Abstract

This paper considers the pricing of a European option using a (𝐵,𝑆)-market in which the stock price and the asset in the riskless bank account both have hereditary price structures described by the authors of this paper (1999). Under the smoothness assumption of the payoff function, it is shown that the infinite dimensional Black-Scholes equation possesses a unique classical solution. A spectral approximation scheme is developed using the Fourier series expansion in the space 𝐶[,0] for the Black-Scholes equation. It is also shown that the 𝑛th approximant resembles the classical Black-Scholes equation in finite dimensions.

1. Introduction

The pricing of contingent claims in the continuous-time financial market that consists of a bank account and a stock account has been a subject of extensive research for the last decades. In the literature (e.g., [15]), the equations that describe the bank account and the price of the stock are typically written, respectively, as𝑑𝐵(𝑡)=𝑟𝐵(𝑡)𝑑𝑡,𝐵(0)=𝑥,𝑑𝑆(𝑡)=𝛼𝑆(𝑡)𝑑𝑡+𝜎𝑆(𝑡)𝑑𝑊(𝑡),𝑆(0)=𝑦,(1.1) where 𝑊={𝑊(𝑡),𝑡0} is a one-dimensional standard Brownian motion defined on a complete filtered probability space (Ω,𝐹,𝐏;{𝐹(𝑡),𝑡0}) and 𝑟, 𝛼, and 𝜎 are positive constants that represent, respectively, the interest rate of the bank account, the stock appreciation rate, and the stock volatility rate. The financial market that consists of one bank account and one stock account will be referred to as a (𝐵,𝑆)-market, where 𝐵 stands for the bank account and 𝑆 stands for the stock.

A European option contract is a contract giving the buyer of the contract the right to buy (sell) a share of a particular stock at a predetermined price at a predetermined time in the future. The European option problem is, briefly, to determine the fee (called the rational price) that the writer of the contract should receive from the buyer for the rights of the contract and also to determine the trading strategy the writer should use to invest this fee in the (𝐵,𝑆)-market in such a way as to ensure that the writer will be able to cover the option if it is exercised. The fee should be large enough that the writer can, with riskless investing, cover the option, but be small enough that the writer does not make an unfair (i.e., riskless) profit.

In [6], we noted reasons to include hereditary price structures to a (𝐵,𝑆)-market model and then introduced such a model using a functional differential equation to describe the dynamics of the bank account and a stochastic functional differential equation to describe those of the stock account. The paper then obtained a solution to the option pricing problem in terms of conditional expectation with respect to a martingale measure. The importance of including hereditary price structure in the stock price dynamics was also recognized by other researchers in recent years (see, e.g., [714]).

In particular, [6] was one of the firsts that took into consideration hereditary structure in studying the pricing problem of European option. There the authors obtained a solution to the option pricing problem in terms of conditional expectation with respect to a martingale measure. The two papers [7, 9] developed an explicit formula for pricing European options when the underlying stock price follows a nonlinear stochastic delay equation with fixed delays (resp., variable delays) in the drift and diffusion terms. The paper [8] computed the logarithmic utility of an insider when the financial market is modelled by a stochastic delay equation. There the author showed that, although the market does not allow free lunches and is complete, the insider can draw more from his wealth than the regular trader. The paper also offered an alternative to the anticipating delayed Black-Scholes formula, by proving stability of European call option proces when the delay coefficients approach the nondelayed ones. The paper [10] derived the infinite-dimensional Black-Scholes equation for the (𝐵,𝑆)-market, where the bank account evolves according to a linear (deterministic) functional differential equation and the stock dynamics is described by a very general nonlinear stochastic functional differential equation. A power series solution is also developed for the equation. Following the same model studied in [10], the work in [11] shows that under very mild conditions the pricing function is the unique viscosity solution of the infinite-dimensional Black-Scholes equation. A finite difference approximation scheme for the solution of the equation is developed and convergence result is also obtained. We mention here that option pricing problems were also considered by [1214] for a financial market that is more restricted than those of [10, 11].

This paper considers the pricing of a European option using a (𝐵,𝑆)-market, such as those in [6], in which the stock price and the asset in the riskless bank account both have hereditary price structures. Under the smoothness assumption of the payoff function, it is shown that the pricing function is the unique classical solution of the infinite-dimensional Black-Scholes equation. A spectral approximation scheme is developed using the Fourier series expansion in the space 𝐶[,0] for the Black-Scholes equation. It is also shown that the 𝑛th approximant resembles the celebrated classical Black-Scholes equation in finite dimensions (see, e.g., [4, 5]).

This paper is organized as follows. Section 2 summarizes the definitions and key results of [6] that will be used throughout this paper. The concepts of Fréchet derivative and extended Fréchet derivative are introduced in Section 3, along with results needed to make use of these derivatives. In Section 4, the results regarding the infinite-dimensional Black-Scholes equation and its corollary are restated from [6, 10]. Section 5 details the spectral approximate solution scheme for this equation. Section 6 is the paper's conclusion, followed by an appendix with the proof of Proposition 3.2.

2. The European Option Problem with Hereditary Price Structures

To describe the financial model with hereditary price structures, we start by defining our probability space. Let 0<< be a fixed constant. This constant will be the length of the time window in which the hereditary information is contained. If 𝑎,𝑏 with 𝑎<𝑏, denote the space of continuous functions 𝜙[𝑎,𝑏] by 𝐶[𝑎,𝑏]. Define 𝐶+[][][]𝑎,𝑏={𝜙𝐶𝑎,𝑏𝜙(𝜃)0𝜃𝑎,𝑏}.(2.1) Note that 𝐶[𝑎,𝑏] is a real separable Banach space equipped with the uniform topology defined by the sup-norm 𝜙=sup𝑡[𝑎,𝑏]|𝜙(𝑡)| and 𝐶+[𝑎,𝑏] is a closed subset of 𝐶[𝑎,𝑏]. Throughout the end of this paper, we let 𝐂=𝐶[,0] and 𝐂+[]}={𝜙𝐂𝜙(𝜃)0𝜃,0(2.2) for simplicity. If 𝜓𝐶[,) and 𝑡[0,), let 𝜓𝑡𝐂 be defined by 𝜓𝑡(𝜃)=𝜓(𝑡+𝜃), 𝜃[,0].

Let Ω=𝐶[,), the space of continuous functions 𝜔[,), and let 𝐹=𝐵(𝐶[,)), the Borel 𝜎-algebra of subsets of 𝐶[,) under the topology defined by the metric 𝑑Ω×Ω, where𝑑𝜔,𝜔=𝑛=112𝑛sup𝑡𝑛||𝜔(𝑡)𝜔||(𝑡)1+sup𝑡𝑛||𝜔(𝑡)𝜔||.(𝑡)(2.3)

Let 𝐏 be the Wiener measure defined on (Ω,𝐹) with []𝐏{𝜔Ω𝜔(𝜃)=0𝜃,0}=1.(2.4) Note that the probability space (Ω,𝐹,𝐏) is the canonical Wiener space under which the coordinate maps 𝑊={𝑊(𝑡),𝑡0}, 𝑊(𝑡)𝐶[,), defined by 𝑊(𝑡)(𝜔)=𝜔(𝑡) for all 𝑡 and 𝜔Ω is a standard Brownian motion and 𝐏{𝑊0=0}=1. Let the filtration 𝐹𝑊={𝐹(𝑡),𝑡} be the 𝐏-augmentation of the natural filtration of the Brownian motion 𝑊, defined by 𝐹(𝑡)={,Ω} for all 𝑡[,0] and 𝐹(𝑡)=𝜎(𝑊(𝑠),0𝑠𝑡),𝑡0.(2.5)

Equivalently, 𝐹(𝑡) is the smallest sub-𝜎-algebra of subsets of Ω with respect to which the mappings 𝑊(𝑠)Ω are measurable for all 0𝑠𝑡. It is clear that the filtration 𝐹𝑊 defined above is right continuous in the sense of [15].

Consider the 𝐂-valued process {𝑊𝑡,𝑡0}, where 𝑊0=0 and 𝑊𝑡(𝜃)=𝑊(𝑡+𝜃), 𝜃[,0] for all 𝑡0. That is, for each 𝑡0, 𝑊𝑡(𝜔)=𝜔𝑡 and 𝑊0=0. In [10], it is shown that 𝐹0=𝐹(𝑡) for 𝑡[,0] and 𝐹(𝑡)=𝐹𝑡 for 𝑡0, where 𝐹𝑡𝑊=𝜎𝑠,0𝑠𝑡,𝑡0.(2.6)

The new model for the (𝐵,𝑆)-market introduced in [6] has a hereditary price structure in the sense that the rate of change of the unit price of the investor's assets in the bank account 𝐵 and that of the stock account 𝑆 depend not only on the current unit price but also on their historical prices. Specifically, we assume that 𝐵 and 𝑆 evolve according to the following two linear functional differential equations: 𝐵𝑑𝐵(𝑡)=𝐿𝑡𝑆𝑑𝑡,𝑡0,(2.7)𝑑𝑆(𝑡)=𝑀𝑡𝑆𝑑𝑡+𝑁𝑡𝑑𝑊(𝑡),𝑡0,(2.8) with initial price functions 𝐵0=𝜙 and 𝑆0=𝜓, where 𝜙 and 𝜓 are given functions in 𝐂+. In the model, 𝐿, 𝑀, and 𝑁 are bounded linear functionals on the real Banach space 𝐂. The bounded linear functionals 𝐿,𝑀,𝑁𝐂 can be represented as (see [6]) 𝐿(𝜙)=0𝜙(𝜃)𝑑𝜂(𝜃),𝑀(𝜙)=0𝑁𝜙(𝜃)𝑑𝜉(𝜃),(2.9)(𝜙)=0𝜙(𝜃)𝑑𝜁(𝜃),𝜙𝐂,(2.10) where the above integrals are to be interpreted as Lebesgue-Stieltjes integrals and 𝜂, 𝜉, and 𝜁 are functions that are assumed to satisfy the following conditions.

Assumption 2.1. The functions 𝜂,𝜉[,0], are nondecreasing functions on [,0] such that 𝜂(0)𝜂()>0 and 𝜉(0)𝜉()>0, and the function 𝜁[,0] is a function of bounded variation on [,0] such that 0𝜙(𝜃)𝑑𝜁(𝜃)𝜎>0 for every 𝜙𝐂+.

We will, throughout the end, extend the domain of the above three functions to 𝑅 by defining 𝜂(𝜃)=𝜂() for 𝜃 and 𝜂(𝜃)=𝜂(0) for 𝜃0, and so forth.

Proposition 2.3 in [6] provides an existence and uniqueness result under mild conditions, so the model makes sense mathematically to use. Note that the equations described by (2.7)-(2.8) include (1.1) as a special case. Therefore, the model considered in this paper is a generalization of that considered in most of the existing literature (see, e.g., [5]).

