Next Article in Journal
Some Properties of Operator Valued Frames in Quaternionic Hilbert Spaces
Next Article in Special Issue
ADI Method for Pseudoparabolic Equation with Nonlocal Boundary Conditions
Previous Article in Journal
Deep Neural Network-Based Footprint Prediction and Attack Intention Inference of Hypersonic Glide Vehicles
Previous Article in Special Issue
Reclamation of a Resource Extraction Site Model with Random Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Alternative Numerical Scheme to Approximate the Early Exercise Boundary of American Options

by
Denis Veliu
1,†,
Roberto De Marchis
2,†,
Mario Marino
3,† and
Antonio Luciano Martire
4,*,†
1
Departament of Finance-Banking, Metropolitan University of Tirana, 1000 Tirana, Albania
2
MEMOTEF Department, Sapienza University of Rome, 00185 Rome, Italy
3
DEAMS “Bruno De Finetti”, University of Trieste, 34127 Trieste, Italy
4
Department of Business Economics, Roma Tre University, 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Mathematics 2023, 11(1), 187; https://doi.org/10.3390/math11010187
Submission received: 25 November 2022 / Revised: 21 December 2022 / Accepted: 26 December 2022 / Published: 29 December 2022

Abstract

:
This paper deals with a new numerical method for the approximation of the early exercise boundary in the American option pricing problem. In more detail, using the mean-value theorem for integrals, we provide a flexible algorithm that allows for reaching a more accurate numerical solution with fewer calculations rather than other previously described methods.

1. Introduction

Volterra integral equations have practical applications in many fields, including biology, medicine, finance and engineering. Although various methods are available to find analytical solutions for these integral equations, in most cases, finding a closed-form solution may be unfeasible. To this end, several numerical methods have been developed in recent years (see, e.g., [1,2]). In the present work, we deal specifically with the Volterra integral equation of the second kind, defined below
ϕ ( t ) = f ( t , ϕ ( t ) ) + 0 t k ( t , s ) ψ ( t , s , ϕ ( t ) , ϕ ( s ) ) d s ,
with t I = [ 0 , T ] . Such an equation in which the integrand depends both on ϕ ( t ) and ϕ ( s ) is also a nonstandard Volterra integral equation ([3]).
In the field of mathematical finance, Equation (1) and its solution are a matter of interest in order to determine the value of an American financial option. More generally, a financial option is a derivative contract allowing for the holder to buy or sell an underlying financial asset at a fixed price, namely, the strike price. If the holder can buy or sell the underlying asset only at the expiration date, the option is called European; conversely, an American financial option provides the holder with the possibility to also exercise its right before the expiration date. Depending on the value of the underlying asset, it may become optimal for the holder to exercise its right before the expiration time. Such an asset value is the optimal exercise price, and by considering the optimal exercise times during the option duration, we can achieve a collection of optimal exercise prices, namely, the optimal exercise boundary. From a mathematical perspective, the definition of optimal exercise boundary identifies a free boundary problem whose solution must be determined numerically. For instance, Kim [4] showed that the early exercise boundary of an American put option is the solution of the integral equation having the form (1).
Among others, the authors in [5,6], Barone-Adesi, Whaley [7], Bunch et al. [8], Aitsahlia and Lai [9,10] considered integral equations and/or the optimal stopping problem for pricing American options. The integral representation proposed by Kim gives a suggestive economic interpretation, but a closed-form solution for these kind of problems is not available. In particular, Kim’s equation ([4]) contains a double integral due to the presence of a cumulative normal distribution term. To avoid such a problem, the authors in [11,12] suggested a transformation of the original equation in order to reduce it to an one-dimensional integral equation and afterwards consider its numerical solution. The early exercise boundary shows some kind of singularity near expiration (see, e.g., [13,14]). Considering the numerical approximation problem of the early boundary exercise for the American put, there have been more efforts in recent years. The authors in [15] proposed a four-point extrapolation scheme. Bunch and Johnson [8] proposed a modified version of the two-point extrapolation scheme of Geske and Johnson. Broadie and Detemple [16] suggested a lower and upper bound approximation method. The authors in [17] showed the exponential function method combined with a three-point Richardson extrapolation. Starting from an integral equation representation for the early boundary exercise, the authors in [12] used an iteration method. The authors in [18] proposed trapezoidal formulas approximations combined with the Newton–Raphson iteration. In [19], the authors solved a variational inequality representation of the American put option pricing problem by applying the projected successive over-relaxation method. The authors in [20] suggested an analytical expression to the value of American put options and their optimal exercise boundary. The authors in [21] derived a local iterative numerical scheme based on a solution of the integral equation proposed by [13]. More recently, Nedaiasl et al. [22], starting from the cited one-dimensional reformulation of Kim’s integral equations, obtained the numerical solution by using a modified version of the Nyström method in order to take into account the singularity near expiry presented by the early exercise boundary. Besides the method of Zhu [20] and others involving semianalytical approximations that are not cited for the brevity of treatment, in the American option pricing literature, numerical methods aim to numerically solve the integral equations describing the early exercise boundary. We proceed in the latter direction.
More recently, some research studies applied the integral equation approach to deal with particular financial evaluations. For instance, the authors in [23] exploited the Mellin’s transform aiming to provide an integral representation for the barrier option price. In [24], the authors gave an integral equation representation for a two-free-boundaries problem arising in the American better-of option on two assets.
In the present work, using a very simplified approach, we straightforward solve the nonstandard Volterra integral equation of the form (1). In more detail, using the mean-value theorem for integrals, we provide a flexible algorithm that allows for reaching a more accurate numerical solution with fewer calculations rather than other previously described methods.
The paper is organized as follows. In Section 2, we propose our numerical method for nonstandard Volterra integral equations. In Section 3, such a method is applied to the case of the American put. In Section 4, we present some numerical results confirming the accuracy of the proposed method. Lastly, in Section 5 we show the conclusions.

