our method the implied volatili-
ties are relatively stable and the expression for the local volatility involves relatively few
1
This result was derived following discussions with S. H. Babbs (of Bank One) and has since been inde-
pendently presented in Andersen and Brotherton-Ratcliffe (1997).
168 m. a. h. dempster and d. g. richards
j
X
(3.17) P (S, t; Kj +1 , T ) ≈ P (t, T ) (Kj +1 − Ki )p(Ki )1K
i=0
j −1
X
= P (t, T ) (Kj +1 − Ki )p(Ki )1K + P (t, T )(Kj +1 − Kj )p(Kj )1K,
i=0
which given that the P (S, t; Ki , T ) are known for all i ≤ j (from put-call parity) allows
us to find a value of the discrete probability p(Kj ) consistent with the market data. If
this probability is also negative, its value is set to zero. The other possible inconsistency
in the data occurs when the numerator of (3.16) is negative, when we set σ 2 (K, T ) to
zero.
pricing american options fitting the smile 169
4. NUMERICAL RESULTS
In this section we present empirical results for the procedure outlined in Section 3 for
pricing American options consistent with the observed market volatility smile, together
with benchmark results for their constant volatility equivalents. To facilitate comparisons
we use as our underlying implied volatility surfaces market data appearing previously
in the literature and we price European, American, and exotic options with respect to it.
All solution times quoted are for calculations on an IBM RS6000/590 workstation with
1 GB RAM running under AIX 4.3, although only a small proportion of this memory is
utilized. Results are quoted in Dempster et al. (1998) for solution on a Pentium II 400
Mhz PC which gives significant speed-ups over those presented here for most levels of
domain discretization.
Table 4.1
Comparison of Tridiagonal Simplex Solvers with PSOR
Tridiagonal simplex
Space Constant Recalculation
steps coefficients UL update simplex PSOR
75 0.02 0.05 0.10 0.07
150 0.05 0.08 0.17 0.13
300 0.10 0.14 0.33 0.27
600 0.19 0.24 0.65 1.25
1200 0.38 0.47 1.26 6.37
2400 0.77 1.00 2.47 37.55
4800 1.61 2.24 5.12 255.06
9600 3.71 5.09 10.77 1856.91
Solution CPU times in seconds for M := 1000.
2
See Dempster and Hutton (1999) for comparison of the commercial IBM Optimization Subroutine Library
(OSL) with PSOR, interior-point, and simplex methods.
170 m. a. h. dempster and d. g. richards
Table 4.2
Comparison of Lookback Valuation Results for the Tridiagonal Method against the
Similarity Transformed Method
Tridiagonal
Similarity
200 × 300 × 200 400 × 600 × 400 800 × 1200 × 800
Transform
Scheme Solution Time Solution Time Solution Time Solution
A 10.532 6.46 10.550 53.88 10.555 477.32 10.5
B 9.445 6.69 9.454 59.05 9.457 508.95 9.5
C 8.114 6.59 8.116 59.32 8.116 489.60 8.1
pricing american options fitting the smile 171
Figure 4.1. Implied volatility surface for European call options on the FTSE 100 index.
described in Section 3. We also assume that a constant rate of interest, r = 10%, applies
throughout for all maturities.
As described in Section 3 the pricing procedure occurs in two stages. The first—
calibrating the local volatility—uses the estimates of the implied volatility derivatives to
calculate the surface σ (S, t) for use in pricing. In the sequel this calibration takes place
on the same mesh that is used in the pricing procedure. There are significant compu-
tational savings to be made by calculating the local volatilities on a coarser mesh and
performing some type of interpolation between the calibrated values. The initial cali-
bration of the surface is computationally expensive. To calculate a volatility surface for
a discretization of 800 time-steps and 200 space-steps takes approximately 20 seconds,
although the time required for subsequent valuations on the same grid is a fraction of
this if the surface is stored for future use. We quote solution times for the option val-
uation only, assuming that the precomputed local volatility surface is in memory. After
calibrating the volatility surface, the options are valued using the UL update algorithm
of Section 2 for options with American exercise and a simple linear equation solver
for European options. At each time-step for the former, the basis decomposition was
calculated and the UL update applied only when new variables entered the basis.
Empirical tests showed that the underlying discretization error due to the Crank–
Nicolson numerical approximation of the derivatives was less than 8 basis points on a
grid of 3200 time steps and 200 space steps; this must be accounted for in any accuracy
comparisons.
172 m. a. h. dempster and d. g. richards
Table 4.3
Pricing errors ×100 of FTSE 100 European Call Options Fitting the Smile
2975 158.84 158.87 3.50 189.08 189.09 1.00 221.23 221.25 2.30
3025 114.92 114.88 4.40 148.82 148.80 1.80 181.84 181.82 1.70
3075 74.54 74.52 1.63 112.51 112.52 0.70 146.10 146.11 1.20
3125 44.43 44.46 2.89 80.69 80.72 3.32 114.15 114.19 4.00
3175 23.50 23.51 1.09 54.92 54.91 0.54 86.11 86.13 1.93
3225 9.77 9.79 2.34 35.13 35.08 4.88 62.02 61.98 3.99
3275 4.15 4.17 2.10 21.24 21.21 2.62 42.85 42.84 0.94
3325 1.09 1.06 2.79 11.85 11.82 3.48 27.83 27.80 3.36
M = 3200, I = 200. Actual value given by the Black–Scholes call pricing equation.
pricing american options fitting the smile 173
Figure 4.2. Local volatilities σ (S, t) for FTSE 100 index options on a truncated strike
domain. Parameters M = 200, I = 200 for the full strike range.