For the purpose of analyzing the discount rate for the bank account, let us assume that the solution process 𝐵(𝐿;𝜙)={𝐵(𝑡),𝑡<} of (2.7) with the initial function 𝜙𝐂+ takes the following form: 𝐵(𝑡)=𝜙(0)𝑒𝑟𝑡,𝑡0,(2.11) and 𝐵0=𝜙𝐂+. Then the constant 𝑟 satisfies the following equation: 𝑟=0𝑒𝑟𝜃𝑑𝜂(𝜃).(2.12) The existence and uniqueness of a positive number 𝑟 that satisfies the above equation is shown in [6].

Throughout the end, we will fix the initial unit price functions 𝜙, and 𝜓𝐂+, and the functional 𝑁𝐂 for the stock price described in (2.8) and (2.10). For the purpose of making the distinction when we interchange the usage of 𝑀𝐂 and 𝐿𝐂 in (2.8), we write the stock price process 𝑆(𝑀,𝑁;𝜓) as 𝑆(𝑀)={𝑆(𝑡),𝑡} for simplicity. And, when the functional 𝐾𝐂, 𝐾(𝜙𝑡)=𝑟𝜙(𝑡) is used in place of 𝑀𝐂 in (2.8), its solution process will be written as 𝑆(𝐾)={𝑆(𝑡),𝑡}.

In [6], the basic theory of European option pricing using the (𝐵,𝑆)-market model described in (2.7)-(2.8) is developed. We summarize the key definitions and results below.

A trading strategy in the (𝐵,𝑆)-market is a progressively measurable vector process 𝜋={(𝜋1(𝑡),𝜋2(𝑡)),0𝑡<} defined on (Ω,𝐹,𝐏;𝐹𝑊) such that for each 𝑎>0, 𝑎0𝐄𝜋2𝑖(𝑡)𝑑𝑡<,𝑖=1,2,(2.13) where 𝜋1(𝑡) and 𝜋2(𝑡) represent, respectively, the number of units of the bank account and the number of shares of the stock owned by the writer at time 𝑡0, and 𝐄 is the expectation with respect to 𝐏.

The writer's total asset is described by the wealth process 𝑋𝜋(𝑀)={𝑋𝜋(𝑡),0𝑡<} defined by 𝑋𝜋(𝑡)=𝜋1(𝑡)𝐵(𝑡)+𝜋2(𝑡)𝑆(𝑡),0𝑡<,(2.14) where again 𝐵(𝐿;𝜙) and 𝑆(𝑀,𝑁;𝜓) are, respectively, the unit price of the bank account and the stock described in (2.7) and (2.8). This wealth process can clearly take both positive and negative values, since it is permissible that (𝜋1(𝑡),𝜋2(𝑡))2.

We will make the following basic assumption throughout this paper.

Assumption 2.2 (self-financing condition). In the (𝐵,𝑆)-market, it is assumed that all trading strategies 𝜋 satisfy the following self-financing condition: 𝑋𝜋(𝑡)=𝑋𝜋(0)+𝑡0𝜋1(𝑠)𝑑𝐵(𝑠)+𝑡0𝜋2(𝑠)𝑑𝑆(𝑠),0𝑡<,a.s.(2.15) or equivalently, 𝑑𝑋𝜋(𝑡)=𝜋1(𝑡)𝑑𝐵(𝑡)+𝜋2(𝑡)𝑑𝑆(𝑡),0𝑡<.(2.16)

Using the same notation as in [6] (see also [10]) the set of all self-financing trading strategies 𝜋 will be denoted by SF(𝐿,𝑀,𝑁;𝜙,𝜓) or simply SF if there is no danger of ambiguity.

For the unit price of the bank account 𝐵(𝐿;𝜙)={𝐵(𝑡),𝑡0} and the stock 𝑆(𝑀,𝑁;𝜓)={𝑆(𝑡),𝑡0} described in (2.7) and (2.8), define 𝑊(𝑡)=𝑊(𝑡)+𝑡0𝛾𝐵𝑠,𝑆𝑠𝑑𝑠,𝑡0,(2.17) where 𝛾𝐂+×𝐂+ is defined by 𝛾(𝜙,𝜓)=𝜙(0)𝑀(𝜓)𝜓(0)𝐿(𝜙).𝜙(0)𝑁(𝜓)(2.18)

Define the process 𝑍(𝐿,𝑀,𝑁;𝜙,𝜓)={𝑍(𝑡),𝑡0} by 𝑍(𝑡)=exp𝑡0𝛾𝐵𝑠,𝑆𝑠1𝑑𝑊(𝑠)2𝑡0||𝛾𝐵𝑠,𝑆𝑠||2𝑑𝑠,𝑡0.(2.19) The following results are proven in [6].

Lemma 2.3. The process 𝑍(𝐿,𝑀,𝑁;𝜙,𝜓)={𝑍(𝑡),𝑡0} defined by (2.19) is a martingale defined on (Ω,𝐹,𝐏;𝐹𝑊).

Lemma 2.4. There exists a unique probability measure 𝐏 defined on the canonical measurable space (Ω,𝐹) such that 𝟏𝐏(𝐴)=𝐄𝐴𝑍(𝑇)𝐴𝐹𝑇,0<𝑇<,(2.20) where 𝟏𝐴 is the indicator function of 𝐴𝐹𝑇.

Lemma 2.5. The process 𝑊 defined by (2.17) is a standard Brownian motion defined on the filtered probability space (Ω,𝐹,𝐏;𝐹𝑊).

From the above, it has been shown (see [6, equation (14)]) that 𝑆𝑑𝑆(𝑡)=𝑟𝑆(𝑡)𝑑𝑡+𝑁𝑡𝑑𝑊(𝑡),(2.21) with 𝑆0=𝜓𝐂+. It is also clear that the probabilistic behavior of 𝑆(𝑀) under the probability measure 𝐏 is the same as that of 𝑆(𝐾) under the probability measure 𝐏; that is, they have the same distribution.

Define the process 𝑌𝜋(𝐿,𝑀,𝑁;𝜙,𝜓)={𝑌𝜋(𝑡),𝑡0}, called the discounted wealth process, by 𝑌𝜋𝑋(𝑡)=𝜋(𝑡)𝐵(𝑡),𝑡0.(2.22) We say that a trading strategy 𝜋 from SF(𝐿,𝑀,𝑁;𝜙,𝜓) belongs to a subclass SF𝜍SF if 𝐏 a.s. 𝑌𝜋𝐄(𝑡)𝜍𝐹𝑡,𝑡0,(2.23) where 𝐄 is the expectation with respect to 𝐏, 𝜍 is a nonnegative 𝐹-measurable random variable such that 𝐄[𝜍]<. We say that 𝜋 belongs to SF+SF if 𝜍0.

In [6, 10], it is shown that 𝑌𝜋 is a local martingale; for 𝜋SF𝜍, 𝑌𝜋 is a supermartingale, and is a nonnegative supermartingale if 𝜋SF+.

Throughout, we assume the reward function Λ is an 𝐹𝑇-measurable nonnegative random variable satisfying the following condition: 𝐄Λ1+𝜖<,(2.24) for some 𝜖>0. Here, 𝑇>0 is the expiration time. (Note that the above condition on Λ implies that 𝐄[Λ]<.)

Let Λ be a nonnegative 𝐹𝑇-measurable random variable satisfying (2.24). A trading strategy 𝜋SF is a (𝑀;Λ,𝑥)-hedge of European type if 𝑋𝜋(0)=𝜋1(0)𝜙(0)+𝜋2(0)𝜓(0)=𝑥(2.25) and 𝐏 a.s. 𝑋𝜋(𝑇)Λ.(2.26) We say that a (𝑀;Λ,𝑥)-hedge trading strategy 𝜋SF(𝑀) is minimal if 𝑋𝜋(𝑇)𝑋𝜋(𝑇)(2.27) for any (𝑀;Λ,𝑥)-hedge strategy 𝜋SF(𝑀).

Let Π(𝑀;Λ,𝑥) be the set of (𝑀;Λ,𝑥)-hedge strategies from SF+(𝑀). Define 𝐶(𝑀;Λ)=inf{𝑥0Π(𝑀;Λ,𝑥)}.(2.28) The value 𝐶(𝑀;Λ) defined above is called the rational price of the contingent claim of European type. If the infimum in (2.28) is achieved, then 𝐶(𝑀;Λ) is the minimal possible initial capital for which there exists a trading strategy 𝜋SF+(𝑀) possessing the property that 𝐏 a.s. 𝑋𝜋(𝑇)Λ.

Let 𝑌(𝑀)={𝑌(𝑡),0𝑡𝑇} be defined by 𝐄Λ𝑌(𝑡)=𝐹𝐵(𝑇)𝑡,0𝑡𝑇,(2.29) where 𝐹𝑡𝑊=𝜎(𝑠,0𝑠𝑡). In [10], it is shown that the process 𝑌(𝑀) is a martingale defined on 𝑊)(Ω,𝐹,𝐏;𝐹 and can be represented by 𝑌(𝑡)=𝑌(0)+𝑡0𝛽(𝑠)𝑑𝑊(𝑠),(2.30) where 𝛽={𝛽(𝑡),0𝑡𝑇} that is 𝐹𝑊-adapted and 𝑇0𝛽2(𝑡)𝑑𝑡< (𝐏 a.s.).

The following lemma and theorem provide the main results of [6, 10]. Let 𝜋={(𝜋1(𝑡),𝜋2(𝑡)),0𝑡𝑇} be a trading strategy, where 𝜋2(𝑡)=𝛽(𝑡)𝐵(𝑡)𝑁𝑆𝑡,𝜋1𝑆(𝑡)=𝑌(𝑡)(𝑡)𝜋𝐵(𝑡)2[].(𝑡),𝑡0,𝑇(2.31)

Lemma 2.6. 𝜋SF(𝑀) and for each 𝑡[0,𝑇], 𝑌(𝑡)=𝑌𝜋(𝑡) for each 𝑡[0,𝑇] where again 𝑌𝜋 is the process defined in (2.22) with the minimal strategy 𝜋 defined in (2.31).