2. A New Numerical Scheme for Nonstandard Volterra Integral Equations

Let us fix interval I = [ 0 , T ] where T > 0 and consider the following nonstandard Volterra integral equation:
ϕ ( t ) = f ( t , ϕ ( t ) ) + 0 t k ( t , s ) ψ ( t , s , ϕ ( t ) , ϕ ( s ) ) d s ,
with t I = [ 0 , T ] , where the kernel function k is continuous in I × I , ψ is continuous in I × I × R × R .
Let n be a positive integer, and let us consider the following partition Γ of the interval [ 0 , T ] into n intervals of equal length Δ = T / n :
0 = t 0 < t 1 < t 2 < < t n 1 < t n = T .
Let us consider Equation (2). By means of the additive properties for integrals, we can rewrite for each t i , i = 1 , 2 , , n Equation (2) in the following way:
ϕ ( t i ) = f ( t i , ϕ ( t i ) ) + m = 1 i t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s .
The following result holds.
Proposition 1. 
Let:
(a) 
Function k ( t , s ) be continuous such that it never changes sign for all ( t , s ) I × I . Let L > 0 be a constant such that | k ( t , s ) | L for each ( t , s ) I × I .
(b) 
Function ϕ ( s ) be continuous in I.
(c) 
Function ψ ( t , s , x , y ) be continuous in I × I × R × R .
Then, for each m = 1 , 2 , i , the following approximation holds:
t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s = = 1 2 ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s + + 1 2 ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s + ϵ m , i ,
where error | ϵ m , i | satisfies | ϵ m , i | h m / ( 2 n ) , with h m a constant, and ϵ m , i 0 for n .
Proof. 
Using the mean value theorem for integrals, it follows that
t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s = ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) t m 1 t m k ( t i , s ) d s ,
where ξ m is, for each m, a number belonging to the interval [ t m 1 , t m ] . It follows
t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s = 1 2 ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s + + 1 2 ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s 1 2 ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s + 1 2 ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s + 1 2 ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) t m 1 t m k ( t i , s ) d s + + 1 2 ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) t m 1 t m k ( t i , s ) d s .
Let us rearrange the elements in (6) in the following way:
t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s = 1 2 ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s + + 1 2 ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s + ϵ m , i A + ϵ m , i B ,
where we have set
ϵ m , i A = 1 2 ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s
and
ϵ m , i B = 1 2 ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s .
Then, it follows that
| ϵ m , i A | = 1 2 | ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s | = 1 2 | ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) | · | t m 1 t m k ( t i , s ) d s | 1 2 | ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) | · t m 1 t m | k ( t i , s ) | d s 1 2 | ψ ( t i , ξ m , ϕ ( t i ) , ϕ ( ξ m ) ) ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) | · T L n = h m , i A 2 n ,
where h m , i A is a positive constant because the argument in modulus is a number for each m and for any fixed t i .
As n , it follows ϵ m , i A 0 , and the same considerations apply to ϵ m , i B with a positive constant h m , i B . If we consider the sum ϵ m , i = ϵ m , i A + ϵ m , i B , it follows ϵ m , i 0 , as n .
In addition, it follows easily that | ϵ m , i | h m , i / ( 2 n ) , with h m , i = h m , i A + h m , i B . □
Remark. 
If we consider Equation (4), it follows, through Proposition 1 for each fixed t i , that
m = 1 i t m 1 t m k ( t i , s ) ψ ( t i , s , ϕ ( t i ) , ϕ ( s ) ) d s = = 1 2 m = 1 i [ ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) , d s + + ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) , d s ] + ϵ i ,
where
ϵ i = m = 1 i ϵ m , i .
From Proposition 1, it is easy to check that: | ϵ i | H / ( 2 n ) , with H = m = 1 i h m , i . It is straightforward to notice that ϵ i 0 for n .
We provide the following algorithm in order to find the numerical solution.
Step 1. 
Let n be a positive integer. Let us consider the following partition Γ of interval [ 0 , T ] into n intervals of equal length Δ = T / n :
0 = t 0 < t 1 < t 2 < < t n 1 < t n = T .
If Proposition 1 is verified, and from the additive property for integrals, for each t i , with i = 0 , 1 , 2 , , n , we can rewrite Equation (2) in the following way:
ϕ ( t i ) = f ( t i , ϕ ( t i ) ) + 1 2 [ m = 1 i ψ ( t i , t m 1 , ϕ ( t i ) , ϕ ( t m 1 ) ) t m 1 t m k ( t i , s ) d s + + ψ ( t i , t m , ϕ ( t i ) , ϕ ( t m ) ) t m 1 t m k ( t i , s ) d s ]
Step 2. 
We observe that
ϕ ( t 0 ) = f ( t 0 , ϕ ( t 0 ) ) + t 0 t 0 k ( t 0 , s ) ψ ( t 0 , s , ϕ ( t 0 ) , ϕ ( s ) ) d s = f ( t 0 , ϕ ( t 0 ) ) .
Then, for i = 0 , , n we consider the following nonlinear system.
ϕ ( t 0 ) = f ( t 0 , ϕ ( t 0 ) ) , ϕ ( t 1 ) = f ( t 1 , ϕ ( t 1 ) ) + 1 2 [ ψ ( t 1 , t 0 , ϕ ( t 1 ) , ϕ ( t 0 ) ) t 0 t 1 k ( t 1 , s ) d s + + ψ ( t 1 , t 1 , ϕ ( t 1 ) , ϕ ( t 1 ) ) t 0 t 1 k ( t 1 , s ) d s ] , ϕ ( t 2 ) = f ( t 2 , ϕ ( t 2 ) ) + 1 2 [ m = 1 2 ψ ( t 2 , t m 1 , ϕ ( t 2 ) , ϕ ( t m 1 ) ) t m 1 t m k ( t 2 , s ) d s + + ψ ( t 2 , t m , ϕ ( t 2 ) , ϕ ( t m ) ) t m 1 t m k ( t 2 , s ) d s ] , ϕ ( t n ) = f ( t n , ϕ ( t n ) ) + 1 2 [ m = 1 n ψ ( t n , t m 1 , ϕ ( t n ) , ϕ ( t m 1 ) ) t m 1 t m k ( t n , s ) d s + + ψ ( t n , t m 1 , ϕ ( t n ) , ϕ ( t m ) ) t m 1 t m k ( t n , s ) d s ] .
The above nonlinear system is solved with a numerical method that provides the approximate solution { ϕ ˜ ( t 0 ) , ϕ ˜ ( t 2 ) , , ϕ ˜ ( t n ) } .