The results for the FTSE 100 index in Table 4.4 show American put valuations for the
LP tridiagonal solver. As a benchmark we use the original LP valuation on a very fine
solution mesh using the at-the-money BS implied volatility. American option values are
higher than the corresponding constant volatility values to within the numerical tolerance
previously evaluated, with the LP solution time being 0.6 seconds. Since both solution
algorithms converge to within the same numerical accuracy, the discrepancies between
the smile-fitting and at-the-money implied values are due to the volatility surface. It was
noted in Section 2 that the convex shape of the optimal exercise boundary for the vanilla
American put problem was useful in increasing the efficiency of the tridiagonal solver.
Figure 4.3 highlights the reason why the volatility smile fitting option value is different
from its vanilla counterpart by illustrating the shifted nonconvex shape of its optimal
exercise boundary. When we take account of local volatility, the exercise boundary is
no longer a convex function of the asset price, but is shifted horizontally by changes in
local volatility. Although this poses no problem for the accuracy of the pricing algorithm,
it does radically affect the realized option price. Similar results for the S&P 500 index
options are given in Dempster and Richards (1999).
Figure 4.3. American put optimal exercise boundary for the smile fit compared with the
1-factor LP method using ATM BS implied volatility. Parameters: M = 1000, I = 3000,
S0 = 3129.5, K = 3125, σBS = 15.89%. Dotted lines show market European call
maturities.
Black–Scholes at-the-money implied volatility for the options in question. The columns
labeled “Smile value” illustrate the results obtained by fitting the volatility smile and
term structure.
As can be seen, the option price fitting the volatility smile is significantly different
from the constant volatility price. Unlike the vanilla American put the American fixed-
strike Asian option results on the FTSE 100 index (Table 4.5) show a mixed effect. For
a low arithmetic average sampling rate the smile-fitting value is less than the Black–
Scholes implied value, but the converse relationship holds for higher sampling rates.
This effect is likely due to the time-to-maturity variability of implied volatility at strike
2975 depicted in Figure 4.1.
5. CONCLUSION
After surveying earlier results, we applied a fast accurate linear programming valua-
tion algorithm to pricing American exotic options fitting the volatility smile implied by
the market prices of vanilla European call options. We have demonstrated first that the
basic Crank–Nicolson finite difference methods have low discretization error and that
the quoted vanilla options are accurately priced by the fitted local volatility surface.
Subsequently we have seen that, due to local volatility effects on the computed optimal
exercise boundary, prices of American options fitted to the smile differ significantly from
pricing american options fitting the smile 175
Table 4.4
American Put Option Valuation Results Fitting the FTSE 100 Volatility Smile
those with constant volatilities. Finally, we have seen similar effects for American exotic
options, as represented by discretely sampled fixed-strike Asian options.
Current research extends the testing of these methods to lookbacks and barriers, includ-
ing both digitals and knock-in and knock-out features for Asians and lookbacks. An inter-
esting area of related research involves the Kalman filtering of local volatility surfaces—
as for example computed in this paper—from one market epoch (day) to the next in order
to achieve better long-run hedging. Another line of our current research with PDE-based
valuation methods concerns wavelet basis techniques for discontinuous option payoffs,
including barriers (Dempster et al. 1999), and high-dimensional Bermudan and American
fixed income derivatives (see Dempster and Hutton (1997b)).
Table 4.5
Discretely Sampled American Fixed-Strike Asian Put Option Results Fitting the
FTSE 100 Smile
2 Samples 12 Samples
Implied Smile Implied Smile
M I J value value M I J value value
400 200 200 0.360 0.295 270 200 200 3.134 3.710
800 200 200 0.360 0.296 540 200 200 3.135 3.759
800 400 400 0.353 0.278 540 400 400 3.127 3.737
Parameters: K = 2975, S0 = 3129.5 and T = 0.211 years. Solution time for
M = 200, I = J = 200 is approximately 18 seconds.
176 m. a. h. dempster and d. g. richards
references
Dupire, B. (1997): Pricing and Hedging with Smiles; in Mathematics of Derivative Securities,
M. A. H. Dempster and S. R. Pliska, eds. Cambridge: Cambridge University Press, 103–
111.
Harrison, J. M., and D. Kreps (1979): Martingales and Arbitrage in Multi-Period Security
Markets, J. Econ. Theory 20, 381–408.
Harrison, J. M., and S. R. Pliska (1981): Martingales and Stochastic Integrals in the Theory
of Continuous Trading. Stoch. Process. Appl. 11, 215–260.
Hull, J., and A. White (1987): The Pricing of Options on Assets with Stochastic Volatilities,
J. Financial Quant. Anal. 3, 281–300.
Hutton, J. P. (1995): Fast Pricing of Derivative Securities, Ph.D. thesis, Department of
Mathematics, University of Essex.
Jackwerth, J. C., and M. Rubinstein (1996): Recovering Probability Distributions from
Option Prices, J. Finance 51, 62–74.
Jaillet, P., D. Lamberton, and B. Lapeyre (1990): Variational Inequalities and the Pricing
of American Options, Acta Appl. Math. 21, 263–289.
Lagnado, R., and S. Osher (1997): A Technique for Calibrating Derivative Security Pricing
Models: Numerical Solution of an Inverse Problem, J. Computat. Finance 1, 13–25.
Richards, D. G. (1999): Pricing American Exotic Options, Ph.D. thesis, Judge Institute of
Management Studies, University of Cambridge.
Rubinstein, M. (1994): Implied Binomial Trees, J. Finance 49, 771–818.
Wilmott, P., S. Howison, and J. Dewynne (1993): Option Pricing: Mathematical Models
and Computation. Oxford: Oxford Financial Press.