Theorem 2.7. Let Λ be an 𝐹𝑇-measurable random variable defined on the filtered probability space (Ω,𝐹,𝐏;𝐹𝑊) that satisfies (2.24). Then the rational price 𝐶(𝑀;Λ) defined in (2.28) is given by 𝐄𝑒𝐶(𝑀;Λ)=𝑟𝑇Λ,(2.32) where 𝑟 is the positive constant that satisfies (2.12). Furthermore, there exists a minimal hedge 𝜋={(𝜋1(𝑡),𝜋2(𝑡)),0𝑡𝑇}, where 𝜋2(𝑡)=𝛽(𝑡)𝐵(𝑡)𝑁𝑆𝑡,𝜋1(𝑡)=𝑌𝜋(𝑡)𝜋2𝑆(𝑡)(𝑡),𝐵(𝑡)(2.33) and the process 𝛽={𝛽(𝑡),0𝑡𝑇} is given by (2.30).
If in addition, the reward Λ is intrinsic, that is, Λ=Γ(𝑆(𝑀)) for some measurable function Γ𝐂+, then the rational price 𝐶(𝑀;Λ) does not depend on the mean growth rate 𝑀 of the stock and 𝐄𝑒𝐶(Λ)=𝑟𝑇Λ.(2.34)

3. Fréchet and Extended Fréchet Derivatives

In this section, results are proven that allow the use of a Dynkins formula for stochastic functional differential equation as found in [16, 17]. We assume contingent claims of European type in which the 𝐹𝑇-measurable reward function Λ has the explicit expression Λ=𝑓(𝑆𝑇), where again 𝑆𝑇(𝜃)=𝑆(𝑇+𝜃), 𝜃[,0] and 𝑆(𝐾)={𝑆(𝑡),𝑡0} is the unit price of the stock described by the following equation: 𝑆𝑑𝑆(𝑡)=𝑟𝑆(𝑡)𝑑𝑡+𝑁𝑡𝑑𝑊(𝑡),𝑡0,(3.1) where 𝑆0=𝜓𝐂+. Throughout this section, we assume that 𝑆(𝑡), and therefore 𝑁(𝑆𝑡), are uniformly bounded almost surely. This assumption is realistic for the price of a stock during time interval [0,𝑇] in a financial system with finite total wealth.

The remaining sections make extensive use of Fréchet derivatives. Let 𝐂 be the space of bounded linear functionals Φ𝐂. 𝐂 is a real separable Banach space under the supremum operator norm Φ=sup𝜙0||||Φ(𝜙).𝜙(3.2) For Ψ[0,𝑇]×𝐂, we denote the Fréchet derivative of Ψ at 𝜙𝐂 by 𝐷Ψ(𝑡,𝜙). The second Fréchet derivative at 𝜙 is denoted as 𝐷2Ψ(𝑡,𝜙).

Let Γ be the vector space of all simple functions of the form 𝑣𝟏{0}, where 𝑣𝑅 and 𝟏{0}[,0] is defined by 𝟏{0}[(𝜃)=0,for𝜃,0),1,for𝜃=0.(3.3) Form the direct sum 𝐂Γ and equip it with the complete norm 𝜙+𝑣𝟏{0}=sup𝜃[,0]||||𝜙(𝜃)+|𝑣|,𝜙𝐂,𝑣.(3.4) Then 𝐷Ψ(𝑡,𝜙) has a unique continuous linear extension from 𝐂Γ to which we will denote by 𝐷Ψ(𝑡,𝜙), and similarly for 𝐷2Ψ(𝑡,𝜙); see [16] or [17] for more details.

Finally, we define 𝐆(Ψ)𝑡,𝜓𝑡=lim𝑢0+1𝑢Ψ𝑡,𝜓𝑡+𝑢Ψ𝑡,𝜓𝑡(3.5) for all 𝑡[0,) and 𝜓𝐂+, where 𝜓[,) is defined by [𝜓(𝑡)=𝜓(𝑡)if𝑡,0)𝜓(0)if𝑡0.(3.6) Let 𝑓𝐂. We say that 𝑓𝐶1(𝐂) if 𝑓 has a continuous Fréchet derivative. Similarly, 𝑓𝐶𝑛(𝐂) if 𝑓 has a continuous 𝑛th Fréchet derivative. For 𝑓𝑅+×𝐂, we say that 𝑓𝐶,𝑛([0,)×𝐂) if 𝑓 is infinitely differentiable in its first variable and has a continuous 𝑛th partial derivative in its second variable.

Proposition 3.1. Let 𝜑𝐂 and 𝑓𝐂 with 𝑓𝐶2(𝐂). Define Ψ[0,𝑇]×𝐂 by Ψ(𝑡,𝜑)=𝑒𝑟(𝑇𝑡)𝐄𝑓𝑆𝑇𝑆𝑡.=𝜑(3.7) Then Ψ𝐶,2([0,)×𝐂).

Proof. That 𝑒𝑟(𝑇𝑡) is 𝐶[0,) is clear, so we have only to show that Υ𝐶2(𝐂), where Υ(𝜑)=𝐄[𝑓(𝑆𝑇)𝑆𝑡=𝜑] given that 𝑓𝐶2(𝐂).
We have that 𝑆𝑑𝑆(𝑡)=𝑟𝑆(𝑡)𝑑𝑡+𝑁𝑡𝑑𝑊(𝑡),𝑡0,(3.8) with 𝑆0=𝜓𝐂+. Under Assumption 2.1 on 𝑁𝐂 and the properties of Υ, it can be shown that there exists 𝐻××𝐂𝐂 such that 𝑆𝑡=𝐻(𝑡,𝑊(𝑡),𝜓). Therefore, 𝐄𝑓𝑆Υ(𝜑)=𝑇𝑆𝑡=1=𝜑2𝜋𝑓(𝐻(𝑇𝑡,𝑦,𝜑))𝑒𝑦2/2𝑑𝑦.(3.9) By Theorem 3.2, Chapter 2 of [16], 𝐻(𝑡,𝑦,)𝐶1(𝐂) as a function of 𝜓. By a second application of the same theorem (since 𝑓𝐶2(𝐂)), we have that 𝐻(𝑡,𝑦,)𝐶2(𝐂) as a function of 𝜓. Define 𝑔××𝐂 by 𝑔=𝑓𝐻. Since 𝑓𝐶2(𝐂) and 𝐻(𝑡,𝑦,)𝐶2(𝐂) in its third variable, 𝑔(𝑡,𝑦,)𝐶2(𝐂). Hence, for 𝜑,𝜙𝐂, 𝐄𝑓𝑆𝑇𝑆𝑡𝐄𝑓𝑆=𝜑+𝜙𝑇𝑆𝑡=1=𝜑2𝜋[]𝑒𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜙))𝑓(𝐻(𝑇𝑡,𝑦,𝜑))𝑦2/2=1𝑑𝑦2𝜋𝐷𝑔(𝑇𝑡,𝑦,𝜑)(𝜙)𝑒𝑦2/21𝑑𝑦+2𝜋𝑜(𝜙)𝑒𝑦2/2𝑑𝑦,(3.10) where 𝑜(𝜙) is a function mapping continuous functions into the reals such that 𝑜(𝜙)𝜙0as𝜙0.(3.11) The last integral is clearly 𝑜(𝜙) and 12𝜋𝐷𝑔(𝑇𝑡,𝑦,𝜑)(𝜙)𝑒𝑦2/2𝑑𝑦(3.12) is bounded and linear in 𝜙, so this integral is the first Fréchet derivative with respect to 𝜑. Since 𝑔(𝑡,𝑦,)𝐶2(𝐂), the process can be repeated, giving a second Fréchet derivative with respect to 𝜑 and so Υ𝐶2(𝐂).

Proposition 3.2. Let 𝜑𝐂 and 𝑓𝐂. Further assume 𝑓𝐶2(𝐂) and let Ψ[0,𝑇]×𝐂 be defined by Ψ(𝑡,𝜑)=𝑒𝑟(𝑇𝑡)𝐸𝑓𝑆𝑇𝑆𝑡.=𝜑(3.13) Then if 𝐷𝑓 and 𝐷2𝑓 are globally Lipschitz, then so is 𝐷2Ψ.

Recall from Proposition 3.1 that 𝑔𝑅×𝑅×𝐂𝑅 is 𝑓𝐻 where 𝑆𝑡=𝐻(𝑡,𝑊(𝑡),𝜓) with 𝑆0=𝜓𝐂+.

Proposition 3.3. Let 𝜑𝐂 and 𝑓𝐂. Further assume 𝑓𝐶2(𝐂) and let Ψ[0,𝑇]×𝐂 be defined by Ψ(𝑡,𝜑)=𝑒𝑟(𝑇𝑡)𝐸𝑓𝑆𝑇𝑆𝑡.=𝜑(3.14) Then if 𝑓 and 𝐆(𝑔)(𝑇𝑡,𝑦,𝜓𝑡) are globally bounded, then so is 𝐆(Ψ)(𝑡,𝜓𝑡).

Proof. We have that 𝐆(Ψ)𝑡,𝜓𝑡=lim𝑢0+1𝑢Ψ𝑡,𝜓𝑡+𝑢Ψ𝑡,𝜓𝑡=lim𝑢0+1𝑢12𝜋𝑔𝑇𝑡,𝑦,𝜓𝑡+𝑢𝑔𝑇𝑡,𝑦,𝜓𝑡𝑒𝑦2/2=1𝑑𝑦2𝜋lim𝑢0+1𝑢𝑔𝑇𝑡,𝑦,𝜓𝑡+𝑢𝑔𝑇𝑡,𝑦,𝜓𝑡𝑒𝑦2/21𝑑𝑦2𝜋𝑀𝑒𝑦2/2𝑑𝑦=𝑀<,(3.15) where we used the assumption that 𝑓 and hence 𝑔 are globally bounded to move the limit inside the integral and 𝐆(𝑔)(𝑇𝑡,𝑦,𝜓𝑡)𝑀<.

Remark 3.4. Note that since 𝐆(Ψ)(𝜓𝑠) is bounded for all 𝑠[0,𝑇], 𝑡0𝐷Ψ(𝑠,𝜓𝑠)(𝑑𝜓𝑠) exits. Also, if 𝐷2𝑓 is bounded, 𝑡0𝐷2Ψ(𝑠,𝜓𝑠)(𝑑𝜓𝑠,𝑑𝜓𝑠) exits (see [18]).

4. The Infinite-Dimensional Black-Scholes Equation

It is known (e.g., [4, 5]) that the classical Black-Scholes equation is a deterministic parabolic partial differential equation (with a suitable auxiliary condition) the solution of which gives the value of the European option contract at a given time. Propositions 3.1 through 3.3 allow us to use the Dynkin formula in [16]. With it, a generalized version of the classical Black-Scholes equation can be derived for when the (𝐵,𝑆)-market model is (2.7) and (2.8). The following theorem is a restatement of Theorem 3.1 in [10].