3. Approaching the American Put Pricing Problem

Let us assume the asset price process { S ( t ) , t 0 } following the log-normal distribution of the form
d S t = ( r δ ) S t d t + σ S t d W t ,
where W t is the standard Wiener process, r is the constant interest rate, and δ is the dividend yield. For the aim of the present work, we follow [25] assuming that the volatility term, σ , is constant. The latter is a simplifying hypothesis, and Equation (11) can be generalized considering nonconstant volatility, as in [26]. In this case, the boundary problem is represented by a multidimensional Volterra integral equation involving more complicated integrals (see, e.g., [27]). We leave the nonconstant volatility problem to future research works.
Let us consider the interval I = [ 0 , T ] . We denote with Φ ( · ) the standard cumulative normal distribution function. Let us denote with B ( t ) the early exercise boundary of an American put option, where t I . Let P the price of an American put. The authors in [28] proved the existence and uniqueness of the pair ( P , B ) , and the continuity and monotonic behaviour of B . In [4], the authors showed that B ( t ) , as a function of time to expiration, satisfies the following weakly singular Volterra integral equation:
K B ( t ) = K e r t Φ ( log ( B ( t ) K ) + ( r δ σ 2 2 ) t σ t ) + B ( t ) e δ t Φ ( log ( B ( t ) K ) + ( r δ + σ 2 2 ) t σ t ) + + K r 0 t e r ( t s ) Φ ( log ( B ( t ) B ( s ) ) + ( r δ σ 2 2 ) ( t s ) σ t s ) d s + δ B ( t ) 0 t e δ ( t s ) Φ ( log ( B ( t ) B ( s ) ) + ( r δ + σ 2 2 ) ( t s ) σ t s ) d s ,
where K is the exercise price.
In Equation (12), let us consider the two integrals and in particular the two functions k 1 ( t , s ) = e r ( t s ) and k 2 ( t , s ) = e δ ( t s ) . It is easy to check that k 1 ( t , s ) and k 2 ( t , s ) , for any ( t , s ) I × I , satisfy Proposition 1 with L = 1 .
We apply the algorithm presented in Section 2. Let us consider the following result.
Proposition 2. 
Assume that the asset price, S t , follows a log-normal distribution process of the form d S t = ( r δ ) S t d t + σ S t d W t , in which W t is the standard Wiener process.
Let B ( t ) be the solution of Integral Equation (12). Then, B ( t ) is a continuously differentiable function on ( 0 , T ] and
  • for r δ :
    lim t 0 B ( t ) = K ;
  • for r > δ :
    lim t 0 B ( t ) = r δ K .
Proof. 
See [4]. □
For our purposes, let us rewrite Equation (12) in contract form:
B ( t ) = K f ( t , B ( t ) ) r K 0 t e r ( t s ) Φ ( d 2 ( t , s , B ( t ) , B ( s ) ) ) d s + + δ B ( t ) 0 t e δ ( t s ) Φ ( d 1 ( t , s , B ( t ) , B ( s ) ) ) d s ,
where
  • f ( t , B ( t ) ) = K e r t Φ ( log ( B ( t ) K ) + ( r δ σ 2 2 ) t σ t ) B ( t ) e δ t Φ ( log ( B ( t ) K ) + ( r δ + σ 2 2 ) t σ t ) ;
  • d 1 ( t , s , B ( t ) , B ( s ) ) = log ( B ( t ) B ( s ) ) + ( r δ + σ 2 2 ) ( t s ) σ t s ;
  • d 2 ( t , s , B ( t ) , B ( s ) ) = log ( B ( t ) B ( s ) ) + ( r δ σ 2 2 ) ( t s ) σ t s .
Let n be a positive integer. Let us consider the following partition Γ of the interval [ 0 , T ] into n intervals of equal length Δ = T / n :
0 = t 0 < t 1 < t 2 < < t n 1 < t n = T .
For each t i , with i = 0 , 1 , 2 , , n , using the additive property of integral, let us rewrite Equation (15) in this way:
B ( t i ) = K f ( t i , B ( t i ) ) + m = 1 i [ r K t m 1 t m e r ( t i s ) Φ ( d 2 ( t i , s , B ( t i ) , B ( s ) ) ) d s + + δ B ( t ) t m 1 t m e δ ( t i s ) Φ ( d 1 ( t i , s , B ( t i ) , B ( s ) ) ) d s ] .
Proposition 3. 
For the two integrals appearing in Equation (16), the following equalities hold:
t m 1 t m e r ( t i s ) Φ ( d 2 ( t i , s , B ( t i ) , B ( s ) ) ) d s = = 1 2 Φ ( d 2 ( t i , t m 1 , B ( t i ) , B ( t m 1 ) ) t m 1 t m e r ( t i s ) d s + + 1 2 Φ ( d 2 ( t i , t m , B ( t i ) , B ( t m ) ) t m 1 t m e r ( t i s ) d s + ϵ m , i A
and
t m 1 t m e δ ( t i s ) Φ ( d 1 ( t i , s , B ( t i ) , B ( s ) ) ) d s = = 1 2 Φ ( d 1 ( t i , t m 1 , B ( t i ) , B ( t m 1 ) ) ) t m 1 ζ m e δ ( t i s ) d s + + 1 2 Φ ( d 1 ( t i , t m , B ( t i ) , B ( t m ) ) ) t m 1 t m e δ ( t i s ) d s + ϵ m , i B ,
where ϵ m , i A and ϵ m , i B are the errors as defined in Proposition 1.
Proof. 
The result follows immediately from the continuity of Φ and B and by means of Proposition 1. □
Equation (12) represents a nonstandard Volterra integral equation of the second kind having the form (1). We solve it by using of the method described in Section 2. By applying this method, when s = t i , the indeterminate form
Φ ( log ( B ( t i ) B ( t i ) ) + ( r δ σ 2 2 ) ( t i t i ) σ t i t i ) = Φ ( 0 0 )
arises. Observing that
lim s t Φ ( log ( B ( t ) B ( s ) ) + ( r δ σ 2 2 ) ( t s ) σ t s ) = 1 2 ,
we can define the continuous function
Φ ( log ( u ( t ) u ( s ) ) + ( r δ σ 2 2 ) ( t s ) σ t s ) = Φ ( log ( u ( t ) u ( s ) ) + ( r δ σ 2 2 ) ( t s ) σ t s ) 0 s < t 1 2 t = s .
The same considerations apply to
Φ ( log ( u ( t ) u ( s ) ) + ( r δ + σ 2 2 ) ( t s ) σ t s ) .
Using aforementioned notations, from the numerical approximation of the early boundary exercise, the price of the American put may be found by means of the following formula:
P ( t , S ) = f ( t , S ) + r K 0 t e r ( t s ) Φ ( d 2 ( t , s , S , B ( s ) ) ) d s + δ S 0 t e δ ( t s ) Φ ( d 1 ( t , s , S , B ( s ) ) ) d s .
In Equation (21), S represents the current price of the underlying asset (see [4]). The authors in [29] showed that the value of American options can be written as the sum of the corresponding European option price and the early exercise premium. Equation (21) lends itself to this interpretation being f ( t , S ) the price of an European put.