Theorem 4.1. Let Ψ(𝑡,𝜑)=𝑒𝑟(𝑇𝑡)𝐄[𝑓(𝑆𝑇)𝑆𝑡=𝜑], where 𝑆0=𝜓𝐂+ and 𝑡[0,𝑇]. Let 𝑓 be a 𝐶2(𝐂) function with 𝐷𝑓 and 𝐷2𝑓 globally Lipschitz and let Λ=𝑓(𝑆𝑇) and 𝑥=𝑋𝜋(0). Finally, let 𝑓 and 𝐆(𝑔)(𝑇𝑡,𝑦,𝜓𝑡) be globally bounded. Then if 𝑋𝜋(𝑡)=Ψ(𝑡,𝑆𝑡) is the wealth process for the minimal (Λ,𝑥)-hedge, one has 𝜕𝑟Ψ(𝑡,𝜑)=𝜕𝑡Ψ(𝑡,𝜑)+𝐆(Ψ)𝑡,𝜑𝑡+𝐷Ψ(𝑡,𝜑)𝑟𝜑(0)𝟏{0}+12𝐷2Ψ(𝑡,𝜑)𝑁(𝜑)𝟏{0},𝑁(𝜑)𝟏{0}[,a.s.(𝑡,𝜑)0,𝑇)×𝐂+,(4.1) where Ψ(𝑇,𝜑)=𝑓(𝜑)𝜑𝐂+,(4.2) and the trading strategy (𝜋1(𝑡),𝜋2(𝑡)) is defined by 𝜋2(𝑡)=𝟏𝐷Ψ(𝑡,𝜑){0}𝜋a.s.,11(𝑡)=𝑋𝐵(𝑡)𝜋(𝑡)𝜑(0)𝜋2.(𝑡)(4.3) Furthermore, if (4.1) and (4.2) hold, then Ψ(𝑡,𝑆𝑡) is the wealth process for the (Λ,𝑥)-hedge with 𝜋2(𝑡)=𝐷Ψ(𝑡,𝑆𝑡)(𝟏{0}) and 𝜋1(𝑡)=(1/𝐵(𝑡))[𝑋𝜋(𝑡)𝑆(𝑡)𝜋2(𝑡)].

Proof. The theorem is a restatement of Theorem 3.1 in [10] and is therefore omitted.

Note
Equations (4.1) and (4.2) are the generalized Black-Scholes equation for the (𝐵,𝑆)-market with hereditary price structure as described by (2.7) and (2.8).

5. Approximation of Solutions

In this section, we will solve the generalized Black-Scholes equation (4.1)-(4.2) by considering a sequence of approximations of its solution. By a (classical) solution to (4.1)-(4.2), we mean Ψ[0,𝑇]×𝐂 satisfying the following conditions:

(i)Ψ𝐶1,2([0,𝑇]×𝐂),(ii)Ψ(𝑇,𝜑)=𝑓(𝜑) for all 𝜑𝐂,(iii)Ψ satisfies (4.1).

The sequence of approximate solutions is constructed by looking at finite-dimensional subspaces of 𝐂, solving (4.1)-(4.2) on these subspaces, and then showing that as the dimension of the subspaces goes to infinity, the finite-dimensional solutions converge to a solution of the original problem. Theorem 5.2, Remark 5.3, and Corollary 5.4 show that the generalized Black-Scholes equation can be solved by solving two simpler equations. The first of these, a first-order partial differential equation, can be handled by traditional techniques once the second equation is solved. Theorem 5.5 provides a solution to the second. Proposition 5.7, which uses Lemma 5.6, gives a generalized Black-Scholes formula for the standard European call option when used in conjunction with Theorem 5.2.

We start by noting that 𝐂𝐿2[,0] where 𝐿2[,0] is the space of all square-integrable functions on the interval [,0]. Furthermore, 𝐂 is dense in 𝐿2[,0]. It is well known (e.g., [19]) that even extensions of a function 𝜑 in 𝐿2[,0] may be represented by a cosine Fourier series where 𝜑𝑁𝑖=0𝑎𝑖cos2𝜋𝑖20(5.1) as 𝑁 where 𝑎0=10𝑎𝜑(𝜃)𝑑𝜃,𝑖=20𝜑(𝜃)cos2𝜋𝑖𝜃𝑑𝜃,𝑖=1,2,3,.(5.2) Here, 𝑓22=0𝑓2(𝜃)𝑑𝜃(5.3) for 𝑓𝐿2[,0]. If 𝜑 is Hölder-continuous, then the convergence is also point wise (see, e.g., [20]).

Throughout this section, we let 𝐿2𝑛[,0] be the subspace of 𝐿2[,0] consisting of functions that can be represented as a finite Fourier series, that is, 𝜑(𝑛)𝐿2𝑛[,0] if 𝜑(𝑛)(𝜃)=𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝜃(5.4) for all 𝜃[,0].

We will see that it is convenient having a spanning set {𝑓𝑖}𝑖=0 for 𝐿2𝑛[,0] where 𝑓𝑖[,0] for 𝑖=0,1, such that 𝑓𝑖(0)=1 for all 𝑖 and 𝑁(𝑓𝑖)=0𝑓𝑖(𝜃)𝑑𝜁(𝜃)=𝛿 for all 𝑖. Here, 𝛿=0𝑑𝜁(𝜃). Let 𝑞[,0] be any function such that 𝑁(𝑞)=𝑞(0)0. For example, let 𝑞(𝜃)=1+1𝛿𝑑2𝑑1𝜃+𝜃2,(5.5) where 𝑑1=0𝜃𝑑𝜁(𝜃) and 𝑑2=0𝜃2𝑑𝜁(𝜃). To this end, we define the following functions. Let 𝑓0[],𝑓(𝜃)=1𝜃,01(𝜃)=𝛼1,1+𝛼1,2𝑞[],𝑓(𝜃)𝜃,02(𝜃)=𝛼2,1𝑞(𝜃)+𝛼2,2cos2𝜋𝜃[],𝜃,0(5.6) and for 𝑖=3,4,, 𝑓𝑖(𝜃)=𝛼𝑖,1cos2𝜋(𝑖2)𝜃+𝛼𝑖,2cos2𝜋(𝑖1)𝜃[]𝜃,0.(5.7) Recall that 𝑁𝐿2[,0] is defined by 𝑁(𝜑)=0𝜑(𝜃)𝑑𝜁(𝜃),(5.8) and let 𝑐𝑖=𝑁cos2𝜋𝑖=0cos2𝜋𝑖𝜃𝑑𝜁(𝜃).(5.9) Here again 𝑞[,0] is any function such that 𝑁(𝑞)=𝑞(0)0. For example, 𝑞 can be chosen as in (5.5). In this case, the constant 𝛼1,2 is nonzero but otherwise arbitrary, 𝛼1,1=1𝛼1,2𝛼𝑞(0),2,1=𝛿𝑐1𝑞(0)1𝑐1,𝛼2,2=1𝛼2,1𝑞(0),(5.10) and so on with 𝛼𝑖,2=𝛿𝑐𝑖2𝑐𝑖1𝑐𝑖2,𝛼𝑖,1=1𝛼𝑖,2(5.11) for 𝑖3.

Lemma 5.1. The set {𝑓𝑖}𝑖=0 defined in (5.6) and (5.7) forms a spanning set for 𝐿2[,0] in the sense that 𝜑𝑛+1𝑖=0𝑥𝑖𝑓𝑖20(5.12) as 𝑛, where the 𝑥𝑖 are defined by 𝑥𝑛+1=𝑎𝑛𝛼𝑛+1,2,𝑥𝑛=𝑎𝑛1𝑥𝑛+1𝛼𝑛+1,1𝛼𝑛,2,(5.13) and continuing using 𝑥𝑖=𝑎𝑖1𝑥𝑖+1𝛼𝑖+1,1𝛼𝑖,2(5.14) until 𝑥1𝑥=2𝛼2,1𝛼1,2,𝑥0=𝑎0𝑥1𝛼1,1.(5.15) This set of functions has the properties that 𝑓𝑖(0)=1 and 𝑁(𝑓𝑖)=𝛿 for all 𝑖=0,1,.

Proof. For any 𝜑𝐿2[,0], we can construct an even extension 𝜙𝐿2[,] where 𝜙(𝜃)=𝜑(𝜃) for all 𝜃[,0] and 𝜙(𝜃)=𝜑(𝜃) for all 𝜃[0,]. The function 𝜙 may be represented by a Fourier series of cosine functions 𝜙(𝜃)𝑁𝑖=0𝑎𝑖cos2𝜋𝑖𝜃,(5.16) where the “’’ is used to indicate that 𝜙𝑁𝑖=0𝑎𝑖cos2𝜋𝑖20(5.17) as 𝑁. In what mentioned before, 𝑎0=10𝑎𝜙(𝜃)𝑑𝜃,𝑖=20𝜙(𝜃)cos2𝜋𝑖𝜃𝑑𝜃(5.18) for all 𝑖=1,2,. For simplicity, we will replace the “’’ with an equality sign knowing that mean-square convergence is implied.
For the Fourier series, the basis is cos2𝜋𝑖𝜃𝑖=0,(5.19) so the first term of this basis and {𝑓𝑖}𝑖=0 are the same, namely, the constant “1.’’ Clearly 𝑓0(0)=1 and 𝑁(𝑓0)=𝛿. The first part of this proof is to show that for all 𝑖=0,1,, 𝑓𝑖(0)=1 and 𝑁(𝑓𝑖)=𝛿.
For 𝑓1, we have that 𝑓1(0)=𝛼1,1+𝛼1,2𝑞(0)=1 which implies that 𝛼1,1=1𝛼1,2𝑞(0).(5.20) Also, 𝑁(𝑓1)=𝛼1,1𝛿+𝛼1,2𝑁(𝑞)=𝛿. Since we do not want 𝛼1,2=0, we require that 𝑁(𝑞)=𝑞(0).(5.21) There are no restrictions on 𝛼1,2 other than 𝛼1,20.
For 𝑓2, 𝛼2,1𝑞(0)+𝛼2,2=1 requires that 𝛼2,2=1𝛼2,1𝑞(0).(5.22) Since we want 𝛼2,1𝑁(𝑞)+𝛼2,2𝑐1=𝛿, then 𝛼2,1=𝛿𝑐1𝑁(𝑞)𝑞(0)𝑐1=𝛿𝑐1𝑞(0)1𝑐1.(5.23)
The rest of the 𝑓𝑖, that is, where 𝑖3, are handled alike. In order that 𝑓𝑖(0)=1, we require that 𝛼𝑖,1=1𝛼𝑖,2. To ensure that 𝑁(𝑓𝑖)=𝛿, 𝛼𝑖,2=𝛿𝑐𝑖2𝑐𝑖1𝑐𝑖2.(5.24) We have now shown that the sequence of functions {𝑓𝑖}𝑖=0 is such that 𝑓𝑖(0)=1 and 𝑁(𝑓𝑖)=𝛿 for all 𝑖=0,1,. Now it must be shown that this sequence is a spanning set for 𝐿2[,0]. To do this, we will compare this sequence of functions with the cosine Fourier sequence of functions.
Consider 𝜑(𝑛)[,0] where 𝜑(𝑛)(𝜃)=𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝜃.(5.25) We would like 𝜑(𝑛)(𝜃)=𝑛+1𝑖=0𝑥𝑖𝑓𝑖(𝜃)(5.26) for some set {𝑥𝑖}𝑛+1𝑖=0 of real numbers. By the Fourier expansion, 𝜑(𝑛)(𝜃)=𝑎0+𝑎1cos2𝜋𝜃++𝑎𝑛cos2𝜋𝑛𝜃.(5.27) We want {𝑥𝑖}𝑛+1𝑖=0 where 𝜑(𝑛)(𝜃)=𝑥0+𝑥1𝛼1,1+𝛼1,2𝑞(𝜃)+𝑥2𝛼2,1𝑞(𝜃)+𝛼2,2cos2𝜋𝜃+𝑥3𝛼3,1cos2𝜋𝜃+𝛼3,2cos4𝜋𝜃++𝑥𝑛𝛼𝑛,1cos2𝜋(𝑛2)𝜃+𝛼𝑛,2cos2𝜋(𝑛1)𝜃+𝑥𝑛+1𝛼𝑛+1,1cos2𝜋(𝑛1)𝜃+𝛼𝑛+1,2cos2𝜋𝑛𝜃=𝑥0+𝑥1𝛼1,1𝑥+𝑞(𝜃)1𝛼1,2+𝑥2𝛼2,1+cos2𝜋𝜃𝑥2𝛼2,2+𝑥3𝛼3,1++cos2𝜋𝑖𝜃𝑥𝑖+1𝛼𝑖+1,2+𝑥𝑖+2𝛼𝑖+2,1++cos2𝜋(𝑛1)𝜃𝑥𝑛𝛼𝑛,2+𝑥𝑛+1𝛼𝑛+1,1+cos2𝜋𝑛𝜃𝑥𝑛+1𝛼𝑛+1,2.(5.28) Equating the last coefficients gives 𝑥𝑛+1=𝑎𝑛𝛼𝑛+1,2,𝑥𝑛=𝑎𝑛1𝑥𝑛+1𝛼𝑛+1,1𝛼𝑛,2.(5.29) Continuing, 𝑥𝑖=𝑎𝑖1𝑥𝑖+1𝛼𝑖+1,1𝛼𝑖,2,(5.30) and finally 𝑥1𝑥=2𝛼2,1𝛼1,2,𝑥0=𝑎0𝑥1𝛼1,1.(5.31) Hence, with the above choice of {𝑥𝑖}𝑛+1𝑖=0, 𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝜃=𝑛+1𝑖=0𝑥𝑖𝑓𝑖(𝜃),(5.32) and so 𝜑𝑛+1𝑖=0𝑥𝑖𝑓𝑖2=𝜑𝑛+1𝑖=0𝑥𝑖𝑓𝑖+𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝑛𝑖=0𝑎𝑖cos2𝜋𝑖2=𝜑𝑛𝑖=0𝑎𝑖cos2𝜋𝑖+𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝑛+1𝑖=0𝑥𝑖𝑓𝑖2𝜑𝑛𝑖=0𝑎𝑖cos2𝜋𝑖2+𝑛𝑖=0𝑎𝑖cos2𝜋𝑖𝑛+1𝑖=0𝑥𝑖𝑓𝑖2=𝜑𝑛𝑖=0𝑎𝑖cos2𝜋𝑖20(5.33) as 𝑛.

To find an approximate solution to the generalized Black-Scholes equation we start by letting 𝑋𝜋(𝑡)=𝐄[𝑓(𝑆𝑇)𝑆𝑡=𝜑] (from [6]) and approximating 𝜑 by 𝜑(𝑛)=𝑛+1𝑖=0𝑥𝑖𝑓𝑖.(5.34) We define the space 𝐂𝑛 as the set of all continuous functions that can be represented by this summation for some {𝑥𝑖}𝑖=0. Note that 𝐂𝑛𝐿2𝑛[,0]. Also define 𝑒𝑛𝑛+2 by 𝑒𝑛=𝑥𝑛+1𝑖=0𝑥𝑖𝑓𝑖,(5.35) so that Ψ(𝑡,𝜑(𝑛))=Ψ(𝑡,𝑒𝑛(𝑥)). Define Ψ𝑛[0,𝑇]×𝑛+2 by Ψ𝑛(𝑡,𝑥)=Ψ(𝑡,𝜑(𝑛)) provided that the 𝑥 is formed by the coefficients of 𝜑(𝑛) in the spanning set {𝑓𝑖}𝑖=0. In general, 𝑥(𝑡) is formed by the coefficients of 𝜑𝑡(𝑛) in the spanning set {𝑓𝑖}𝑖=0. Also, define 𝑣𝑛[,0] by 𝑣𝑛1(𝜃)=0,for𝜃,𝑛,1𝑛𝜃+1,for𝜃=𝑛.,0(5.36) Last, let 𝑔𝑛[0,𝑇]×𝐂𝑛×𝑛+1×𝐶1,2([0,𝑇]×𝐂) be defined by 𝑔𝑛𝑡,𝜑(𝑛),𝑥,Ψ=𝑟𝑛+1𝑖=0𝑥𝑖𝐷Ψ(𝑡,𝜑(𝑛))𝟏{0}𝑛+1𝑖=0𝑘𝑖𝜕𝜕𝑥𝑖Ψ𝑛+𝛿𝑡,𝑥22𝑛+1𝑖=0𝑥𝑖2𝐷2Ψ(𝑡,𝜑(𝑛))𝟏{0},𝟏{0}𝑛+1𝑖,𝑗=0𝑘𝑖𝑘𝑗𝜕2𝜕𝑥𝑖𝜕𝑥𝑗Ψ𝑛,𝑡,𝑥(5.37) where the 𝑘𝑖 are the coefficients of 𝑣𝑛 using the spanning set {𝑓𝑖}𝑖=0. Finally, define the operator ()𝑛𝐂𝐂𝑛 by (𝜑)𝑛=𝑛+1𝑖=0𝑥𝑖𝑓𝑖,(5.38) where the right-hand side is the first 𝑛+2 terms of the {𝑓𝑖}-expansion of 𝜑.

We are now ready for a theorem which enables us to approximate the solution of the infinite-dimensional Black-Scholes equation by solving a first-order real-valued partial differential equation and an equation similar to the generalized Black-Scholes equation but without the 𝐆(Ψ)(𝑡,𝜑𝑡) term. The lack of this term allows approximate solutions to be found using traditional techniques.

Theorem 5.2. Let 𝑆0=𝜓𝐂+ and 𝑡[0,𝑇]. Let 𝑓 be a 𝐶2(𝐂) function satisfying the conditions of Theorem 4.1 and let Λ=𝑓(𝑆𝑇). Then 𝑟Ψ𝑡,𝜑(𝑛)=𝜕Ψ𝜕𝑡𝑡,𝜑(𝑛)+𝐆(Ψ)𝑡,𝜑𝑡𝑛+𝐷Ψ𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2Ψ𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0},𝑡,𝜑(𝑛)[0,𝑇)×𝐂𝑛,(5.39) where Ψ𝑇,𝜑(𝑛)𝜑=𝑓(𝑛)𝜑(𝑛)𝐂𝑛(5.40) has a solution of the form 𝑉𝑡,𝜑𝑡𝑛𝐹𝑡,𝜑(𝑛).(5.41) Here, 𝑉(𝑡,(𝜑𝑡)𝑛)=𝑤𝑛(𝑡,0) is a solution to 𝐹𝑡,𝜑(𝑛)𝜕𝑤𝑛𝜕𝑡(𝑡,𝑢)+𝐹𝑡,𝜑𝑡𝑛𝜕𝑤𝑛𝜕𝑢(𝑡,𝑢)(5.42)+𝐆(𝐹)𝑡,𝜑𝑡𝑛𝑤𝑛(𝑡,𝑢)=0(5.43) for 𝑡[0,𝑇] and 𝑢[0,𝜖) for some 𝜖>0 and 𝑤𝑛(𝑇,0)=1, and 𝐹+×𝐂𝑛 is a solution of 𝑟𝐹𝑡,𝜑(𝑛)=𝜕𝐹𝜕𝑡𝑡,𝜑(𝑛)+𝐷𝐹𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2𝐹𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0}𝑡,𝜑(𝑛)[0,𝑇)×𝐂𝑛,(5.44) where 𝐹𝑇,𝜑(𝑛)𝜑=𝑓(𝑛)𝜑(𝑛)𝐂𝑛,(5.45) and 𝑓 is a uniformly bounded 𝐶2(𝐂) function satisfying the conditions of Theorem 4.1.

Proof. We assume a solution of the form Ψ(𝑡,𝜑(𝑛))=𝑉(𝑡,(𝜑𝑡)𝑛)𝐹(𝑡,𝜑(𝑛)), then𝑟𝑉𝑡,𝜑𝑡𝑛𝐹𝑡,𝜑(𝑛)=𝜕𝑉𝜕𝑡𝑡,𝜑𝑡𝑛𝐹𝑡,𝜑(𝑛)+𝐆(𝑉𝐹)𝑡,𝜑𝑡𝑛+𝐷(𝑉𝐹)𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2(𝑉𝐹)𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0}=𝐹𝑡,𝜑(𝑛)𝜕𝑉𝜕𝑡𝑡,𝜑𝑡𝑛+𝑉𝑡,𝜑𝑡𝑛𝜕𝐹𝜕𝑡𝑡,𝜑(𝑛)+𝐹𝑡,𝜑(𝑛)𝐆(𝑉)𝑡,𝜑𝑡𝑛+𝑉𝑡,𝜑𝑡𝑛𝐆(𝐹)𝑡,𝜑(𝑛)+𝑉𝑡,𝜑𝑡𝑛×𝐷(𝐹)𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2(𝐹)𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0},𝑡,𝜑(𝑛)[0,𝑇)×𝐂𝑛.(5.46) If 𝐹(𝑡,𝜑(𝑛)) is the solution to (5.44), then 𝐹𝑡,𝜑(𝑛)𝜕𝑉𝜕𝑡𝑡,𝜑𝑡𝑛+𝐹𝑡,𝜑𝑡𝑛𝐆(𝑉)𝑡,𝜑𝑡𝑛+𝐆(𝐹)𝑡,𝜑𝑡𝑛𝑉𝑡,𝜑𝑡𝑛=0.(5.47) Define 𝑇𝑢𝐂𝐂 by 𝑇𝑢(𝜑)=𝜑𝑢, that is, 𝑇𝑢 is a shift operator. Now let 𝑉(𝑡,(𝜑𝑡+𝑢)𝑛)=𝑉(𝑡,(𝑇𝑢(𝜑𝑡))𝑛)=𝑤𝑛(𝑡,𝑢) for a fixed 𝜑𝐂. Then 𝐆(𝑉)𝑡,𝜑𝑡𝑛=𝜕𝑤𝑛𝜕𝑢𝑡,0+,(5.48) where the superscript + denotes a right-hand derivative with respect to 𝑢. Then 𝐹𝑡,𝜑(𝑛)𝜕𝑤𝑛𝜕𝑡(𝑡,0)+𝐹𝑡,𝜑𝑡𝑛𝜕𝑤𝑛𝜕𝑢𝑡,0++𝐆(𝐹)𝑡,𝜑𝑡𝑛𝑤𝑛(𝑡,0)=0.(5.49)
A slightly more restrictive, but more familiar form is 𝐹𝑡,𝜑(𝑛)𝜕𝑤𝑛𝜕𝑡(𝑡,𝑢)+𝐹𝑡,𝜑𝑡𝑛𝜕𝑤𝑛𝜕𝑢(𝑡,𝑢)+𝐆(𝐹)𝑡,𝜑𝑡𝑛𝑤𝑛(𝑡,𝑢)=0,(5.50) where 𝑡[0,] and 𝑢[0,𝜖) for some 𝜖>0. There is the additional requirement that 𝑤𝑛(𝑇,0)=1 so that (5.44) holds.