4. Numerical Results

In this section, we present some numerical results showing the accuracy of the approximations computed by means of the pricing model presented in Section 2 and Section 3. We denote by T the time expressed in year, and the interest rate, dividend yield, and volatility are expressed on an annual basis. The results were obtained using MATLAB on a MacBook Pro with a 2.6 GHz Intel Core i7 processor with 16 GB RAM. We considered n = 110 steps. The required computational time was about 9 s. The nonlinear system of equations of the form (10), arising in the problem discussed in Section 3, was solved applying the Broyden’s method, a quasi-Newtonian method.
The obtained approximated values of B ( t ) are plotted in Figure 1 for several values of σ . We chose S = K = 100 , T = 2 , r = 0.05 , δ = 0 and σ = 0.2 , 0.40 , 0.60 .
For several values of δ , the obtained values of B ( t ) are, instead, plotted in Figure 2. We chose S = K = 100 , T = 2 , r = 0.05 , σ = 0.2 and δ = 0.05 , 0.10 , 0.15 .
The numerical approximation of B ( t ) gained from our method was applied to (21); consequently, the trapezoidal numerical integration rule was engaged to calculate the price of an American put. The outcome values were compared with the benchmarks ones, stemming from a binomial tree with n = 10,000 time steps.
Table 1 displays values of the American put achieved through benchmarks methods. All the considered methods are able to produce an approximation of the boundary. In particular, the Cox, Ross and Rubenstein model ([30]), namely, C-R-R in the first column of Table 1, with n = 10,000 time steps (C-R-R), is considered. The second column contains the American put prices corresponding to different approximated values of B ( t ) using the four-point extrapolation scheme of [15], namely, G-J. The third column of Table 1 reports the modified two-point method of [8], namely, B-J. The fourth column embeds values stemming from the lower and upper bound approximation of [16], indicated as B-D. The fifth column, namely, Ju, shows the exponential functions method combined with a three-point Richardson extrapolation proposed by [17]. In addition, the iteration method proposed by [12] was considered and it is identified as K-J-K. Values within the seventh column in Table 1 represents the trapezoidal formulas approximations of [18] followed by the Newton–Raphson iteration, namely, K-K. N-B-R refers to the more recently product integration method for the approximation of the early boundary in the American option pricing problem exposed in [22]. Lastly, the last column in Table 1, indicated as D-M-M-V, represents the mean-value theorem approach that we proposed in this work.
In Table 1, we considered S as in first column, T = 3 , σ = 0.2 , r = 0.08 , K = 100 and δ = 0.08 and n = 32 . The benchmark values in the first column stem from a binomial tree (C-R-R model) with n = 10,000 time steps. For ease of reading, we underlined the best approximations.
The proposed numerical approach employs about 0.66 s to compute n = 32 points. It is worth underlining that approximations attained with our method are coherent or even better compared to the similar stemming from benchmark methods. The only exception holds for approximations obtained by [22]. Table 2 shows results from our proposal compared to the ones [22] considering, for D-M-M-V method, n = 110 points.
As Table 2 shows, the approximations obtained with our method were coherent with the ones obtained by applying the method in [22]. The resulting errors in the second rows for each value of S were very small and comparable with that relative to the N-B-R method. In particular, [22] numerically solved the one-dimensional reformulation of Kim’s integral equations ([11]) using a modified version of the Nyström method. Such a procedure allows for taking into account the singularity close to expiry presented by the early exercise boundary. In our method, instead, we directly dealt with Equation (15). In addition, our method may be easily extended to integral equations coming from more complicated dynamics.
In [22], taking into account n = 32 points, an error of range of 10 4 was obtained in about 14 s. Their calculations were performed on a PC with a 4.00 Intel Core i7 GHz processor with 16 GB RAM (see Figures 3–6 as reported by [22]). With our algorithm, also considering n = 110 points, we obtained an error of the same range in about 9 s. Moreover, our calculations were performed using a machine with a less powerful processor, a MacBook Pro with a 2.6 GHz Intel Core i7 processor with 16 GB of RAM. To test the convergence of our algorithm, we report in Table 3 some values of the American put corresponding to as many as the approximated values of B ( t ) . We considered: S = K = 100 , T = 1 , r = 0.08 , σ = 0.2 and δ = 0.08 . In Column 1, we report the time steps. The errors were computed as the difference between our values (D-D-M-V) with the ones obtained considering a binomial tree (C-R-R) model with n = 10,000 used as a benchmark. The error size of order 10 4 obtained when increasing the value of n proves the quality and the precision of the proposed method.

5. Conclusions

The present work proposed a new numerical method for the approximation of the early exercise boundary of an American option. American options are financial contracts allowing for the holder to buy or sell an underlying financial asset before the contract expiration. Thus, the evaluation of an American option requires to consider that the holder may have convenience to interrupt the contract at any time depending on the value of the underlying asset; that is, a collection of optimal exercise prices, one for each optimal exercise time, must be determined. Representing such a boundary by means of a nonstandard Volterra integral equation, we propose a numerical scheme in order to approximate the solution of the free boundary problem concerning the American option. Our proposal allows for managing well-known numerical problems arising in the presence of a singularity close to expiry and of a double integral. First, we provided theoretical instruments that were at the core of the proposed method. Consequently, we compared existing pricing models with results stemming from our method, showing the accuracy of the latter. The direct solution of the integral equation suggests that our method can be applied to more complicated cases coming up in finance.
Future research works will concern the use of an asset price process characterized by a nonconstant volatility term. In particular, we aim to expand our proposal to the case of the Heston model for American options, considering the associated multidimensional Volterra integral equation for the boundary problem.

Author Contributions

Conceptualization, D.V., R.D.M., M.M. and A.L.M. The authors contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wazwaz, A.M. Linear and Nonlinear Integral Equations; Springer: Berlin, Germany, 2011. [Google Scholar]
  2. Brunner, H. Volterra Integral Equations: An Introduction to Theory and Applications. In Cambridge Monographs on Applied and Computational Mathematics; Cambridge University Press: Cambridge, UK, 2017; No. 30. [Google Scholar]
  3. Brunner, H. Collocation Methods for Volterra Integral and Related Functional Differential Equations; Cambridge University Press: Cambridge, UK, 2004; Volume 15. [Google Scholar]
  4. Kim, I.J. The Analytic Valuation of American Options. Rev. Financ. Stud. 1990, 3, 547–572. [Google Scholar]
  5. Jacka, S.D. Optimal stopping and the American Put. Math. Financ. 1991, 1, 1–14. [Google Scholar] [CrossRef]
  6. Carr, P.; Jarrow, R.; Myneni, R. Alternative characterizations of American put options. Math. Financ. 1992, 2, 87–106. [Google Scholar]
  7. Barone-Adesi, G.; Whaley, R.E. Efficient analytic approximation of American option values. J. Financ. 1987, 81, 301–320. [Google Scholar] [CrossRef]
  8. Bunch, D.S.; Johnson, H. A simple and numerically efficient valuation method for American puts using a modified, Geske-Johnson approach. J. Financ. 1992, 47, 809–816. [Google Scholar] [CrossRef]
  9. Aitsahlia, F.; Lai, T.L. A canonical optimal stopping problem for American options and its numerical solution. J. Comput. Financ. 1999, 3, 33–52. [Google Scholar] [CrossRef]
  10. Aitsahlia, F.; Lai, T.L. Exercise Boundaries and Efficient Approximations to American Option Prices and Hedge Parameters. J. Comput. Financ. 2001, 4, 85–104. [Google Scholar] [CrossRef]
  11. Hou, C.; Little, T.; Pant, V. A new integral representation of the early exercise boundary for American put options. J. Comput. Financ. 2000, 3, 73–96. [Google Scholar]
  12. Kim, I.J.; Jang, B.-G.; Kim, K.T. A simple iterative method for the valuation of American options. Quant. Financ. 2013, 13, 885–895. [Google Scholar] [CrossRef]
  13. Stamicar, R.; Ševčovič, D.; Chadam, J. The early exercise boundary for the American put near expiry: Numerical approximation. Can. Appl. Math Q. 1999, 7, 427–444. [Google Scholar]