Remark 5.3. It can be easily shown that 𝑆𝑡 is 𝛼-Hölder continuous a.s. for 0<𝛼<1/2 provided that 𝑆0 is 𝛼-Hölder continuous for the same 𝛼. Therefore, ||𝐹𝑛||𝑡,𝑥𝐹(𝑡,𝜑)0(5.51) for each 𝑡 as 𝑛 where 𝐹(𝑡,𝜑) is a solution to (5.44) and 𝐹𝑛(𝑡,𝑥)=𝐹(𝑡,𝜑(𝑛)) is an approximate solution, since 𝐹 is 𝐶2(𝐂) in its second variable and 𝐹𝑛𝑡,𝑥=𝐹𝑡,𝑒𝑛𝑥=𝐹𝑡,𝜑(𝑛).(5.52)

The proof of the following corollary is identical to that of Theorem 5.2, with the use of Remark 5.3 to obtain Ψ(𝑡,𝜑).

Corollary 5.4. If 𝑆0 is Hölder continuous, then 𝜕𝑟Ψ(𝑡,𝜑)=𝜕𝑡Ψ(𝑡,𝜑)+𝐆(Ψ)𝑡,𝜑𝑡+𝐷Ψ(𝑡,𝜑)𝑟𝜑(0)𝟏{0}+12𝐷2Ψ(𝑡,𝜑)𝑁(𝜑)𝟏{0},𝑁(𝜑)𝟏{0}[,(𝑡,𝜑)0,𝑇)×𝐂+,(5.53) where Ψ(𝑇,𝜑)=𝑓(𝜑)𝜑𝐂+(5.54) has a solution of the form 𝑉(𝑡,𝜑𝑡)𝐹(𝑡,𝜑). Here, 𝑉(𝑡,𝜑𝑡)=𝑤(𝑡,0) is a solution to 𝐹(𝑡,𝜑)𝜕𝑤𝜕𝑡(𝑡,𝑢)+𝐹𝑡,𝜑𝑡𝜕𝑤𝜕𝑢(𝑡,𝑢)+𝐆(𝐹)𝑡,𝜑𝑡𝑤(𝑡,𝑢)=0(5.55) for 𝑡[0,𝑇] and 𝑢[0,𝜖) for some 𝜖>0. 𝐹(𝑡,𝜑) is the solution to (5.44) where one lets 𝑛. In addition, 𝑤(𝑇,0)=1.

Now we must solve (5.44), which is done in the following theorem. With this solution, the first-order partial differential equation can be solved by traditional means.

Theorem 5.5. Let 𝑟Ψ𝑡,𝜑(𝑛)=𝜕Ψ𝜕𝑡𝑡,𝜑(𝑛)+𝐷Ψ𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2Ψ𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0}𝑡,𝜑(𝑛)[0,𝑇)×𝐂𝑛,(5.56) where Ψ𝑇,𝜑(𝑛)𝜑=𝑓(𝑛)𝜑(𝑛)𝐂𝑛,(5.57) and 𝑓 is a uniformly bounded 𝐶2(𝐂) function satisfying the conditions of Theorem 4.1. Let 𝑓𝑛𝑛+2 be defined by 𝑓𝑛=𝑓𝑒𝑛, then Ψ𝑛=𝑒𝑡,𝑥𝑟(𝑇𝑡)2𝜋𝑓𝑛𝛿exp𝑟𝐵22𝐵2(𝑇𝑡)+𝛿𝐵𝑦𝑒𝑇𝑡𝑥𝑦2/2+𝑑𝑦𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠.(5.58) Here, 𝑘𝐵=0𝑘0𝑘0𝑘1𝑘1𝑘1𝑘𝑛+1𝑘𝑛+1𝑘𝑛+1,𝑣𝑛=𝑛+1𝑖=0𝑘𝑖𝑓𝑖(5.59) from (5.36).

Proof. Since Ψ𝑛[0,𝑇]×𝑛+2, the definition of the Fréchet derivatives and the properties of the set {𝑓𝑖}𝑖=0 give 𝑟Ψ𝑛=𝜕𝑡,𝑥Ψ𝜕𝑡𝑛𝑡,𝑥+𝑟𝑛+1𝑖=0𝑥𝑖𝑛+1𝑖=0𝑘𝑖𝜕𝜕𝑥𝑖Ψ𝑛+𝛿𝑡,𝑥22𝑛+1𝑖=0𝑥𝑖2𝑛+1𝑖,𝑗=0𝑘𝑖𝑘𝑗𝜕2𝜕𝑥𝑖𝜕𝑥𝑗Ψ𝑛𝑡,𝑥+𝑔𝑛𝑡,𝜑𝑠(𝑛)[,𝑥,Ψ,𝑡,𝑥0,𝑇)×𝑛+2,Ψ𝑇,𝑥(𝑇)=𝑓𝑛.𝑥(𝑇)(5.60) The 𝑥(𝑇) consists of the first 𝑛+2 coefficients of 𝑆𝑇 in the set of functions {𝑓𝑖}. By the Feynman-Kac theorem (see [15, Theorem  5.7.6]), Ψ𝑛𝑡,𝑥=𝑒𝑟(𝑇𝑡)𝐄𝑓+𝑥(𝑇)𝑥(𝑡)=𝑥𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠,(5.61) where 𝑥(𝑡) is the solution to 𝑑𝑥𝑖(𝑡)=𝑏𝑖𝑡,𝑥(𝑡)𝑑𝑡+𝑛+1𝑗=0𝜎𝑖𝑗𝑑𝑊𝑡,𝑥(𝑡)(𝑗)(𝑡)(5.62) for 𝑖=0,1,,𝑛+1. Noting that 𝑥(𝑡)=𝑥, 𝑏𝑖𝑡,𝑥=𝑟𝑘𝑖𝑛+1𝑖=0𝑥𝑖,𝛿2𝑘𝑖𝑘𝑗𝑛+1𝑖=0𝑥𝑖2=𝑛+1𝑘=0𝜎𝑖𝑘𝜎𝑡,𝑥𝑗𝑘𝑡,𝑥(5.63) with 0𝑖,𝑗𝑛+1. Hence, 𝑏𝑘𝑡,𝑥=𝑟0𝑛+1𝑖=0𝑥𝑖𝑘1𝑛+1𝑖=0𝑥𝑖𝑘𝑛+1𝑛+1𝑖=0𝑥𝑖,(5.64) and so 𝑏𝑥𝑡,𝑥=𝑟𝐵0𝑥1𝑥𝑛+1.(5.65) Also, 𝜎𝜎𝑡,𝑥𝑡,𝑥=𝛿2𝑛+1𝑖=0𝑥𝑖2𝑘20𝑘0𝑘1𝑘0𝑘𝑛+1𝑘0𝑘1𝑘21𝑘1𝑘𝑛+1𝑘0𝑘𝑛+1𝑘1𝑘𝑛+1𝑘2𝑛+1.(5.66) Therefore, 𝜎𝑥𝑡,𝑥=𝛿𝐵0𝑥1𝑥𝑛+1(5.67) as well. Thus, 𝑑𝑥(𝑡)=𝑟𝐵𝑥(𝑡)𝑑𝑡+𝛿𝐵𝑥(𝑡)𝑑𝑊(𝑡)(5.68) which has the solution 𝛿𝑥(𝑡)=exp𝑟𝐵22𝐵2𝑡+𝛿𝐵𝑊(𝑡)𝑥(0).(5.69) Notice that 𝑏(𝑡,𝑥) and 𝜎(𝑡,𝑥) satisfy the conditions necessary for applying the Feynman-Kac formula. Also note that (5.58) satisfies the polynomial growth condition max0𝑡𝑇||Ψ𝑛||𝑡,𝑥𝑀1+𝑥2𝜇(5.70) for 𝑀>0 and 𝜇1 due to the boundedness of 𝑓𝑛 and 𝑔𝑛 for 𝑛 sufficiently large. Therefore, by (5.61), equation (5.58) follows.

Lemma 5.6. Let 𝐴[0,𝑇]×(𝑛+2)×(𝑛+2) be defined by 𝛿𝐴(𝑡,𝑦)=𝑟𝐵22𝐵2(𝑇𝑡)+𝛿𝐵𝑦𝑇𝑡,(5.71) where 𝑘𝐵=0𝑘0𝑘0𝑘1𝑘1𝑘1𝑘𝑛+1𝑘𝑛+1𝑘𝑛+1.(5.72) Then 𝐴(𝑡,𝑦)=𝑄𝐷(𝑡,𝑦)𝑄1 where 𝑄 is the (𝑛+2)×(𝑛+2) matrix defined by 𝑘𝑄=0𝛼121212𝑘1𝛼1002𝑘𝑛+1𝛼12,𝑄001=11112𝑘𝑛+1𝛼2𝑘𝑛+1𝛼2𝑘𝑛+1𝛼2𝛼𝑘𝑛+1𝛼2𝑘𝑛𝛼2𝑘𝑛𝛼2𝑘𝑛𝛼2𝑘𝑛𝛼2𝑘1𝛼2𝛼𝑘1𝛼2𝑘1𝛼2𝑘1𝛼,𝛿𝐷(𝑡,𝑦)=𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑦.𝑇𝑡00000000000(5.73) Here, 𝛼=𝑛+1𝑖=0𝑘𝑖.