  14. Evans, J.D.; Kuske, R.; Keller, J.B. American options on assets with dividends near expiry. Math. Financ. 2002, 12, 219–237. [Google Scholar] [CrossRef]
  15. Geske, R.; Johnson, H.E. The American put option valued analytically. J. Financ. 1984, 39, 1511–1524. [Google Scholar] [CrossRef]
  16. Broadie, M.; Detemple, J. American option valuation: New bounds, approximations, and a comparison of existing methods. Rev. Financ. Stud. 1996, 9, 1211–1250. [Google Scholar] [CrossRef] [Green Version]
  17. Ju, N. Pricing an American option by approximating its early exercise boundary as a multipiece exponential function. Rev. Financ. Stud. 1998, 11, 627–646. [Google Scholar] [CrossRef]
  18. Kallast, S.; Kivinukk, A. Pricing and hedging American options using approximations by Kim integral equations. Eur. Financ. Rev. 2003, 7, 361–383. [Google Scholar] [CrossRef]
  19. Elliott, C.M.; Ockendon, J.R. Weak and Variational Methods for Free and Moving Boundary Problems (Research Notes in Mathematics); Pitman: London, UK, 1982; Volume 59. [Google Scholar]
  20. Zhu, S.-P. A new analytical approximation formula for the optimal exercise boundary of American put options. Int. J. Theor. Appl. Financ. 2006, 9, 1141–1177. [Google Scholar] [CrossRef]
  21. Lauko, M.; Ševčovič, D. Comparison of numerical and analytical approximations of the early exercise boundary of American put options. J. ANZIAM 2010, 51, 430–448. [Google Scholar] [CrossRef] [Green Version]
  22. Nedaiasl, K.; Bastani, A.F.; Rafiee, A. A product integration method for the approximation of the early exercise boundary in the American option pricing problem. Math. Meth. Appl. Sci. 2019, 42, 2825–2841. [Google Scholar] [CrossRef] [Green Version]
  23. Guardasoni, C.; Rodrigo, R.M.; Sanfelici, S. A Mellin transform approach to barrier option pricing. IMA J. Manag. Math. 2020, 31, 49–67. [Google Scholar] [CrossRef]
  24. Jeon, J.; Kim, G. An integral equation representation for American better-of option on two underlying assets. Adv. Contin. Discret. Model. 2022, 2022, 39. [Google Scholar] [CrossRef]
  25. Black, F.; Scholes, M. The pricing of options and corporate liabilities. J. Political Econ. 1973, 81, 637–654. [Google Scholar] [CrossRef] [Green Version]
  26. Heston, S.L. A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev. Financ. Stud. 1993, 6, 327–343. [Google Scholar] [CrossRef] [Green Version]
  27. Chiarella, C.; Ziogas, A.; Ziveyi, J. Representation of American Option Prices Under Heston Stochastic Volatility Dynamics Using Integral Transforms. In Contemporary Quantitative Finance; Chiarella, C., Novikov, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  28. Chen, X.; Chadam, J. A mathematical analysis of the optimal exercise boundary for American put options. SIAM J. Math. Anal. 2006, 38, 1613–1641. [Google Scholar] [CrossRef] [Green Version]
  29. Detemple, J.; Tian, W. The valuation of American options for a class of diffusion processes. Manag. Sci. 2002, 48, 917–937. [Google Scholar] [CrossRef]
  30. Cox, J.C.; Ross, S.A.; Rubenstein, M. Option Pricing: A Simplified Approach. J. Financ. Econ. 1979, 7, 229–263. [Google Scholar] [CrossRef]
Figure 1. The numerical approximation of B ( t ) : K = 100 , T = 2 , r = 0.05 , δ = 0 and σ = 0.2 , 0.40 , 0.60 .
Figure 1. The numerical approximation of B ( t ) : K = 100 , T = 2 , r = 0.05 , δ = 0 and σ = 0.2 , 0.40 , 0.60 .
Mathematics 11 00187 g001
Figure 2. The numerical approximation of B ( t ) : K = 100 , T = 2 , r = 0.05 , σ = 0.2 and δ = 0.05 , 0.10 , 0.15 .
Figure 2. The numerical approximation of B ( t ) : K = 100 , T = 2 , r = 0.05 , σ = 0.2 and δ = 0.05 , 0.10 , 0.15 .
Mathematics 11 00187 g002