Proof. It is clear that 𝐵2=𝛼𝐵, so 𝛿𝐴(𝑡,𝑦)=𝑟22𝛼(𝑇𝑡)+𝛿𝑦𝑇𝑡𝐵.(5.74) The matrix 𝐴(𝑡,𝑦) is an (𝑛+2)×(𝑛+2) matrix with 𝑛+2 eigenvalues. The 𝑛+1 eigenvectors associated with the eigenvalue 0 are 1212000,1201200,(5.75) and so forth. The remaining eigenvalue is 𝛿𝑟22𝛼(𝑇𝑡)+𝛿𝑦𝑇𝑡𝛼(5.76) with the associated eigenvector of 1𝛼𝑘0𝑘1𝑘2𝑘𝑛+1.(5.77) The matrix 𝑄 is simply a matrix with these eigenvectors for columns, the expression for 𝑄1 can easily be verified, and 𝐷(𝑡,𝑦)=𝑄1𝐴(𝑡,𝑦)𝑄.

Next, we will derive an approximate solution to (5.44) with the auxiliary condition being that for the standard European call option. By using Theorem 5.2, a significant piece of the approximate solution is then known.

Proposition 5.7. Let 𝜕𝑟Ψ(𝑡,𝜑)=𝜕𝑡Ψ(𝑡,𝜑)+𝐷Ψ(𝑡,𝜑)𝑟𝜑(0)1{0}+12𝐷2Ψ(𝑡,𝜑)𝑁(𝜑)1{0},𝑁(𝜑)1{0}[,(𝑡,𝜑)0,𝑇)×𝐂+,(5.78) where Ψ(𝑇,𝜑)=(𝜑(0)𝐾)+𝜑𝐂+,(5.79) and 𝐾 is the strike price of the option contract. Then Ψ𝑛𝑡,𝑥=𝑒𝑟(𝑇𝑡)𝑘𝑛+1𝛼[]exp𝑟𝛼(𝑇𝑡)𝑛+1𝑖=0𝑥𝑖Φ𝛿𝛼𝑇𝑡𝑌0𝐾Φ𝑌0+𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠,(5.80) where Ψ𝑛(𝑡,𝑥)Ψ(𝑡,𝜑) pointwise as 𝑛. Here, 1Φ(𝑢)=2𝜋𝑢𝑒𝑦2/2𝑌𝑑𝑦,0=ln𝛼𝐾/𝑘𝑛+1𝑛+1𝑖=0𝑥𝑖𝛿𝛼𝑟𝑇𝑡𝛿𝛿𝛼2𝑇𝑡.(5.81)

Proof. By Theorem 5.5, the equation 𝑟Ψ𝑡,𝜑(𝑛)=𝜕Ψ𝜕𝑡𝑡,𝜑(𝑛)+𝐷Ψ𝑡,𝜑(𝑛)𝑟𝜑(𝑛)(0)𝟏{0}+12𝐷2Ψ𝑡,𝜑(𝑛)𝑁𝜑(𝑛)𝟏{0}𝜑,𝑁(𝑛)𝟏{0},𝑡,𝜑(𝑛)[0,𝑇)×𝐂𝑛,(5.82) with Ψ𝑇,𝜑(𝑛)𝜑=𝑓(𝑛)𝜑(𝑛)𝐂𝑛,(5.83) and 𝑓 being a 𝐶2(𝐂) function has the solution Ψ𝑛=𝑒𝑡,𝑥𝑟(𝑇𝑡)2𝜋𝑓𝑛𝛿exp𝑟𝐵22𝐵2(𝑇𝑡)+𝛿𝐵𝑦𝑒𝑇𝑡𝑥𝑦2/2+𝑑𝑦𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠.(5.84) By Lemma 5.6, 𝛿𝐴(𝑡,𝑦)=𝑟𝐵22𝐵2(𝑇𝑡)+𝛿𝐵𝑦𝑇𝑡(5.85) may be expressed as 𝐴(𝑡,𝑦)=𝑄𝐷(𝑡,𝑦)𝑄1 where 𝑄, 𝑄1, and 𝐷(𝑡,𝑦) are defined in the lemma. Since 𝐷(𝑡,𝑦) is diagonal, 𝑒𝐷(𝑡,𝑦) is straightforward to find and 𝑒𝐴(𝑡,𝑦)=𝑄𝑒𝐷(𝑡,𝑦)𝑄1. Let 𝑣(𝑡,𝑦) be the (𝑛+2)th row of 𝑒𝐴(𝑡,𝑦), expressed as a column vector. Then 𝑣𝑘(𝑡,𝑦)=𝑛+1𝛼𝛿exp𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑦1111𝑇𝑡.(5.86) Let 𝑢𝑛+1(𝑡,𝑦,𝑥) be the (𝑛+2)th component of 𝑒𝐴(𝑡,𝑦)𝑥, we have that 𝑢𝑛+1=𝑣=𝑘𝑡,𝑦,𝑥(𝑡,𝑦)𝑥𝑛+1𝛼𝛿exp𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑦𝑇𝑡𝑛+1𝑖=0𝑥𝑖.(5.87) Therefore, Ψ𝑛=𝑒𝑡,𝑥𝑟(𝑇𝑡)2𝜋𝑌0𝑢𝑛+1𝑒𝑡,𝑦,𝑥𝐾𝑦2/2+𝑑𝑦𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠(5.88) with 𝑘𝑛+1𝛼𝛿exp𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑌0𝑇𝑡𝑛+1𝑖=0𝑥𝑖=𝐾(5.89) since [𝑢𝑛+1(𝑡,𝑦,𝑥)𝐾]+ is 𝐶2 in 𝑥 on the interval (𝑌0,), with the necessary Lipschitz properties. Solving for 𝑌0 we have 𝑌0=ln𝛼𝐾/𝑘𝑛+1𝑛+1𝑖=0𝑥𝑖𝛿𝛼𝑟𝑇𝑡𝛿𝛿𝛼2𝑇𝑡.(5.90) So, Ψ𝑛=𝑒𝑡,𝑥𝑟(𝑇𝑡)2𝜋𝑌0𝑘𝑛+1𝛼𝛿exp𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑦𝑇𝑡𝑛+1𝑖=0𝑥𝑖𝑒𝐾𝑦2/2+𝑑𝑦𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)=𝑒𝑑𝑠𝑟(𝑇𝑡)2𝜋𝑌0𝑘𝑛+1𝛼𝛿exp𝑟22𝛼𝛼(𝑇𝑡)+𝛿𝛼𝑦𝑇𝑡𝑛+1𝑖=0𝑥𝑖𝑒𝑦2/2𝑑𝑦𝐾𝑒𝑟(𝑇𝑡)𝑌1Φ0+𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠=𝑒𝑟(𝑇𝑡)𝑘𝑛+1𝛼𝛿exp𝑟22𝛼×𝛼(𝑇𝑡)𝑛+1𝑖=0𝑥𝑖12𝜋𝑌0𝑒𝑦2/2+𝛿𝛼𝑦𝑇𝑡𝑌𝑑𝑦𝐾1Φ0+𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠=𝑒𝑟(𝑇𝑡)𝑌1Φ0(𝑘𝐾)+𝑛+1𝛼𝛿exp𝑟22𝛼×𝛼(𝑇𝑡)𝑛+1𝑖=0𝑥𝑖𝑒𝛿2𝛼2(𝑇𝑡)/2𝑌1Φ0𝛿𝛼+𝑇𝑡𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠=𝑒𝑟(𝑇𝑡)𝑘𝑛+1𝛼[]exp𝑟𝛼(𝑇𝑡)𝑛+1𝑖=0𝑥𝑖Φ𝛿𝛼𝑇𝑡𝑌0𝐾Φ𝑌0+𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠.(5.91) This result is the solution given 𝜑(𝑛)𝐂𝑛. Notice that 𝑔𝑛𝑠,𝜑𝑠(𝑛),𝑥(𝑠),Ψ0(5.92) pointwise as 𝑛 since lim𝑛𝐷Ψ𝑡,𝜑(𝑛)𝟏{0}=lim𝑛𝐷Ψ𝑛𝑣𝑡,𝑥𝑛=lim𝑛𝑛+1𝑖=0𝑘𝑖𝜕𝜕𝑥𝑖Ψ𝑛,𝑡,𝑥(5.93) and similarly for the second derivatives. By dominated convergence, we have that 𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠0(5.94) pointwise as 𝑛. By Remark 5.3, we have that Ψ𝑛(𝑡,𝑥)Ψ(𝑡,𝜑) pointwise as 𝑛.

6. Summary and Conclusions

In this paper, we have continued [6] by deriving an infinite-dimensional Black-Scholes equation for the European option problem, where the (𝐵,𝑆)-market model is given by (2.7) and (2.8). The resulting deterministic partial differential equation is a new type of equation, one where the partial differentiation contains extended Fréchet derivatives. Given the (𝐵,𝑆)-market model equations, a spanning set for the space of square-integrable function is developed which simplifies finding an approximate solution to this equation. The solution method detail in this paper consists of the following steps.

Step 1. Given 𝑟 and 𝑁(), use (5.6)-(5.7) with (5.5) and (5.10)-(5.11) to find the spanning set {𝑓𝑖}𝑛+1𝑖=0 for 𝑛 sufficiently large. Coefficients in this spanning set of functions are found using (5.13)–(5.15).

Step 2. Use Theorem 5.5 or Proposition 5.7, depending on the reward function, to find Ψ𝑛(𝑡,𝑥). The term 𝑇𝑡𝑔𝑛𝑠,𝜑𝑠(𝑛)𝑒,𝑥(𝑠),Ψ𝑟(𝑠𝑡)𝑑𝑠(6.1) approaches zero as 𝑛 approaches infinity, so this term may be assumed small for sufficiently large 𝑛. The vector 𝑥 is found from (5.69).

Step 3. Having found Ψ𝑛(𝑡,𝑥), solve (5.43) for 𝑤𝑛(𝑡,0), then 𝑤𝑛(𝑡,0)Ψ𝑛(𝑡,𝑥) is an approximate solution to the generalized Black-Scholes equation. By Corollary 5.4, 𝑤𝑛(𝑡,0)Ψ𝑛(𝑡,𝑥)Ψ(𝑡,𝜑) pointwise as 𝑛.

Appendix

Proof of Proposition 3.2

In this appendix, we prove Proposition 3.2, which is stated again as follows.

Proposition  3.2. Let 𝜑𝐂and 𝑓𝐂. Further assume 𝑓𝐶2(𝐂)and let Ψ[0,𝑇]×𝐂be defined by Ψ(𝑡,𝜑)=𝑒𝑟(𝑇𝑡)𝐸𝑓𝑆𝑇𝑆𝑡.=𝜑(A.1)Then if 𝐷𝑓and 𝐷2𝑓are globally Lipschitz, then so is 𝐷2Ψ.

Before proving the proposition, an additional result for Fréchet derivatives is needed.

Lemma A.1. Let 𝐻[0,𝑇]××𝐂𝐂 and 𝑓𝐂. Define 𝑔[0,𝑇]×𝑅×𝐂 by 𝑔(𝑡,𝑦,𝜑)=(𝑓𝐻)(𝑡,𝑦,𝜑) for any 𝑡[0,𝑇], 𝑦, and 𝜑𝐂. Assume that 𝑓 has a second Fréchet derivative and likewise for 𝐻 (with respect to the third variable.) Then 𝐷2𝑔(𝑡,𝑦,𝜑)(𝜙,𝜙)=𝐷2𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷𝑓(𝐻(𝑡,𝑦,𝜑))2.𝐻(𝑡,𝑦,𝜙)(𝜙,𝜙)(A.2)

Proof. We start by considering 𝐷𝑔(𝑡,𝑦,𝜑+𝜙)(𝜙)𝐷𝑔(𝑡,𝑦,𝜑)(𝜙)=𝐷𝑓(𝐻(𝑡,𝑦,𝜑+𝜙))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙))=𝐷𝑓(𝐻(𝑡,𝑦,𝜑+𝜙))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙)).(A.3) From here we have that𝐷𝑓(𝐻(𝑡,𝑦,𝜑+𝜙))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))𝐷𝑓(𝐻(𝑡,𝑦,𝜑)(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)))=𝐷2𝑓(𝐻(𝑡,𝑦,𝜑))(𝐻(𝑡,𝑦,𝜑+𝜙)𝐻(𝑡,𝑦,𝜑),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝑜1(𝐻(𝑡,𝑦,𝜑+𝜙)𝐻(𝑡,𝑦,𝜑),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))=𝐷2𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷2𝑜𝑓(𝐻(𝑡,𝑦,𝜑))2(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)+𝑜1(𝐻(𝑡,𝑦,𝜑+𝜙)𝐻(𝑡,𝑦,𝜑),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))=𝐷2𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷2𝑓𝑜(𝐻(𝑡,𝑦,𝜑))2(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)+𝑜1𝐷𝐻(𝑡,𝑦,𝜑)(𝜙)+𝑜3(.𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)(A.4) Also, 𝐷𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)𝐷𝐻(𝑡,𝑦,𝜑)(𝜙))=𝐷𝑓(𝐻(𝑡,𝑦,𝜑))2(𝐻(𝑡,𝑦,𝜑)(𝜙,𝜙))+𝑜4.(𝜙,𝜙)(A.5) Clearly, 𝐷2𝑜𝑓(𝐻(𝑡,𝑦,𝜑))2(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)=𝑜5(𝜙,𝜙),(A.6) since the second derivative is linear. Also 𝑜1𝐷𝐻(𝑡,𝑦,𝜑)(𝜙)+𝑜3(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)0(A.7) as 𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)0 and 𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)𝜙𝐷𝐻(𝑡,𝑦,𝜑+𝜙)𝜙𝜙=𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)𝐾1(A.8) for some 𝐾1<. Furthermore, 𝑜1𝐷𝐻(𝑡,𝑦,𝜑)(𝜙)+𝑜3(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)+𝑜3(𝜙)0(A.9) as 𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)+𝑜3(𝜙)0, and 𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)+𝑜3(𝜙)𝜙𝐷𝐻(𝑡,𝑦,𝜑+𝜙)𝜙+𝜙𝑜3(𝜙)𝑜𝜙=𝐷𝐻(𝑡,𝑦,𝜑+𝜙)+3(𝜙)𝜙𝐾2+𝑜3(𝜙)𝜙(A.10) for some 𝐾2<. Therefore 𝑜1𝐷𝐻(𝑡,𝑦,𝜑)(𝜙)+𝑜3(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙)=𝑜6(𝜙,𝜙).(A.11) Finally, 𝑜𝐷𝑓(𝐻(𝑡,𝑦,𝜑))4𝑜(𝜙,𝜙)𝐷𝑓(𝐻(𝑡,𝑦,𝜑))4(𝜙,𝜙)𝐾3𝑜4(𝜙,𝜙)(A.12) for some 𝐾3<, so 𝑜𝐷𝑓(𝐻(𝑡,𝑦,𝜑))4(𝜙,𝜙)=𝑜7(𝜙,𝜙).(A.13)
Since 𝐷2𝐷𝑓(𝐻(𝑡,𝑦,𝜑))(𝐷𝐻(𝑡,𝑦,𝜑)(𝜙),𝐷𝐻(𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷𝑓(𝐻(𝑡,𝑦,𝜑))2𝐻(𝑡,𝑦,𝜙)(𝜙,𝜙)(A.14) is bounded and linear, we are done.

Proof of Proposition 3.2. From Proposition 3.1, we have that 𝑒Ψ(𝑡,𝜑)=𝑟(𝑇𝑡)2𝜋𝑓(𝐻(𝑇𝑡,𝑦,𝜑))𝑒𝑦2/2=𝑒𝑑𝑦𝑟(𝑇𝑡)2𝜋𝑔(𝑇𝑡,𝑦,𝜑)𝑒𝑦2/2𝑑𝑦,(A.15) where 𝐻[0,𝑇]××𝐂𝐂 is defined by 𝑆𝑡=𝐻(𝑡,𝑊(𝑡),𝑆0), and 𝑔[0,𝑇]××𝐂 is defined by 𝑔(𝑡,𝑊(𝑡),𝑆0)=(𝑓𝐻)(𝑡,𝑊(𝑡),𝑆0). In the same proposition, we see that 𝑒𝐷Ψ(𝑡,𝜑)(𝜙)=𝑟(𝑇𝑡)2𝜋𝐷𝑔(𝑇𝑡,𝑦,𝜑)(𝜙)𝑒𝑦2/2𝐷𝑑𝑦,2𝑒Ψ(𝑡,𝜑)(𝜙,𝜙)=𝑟(𝑇𝑡)2𝜋𝐷2𝑔(𝑇𝑡,𝑦,𝜑)(𝜙,𝜙)𝑒𝑦2/2𝑑𝑦.(A.16) If 𝐷2𝑔(𝑇𝑡,𝑦,𝜑) is globally Lipschitz, then it is clear that 𝐷2Ψ(𝑡,𝜑) is also globally Lipschitz. We will now look at 𝐷2𝑔(𝑇𝑡,𝑦,𝜑)(𝜙,𝜙).
By Lemma A.1, we have 𝐷2𝑔(𝑇𝑡,𝑦,𝜑)(𝜙,𝜙)=𝐷2𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑)(𝜙),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)(𝜙))+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2,𝐷𝐻(𝑇𝑡,𝑦,𝜙)(𝜙,𝜙)2𝑔(𝑇𝑡,𝑦,𝜑+𝜓)𝐷2𝑔𝐷(𝑇𝑡,𝑦,𝜑)2𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))𝐷2+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙))𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2=𝐷𝐻(𝑇𝑡,𝑦,𝜑)2𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))𝐷2𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))+𝐷2𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))𝐷2+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙))𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐷𝐻(𝑇𝑡,𝑦,𝜑)2𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))𝐷2+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))2𝑓(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓))𝐷2𝑓+𝐷(𝐻(𝑇𝑡,𝑦,𝜑))(𝐷𝐻(𝑇𝑡,𝑦,𝜑),𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙))𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2+𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐷𝐻(𝑇𝑡,𝑦,𝜑)2𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))𝐷2×𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)+2𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑)+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐻𝐷(𝑇𝑡,𝑦,𝜑+𝜓)+𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷2.𝐻(𝑇𝑡,𝑦,𝜑)(A.17) Taking the terms one at a time, 𝐷2𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))𝐷2𝐶𝑓(𝐻(𝑇𝑡,𝑦,𝜑))×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)1𝜓,(A.18) since 𝐷2𝑓 is globally Lipschitz and there is some 𝐶1< such that 𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐶1 and likewise for 𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓).
For the next term, 𝐷2𝑓(𝐻(𝑇𝑡,𝑦,𝜑))×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑)𝐶2𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑).(A.19) But 𝐻(𝑇𝑡,𝑦,𝜑) is linear in 𝜑 as we will now show.
Again using the operator define by ()𝑛𝐂𝐂𝑛 with (𝜑)𝑛=𝑛+1𝑖=0𝑥𝑖𝑓𝑖,(A.20) where the right-hand side is the first 𝑛+2 terms of the {𝑓𝑖}-expansion of 𝜑, we have that 𝑆𝑡𝑛=𝑒𝑛1,𝑥(𝑡)(A.21) where 𝑒𝑛1 is linear. Recall that 𝛿𝑥(𝑡)=exp𝑟𝐵22𝐵2𝑡+𝛿𝐵𝑊(𝑡)𝑥(0)=𝐴𝑡,𝑊(𝑡)𝑥(0),(A.22) where 𝑥(0)=(𝑆0)𝑛. Therefore, 𝑥(𝑡)=𝑛+1𝑖=0𝐴0,𝑖𝑥𝑡,𝑊(𝑡)𝑖𝑛+1𝑖=0𝐴1,𝑖𝑊𝑥𝑡,(𝑡)𝑖𝑛+1𝑖=0𝐴2,𝑖𝑥𝑡,𝑊(𝑡)𝑖𝑛+1𝑖=0𝐴𝑛+2,𝑖𝑊𝑥𝑡,(𝑡)𝑖.(A.23) So, 𝑆𝑡=lim𝑛𝑛𝑖=0lim𝑘𝑘𝑗=0𝐴𝑖,𝑗𝑥𝑡,𝑊(𝑡)𝑗𝑓𝑖(A.24) is linear in 𝑆0=lim𝑛𝑛𝑖=0𝑥𝑖𝑓𝑖.
Since 𝐻(𝑇𝑡,𝑦,𝜑) is linear in 𝜑, 𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)𝐶3𝜓,𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑)𝐶4𝜓.(A.25) Since 𝜓𝑀<, we have that 𝐷2𝑓(𝐻(𝑇𝑡,𝑦,𝜑))𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜙)×𝐷𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷𝐻(𝑇𝑡,𝑦,𝜑)𝐶5𝜓.(A.26)
We also have that 𝐷𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑+𝜓))𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐶6𝜓(A.27) for some 𝐶6< since 𝐷𝑓 is globally Lipschitz. Finally, 𝐷𝐷𝑓(𝐻(𝑇𝑡,𝑦,𝜑))2𝐻(𝑇𝑡,𝑦,𝜑+𝜓)𝐷2𝐻(𝑇𝑡,𝑦,𝜑)𝐶7𝜓(A.28) for some 𝐶7< since 𝐻(𝑇𝑡,𝑦,𝜑) is linear in 𝜑. Combining these, we have that 𝐷2𝑔(𝑇𝑡,𝑦,𝜑) is globally Lipschitz and therefore so is 𝐷2Ψ(𝑡,𝜑).