Table 1. Numerical approximations of American put price: S as in the first column, T = 3 , σ = 0.2 , r = 0.08 , K = 100 , δ = 0.08 , and n = 32 . For each value of S, the resulting approximation errors are displayed in the second rows within brackets. Some data in Table 1 were extracted from [22]; in addition, underlined values represent approximated prices to which the lowest error corresponds.
Table 1. Numerical approximations of American put price: S as in the first column, T = 3 , σ = 0.2 , r = 0.08 , K = 100 , δ = 0.08 , and n = 32 . For each value of S, the resulting approximation errors are displayed in the second rows within brackets. Some data in Table 1 were extracted from [22]; in addition, underlined values represent approximated prices to which the lowest error corresponds.
SC-R-RG-JB-JB-DJuK-J-KK-KN-B-RD-M-M-V
8022.205022.207922.710622.198522.208422.194222.190022.204822.2075
(-)(2.9 × 10−3)(5.1 × 10−1)(6.5 × 10−3)(3.4 × 10−3)(1.1 × 10−2)(1.5 × 10−2)(2 × 10−4)(2.5 × 10−3)
9016.207116.163916.520516.198616.210616.199916.196016.206816.2096
(-)(4.3 × 10−2)(3.6 × 10−1)(5.9 × 10−2)(7.2 × 10−3)(1.1 × 10−2)(3.9 × 10−3)(1.1 × 10−4)(2.5 × 10−3)
10011.703711.705311.810611.698811.706611.699111.695811.703711.7061
(-)(1.6 × 10−3)(1.1 × 10−1)(4.9 × 10−3)(2.9 × 10−3)(4.9 × 10−3)(7.9 × 10−3)(1.0 × 10−5)(4.4 × 10−3)
1108.36718.38868.40728.36308.36958.36388.36138.36698.3689
(-)(2.1 × 10−2)(4.0 × 10−2)(4.1 × 10−3)(2.4 × 10−3)(3.3 × 10−3)(5.8 × 10−3)(2.0 × 10−4)(1.8 × 10−3)
1205.92995.94355.93105.92615.93235.92785.92585.92985.9314
(-)(1.4 × 10−2)(1.1 × 10−3)(3.8 × 10−3)(2.4 × 10−3)(2.1 × 10−3)(4.1 × 10−3)(1.0 × 10−4)(1.5 × 10−3)
Table 2. Comparison between N-B-R and D-M-M-V: S as in the first column, T = 3 , σ = 0.2 , r = 0.08 , K = 100 and δ = 0.08 . The benchmark value was obtained considering the C-R-R model with n = 10,000 time steps. For each value of S, the resulting approximation errors are displayed in the second rows within brackets.
Table 2. Comparison between N-B-R and D-M-M-V: S as in the first column, T = 3 , σ = 0.2 , r = 0.08 , K = 100 and δ = 0.08 . The benchmark value was obtained considering the C-R-R model with n = 10,000 time steps. For each value of S, the resulting approximation errors are displayed in the second rows within brackets.
SC-R-RN-B-RD-M-M-V
8022.205022.204822.2051
(-)(2.0 × 10−4)(1.0 × 10−4)
9016.207116.206816.2072
(-)(1.1 × 10−4)(1.0 × 10−4)
10011.703711.703711.7040
(-)(1.0 × 10−5)(3.0 × 10−4)
1108.36718.36698.3672
(-)(2.0 × 10−4)(1.0 × 10−4)
1205.92995.92995.9299
(-)(1.0 × 10−4)(1.0 × 10−4)
Table 3. Convergence analysis in approximating the American put price. The benchmark value, equal to 7.5009 , stems from a binomial tree (C-R-R model) with n = 10,000 steps. We considered S = K = 100 , T = 1 , r = δ = 0.08 , σ = 0.2 .
Table 3. Convergence analysis in approximating the American put price. The benchmark value, equal to 7.5009 , stems from a binomial tree (C-R-R model) with n = 10,000 steps. We considered S = K = 100 , T = 1 , r = δ = 0.08 , σ = 0.2 .
nD-D-M-VError
107.50590.0050
207.50280.0019
307.50200.0011
407.50178.0 × 10−4
907.50123.0 × 10−4
1107.50121.0 × 10−4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Veliu, D.; De Marchis, R.; Marino, M.; Martire, A.L. An Alternative Numerical Scheme to Approximate the Early Exercise Boundary of American Options. Mathematics 2023, 11, 187. https://doi.org/10.3390/math11010187

AMA Style

Veliu D, De Marchis R, Marino M, Martire AL. An Alternative Numerical Scheme to Approximate the Early Exercise Boundary of American Options. Mathematics. 2023; 11(1):187. https://doi.org/10.3390/math11010187

Chicago/Turabian Style

Veliu, Denis, Roberto De Marchis, Mario Marino, and Antonio Luciano Martire. 2023. "An Alternative Numerical Scheme to Approximate the Early Exercise Boundary of American Options" Mathematics 11, no. 1: 187. https://doi.org/10.3390/math11010187

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop