Abstract

We study the valuation of American-type derivatives in the stochastic volatility model of Barndorff-Nielsen and Shephard (2001). We characterize the value of such derivatives as the unique viscosity solution of an integral-partial differential equation when the payoff function satisfies a Lipschitz condition.

1. Introduction

In their seminal paper, Barndorff-Nielsen and Shephard [1] introduced a model that has been shown to describe particularly well financial assets for which log-returns have heavy tail distributions and display long-range dependence. In this model, the volatility of the asset is described by an Ornstein-Uhlenbeck-type process with a pure jump Lévy process acting as the background driving process. An empirical study was made in [1] and showed from exchange rate data that suitable distributions for the Lévy process are the so-called generalized inverse gaussian distributions from which well-understood examples are the normal inverse gaussian (studied in [2]) and the gamma distribution.

The BNS model has been studied from different points of view. Benth et al. [3] considered the problem of optimal portfolio selection. Nicolato and Vernados [4] have studied European option pricing and described the set of equivalent martingale measures under this model. To evaluate these types of options, the authors propose the transform-based method and a simple Monte Carlo method.

In this paper, we consider the pricing of American options with the use of integral-partial differential equations (IPDEs). Although our technique can be simplified and used for European options and certain path-dependent options such as barrier options (see [5] for a definition and examples), we will mainly concentrate on American type derivatives which have not been studied for this model. The main difficulty in this case is the lack of Lipschitz continuity of some of the coefficients of the IPDE.

The question of whether observed option prices can be calibrated and shown to reproduce stylized features such as smiles using this model is of great practical and theoretical relevance. However the majority of exchange-traded options are of the American type and there do not exist any established numerical methods to compute option prices for this model. One could use Monte-Carlo methods; however it is generally more efficient to characterize the option value function as a solution of a variational inequality and to discretize this inequality in order to get an approximation of the value function. Whereas the existence of a solution to the IPDE suggests the use of finite difference schemes, the uniqueness of the solution is a particularly crucial property which insures the convergence of any such numerical scheme to the correct value function. The characterization of the value function as the unique solution of an IPDE is thus the first step to achieve this goal. The design and implementation of a numerical scheme are beyond the scope of this paper; however we refer the reader to the paper of Levendorskiĭ et al. [6] and references therein for some ideas on how this problem could be approached.

The connection between viscosity solutions of IPDEs and Lévy processes has been studied in the literature by various authors. Pham [7] considered a general stopping time problem of a controlled jump diffusion processes. However, his results do not apply here because the Lipschitz condition on the coefficients is not satisfied in our current setting. Cont and Voltchkova [5] studied barrier options and Barles et al. [8] established the connection between viscosity solutions and backward stochastic differential equations. In these papers, the stock price considered is modeled by a stochastic differential equation with jumps driven by a Lévy process. The main difference between the BNS model and these models is the presence of stochastic volatility. However, we will see that the lack of smoothness of the solution to our IPDE will also lead us to consider the notion of viscosity solutions as presented in [9].

The rest of the paper is organized as follows. In Section 2, we present the model and recall the results of Nicolato and Vernados [4] regarding the set of equivalent martingale measures. Section 3 is devoted to the continuity of the value function. In Section 4, we prove that the value function is the viscosity solution of the associated IPDE, and the uniqueness of the solution is presented in Section 5.

2. Lévy Processes and the BNS Model

Let We consider the stochastic volatility model of Barndorff-Nielsen and Shephard [1] for the price process of an asset, denoted by and defined on a filtered probability space . We thus assume that the log-return of the asset satisfies the following stochastic differential equation: with in which , and is a Brownian motion, and is the background driving Lévy process (BDLP) under the physical measure . In this model, has no gaussian component and the increments are positive. and are assumed to be independent, and is the usual filtration generated by the pair . The positivity of the jumps of insure that the process is always positive. We denote by the Lévy measure of .

Suppose that is a probability measure equivalent to under which is a martingale. We are interested in American-type derivatives of the form in which is the payoff function, and is the set of all stopping times with values less or equal to . Since and are Markov processes, can be written as a function of , say, in which is the process for which and We also denote by the process starting at and at

2.1. Equivalent Martingale Measures

We start by summarizing the results of Nicolato and Vernados [4] concerning the set of equivalent martingale measures. In order to do so, we define the set and as the set of all equivalent martingale measures such that is still a Lévy process under independent of , possibly with a different marginal distribution.

As in [4], we impose the following conditions on the process : (C1)the process is given by the characteristic triplet so that the cumulant transform is given by for values of , for which this expression is defined; (C2); (C3).

Remark 2.1. Assumption (C2) implies that there exists such that For and , we have so that Furthermore, Assumption (C2) is a sufficient condition for the process to have finite moments of all orders.

The following theorem was proved in [4].

Theorem 2.2. For all there exists such that in which and and are, respectively, a Brownian motion and a Lévy process under . is the Lévy density of under and is the cumulant function.

In the remaining part of this paper, all expectations will be with respect to a chosen EMM , unless specified otherwise, and and will denote the associated Lévy measure and the Brownian motion associated to .

Let and assume for a moment that is Lipschitz in and that is is differentiable with respect to and , and twice differentiable with respect to . We can then apply Itô's formula to to find in which and is the -martingale given by in which , and is the random measure of the process . Since is a -martingale, if it can be shown that is also a martingale, we can then expect to satisfy the following integral-partial differential equation (IPDE): if Otherwise and this IPDE can be written as It is clear also that the function satisfies

Condition (2.10) is in fact very restrictive and most of the time not satisfied. Despite this problem, we will see that can still be regarded as a solution of this equation in a weaker sense.

3. Continuity of the Value Function

Recall the definition of the value of an American option with payoff : In the rest of this paper, we will assume that is positive and satisfies the Lipschitz condition, in other words such that For instance, the payoff function for an American put with strike is and satisfies this condition.

Our goal is to show that the function satisfies the IPDE (2.15) in some weak sense. In order to give meaning to this IPDE for a function that does not satisfy basic differentiability conditions, we introduce the idea of viscosity solutions following Crandall and Lions [10]. Let be the set of functions that satisfy

Definition 3.1. The function is a viscosity subsolution (supersolution) of (2.15)-(2.16) if and such that (i) and (ii) The function is a viscosity solution if it is both subsolution and supersolution.

Remark 3.2. As noted in [5, page 317] the condition is sufficient to have a well-defined integral term in In fact if , then for any .

An important property of viscosity solutions is the continuity of the function. It is the content of the following proposition.

Proposition 3.3. When satisfies the Lipschitz condition (3.2), the function is continuous and in .

Proof. In this proof, we will assume for simplicity that . The generalization to is straightforward. Throughout, is a positive constant that can change from line to line.
We start by showing the continuity of with respect to , uniformly in . We have the following representation of the volatility process: We define the integrated variance process started with by By (2.2), we find that so that we have the following representation of the integrated variance process in which .
We also have the following identities: with and .
Using the Lipschitz condition on , we obtain
Then,
Letting , the -field generated by the BDLP up to the maturity , we find that is a -martingale. Thus, is a -submartingale and Doob's theorem applies. In other words, Also,
And thus we proved the continuity of in uniformly in since In particular because of the following inequality:
The next step of the proof is to show that as This is easily obtained by first observing that As for the process , Since for all , and we need to show that when
We mentioned earlier that condition (C2) implies that the moments of are finite for all orders. Thus is uniformly integrable. Since is also continuous in probability, it is continuous in , and the conclusion follows.
Let us now show continuity with respect to time. Let . Take and define Then, From this inequality, we readily find that which converges to zero as
Global continuity follows from the following inequality: and the fact that the first bound is independent of .

4. Viscosity Solutions

This section is devoted to the viscosity solution property of the value function . In order to prove that is a viscosity solution of (2.15), we need the following dynamic programming principle. It is a consequence of the martingale property of the Snell envelope stopped before its optimal stopping time and it is the key property needed in the proof of the subsolution property.

Lemma 4.1. Let , , and define the stopping time Then,

Proof. For some constant , we have that for all . We know that and that from (3.11). As a result, + + + + for some constant large enough. Hence for all As a consequence, converges to as grows to infinity, that is, the collection is uniformly integrable. Hence we find that the process is of Class D, and we can apply the results of [11] to get the result.

The proof of the solution property of makes use of the following lemma.

Lemma 4.2. Let and Suppose that . Then when

Proof. Let such that .
First we show that almost surely. For some sequence , for large enough. In this case, since converges to in , we can take a subsequence if necessary and find that and a.s. with . Taking the limit, we find that
Since is continuous with respect to , we find that for small enough. Then, for small enough, for some constants and . By the continuity in probability of the processes and , we know that this expression goes to zero when

We can now show that is a viscosity solution.

Theorem 4.3. When satisfies the Lipschitz condition (3.2), is a viscosity solution of IPDE (2.15).

Proof. We already know that is continuous and in .
Let us start by showing that is a supersolution of (2.15). Let and satisfy the conditions given in the above definition of supersolutions. By definition, for any in which is the martingale defined by Since is also a martingale, we have the following inequality: in other words, dividing by and taking the limit as , Since, by definition, , satisfies (2.15). To prove that is a viscosity subsolution of (2.15), let and satisfy the conditions of the above definition for subsolutions. If , inequality (3.4) is satisfied. Otherwise, let We know from Lemma 4.1 that for any . Knowing that when by Lemma 4.2, dividing the preceding inequality by and taking the limit to , we get the desired result by Lebesgue's dominated convergence theorem.

5. Comparison Principles and Uniqueness of the Solution

In this section, we prove a comparison result from which we obtain the uniqueness of the solution of the IPDE. In proving comparison results for viscosity solutions, the notion of parabolic superjet and subjet as defined by Crandall et al. [10] is particularly useful. Setting , we define the parabolic superjet and its closure by The subjet and its closure are then defined similarly by

We then have the following lemma which is essentially proved in [8] (Lemma ).

Lemma 5.1. If the function is a viscosity subsolution (resp. supersolution) of (2.15), then for all and for all (resp., ) in which for some and .

Pham [7] obtains the uniqueness of the solution when the coefficients of satisfy Lipschitz conditions on For , define . Then, the coefficients of satisfy the Lipschitz conditions on , and using the ideas of uniqueness proofs in the literature, we can show a comparison principle on This result will then be used to show the uniqueness of the solution on

Theorem 5.2. Let , and let be a subsolution and a supersolution of (2.15) on such that for or Then for all .

Proof. An IPDE of the form for and was shown to have a unique solution in [8] when the coefficients of satisfy some given Lipschitz conditions. In fact when and , we have , and so the operator satisfies the assumptions made in [8]. The extension of the uniqueness result to our current setting is straightforward, and we only give the main details.
We first show that is a subsolution of a related IPDE. Suppose that and attains a maximum at . Set Since and are in , the function attains its maximum (which depends on ) in By a classical argument in the theory of viscosity solutions, we can show that when and when
Applying Theorem of Crandall et al. [9] to the functions and , we find matrices such that with and and for the inequalities are satisfied. Write these two expressions as and Then . Now, and because is Lipschitz when . Thus On the other hand it was shown in [8] that in which Using the fact that we find by letting and then that Consequently,
As shown in [8] (see Lemma ), there exists a function such that for which the maximum is attained at some point . Then Let Then satisfies the properties in the subsolution definition, hence it satisfies (5.12). But Hence, either or or and, in this case, by assumption. Hence, we conclude that Sending to zero we get on . As done in [8], we can repeat this argument as many times as needed to get on .

A solution of (2.15)-(2.16) is said to be minimal if it is less or equal to any other solution of (2.15)-(2.16). To prove uniqueness, we first show that the solution is minimal.

Theorem 5.3. is the minimal viscosity solution of (2.15)-(2.16).

Proof. Let and define in which Then is a viscosity solution of (2.15) on with boundary conditions The proof of this statement is essentially the same as the proof for the viscosity solution property of . The main difference is that the maturity is replaced by Note that for hence for all and
Let be another viscosity solution of (2.15)-(2.16). Then is a viscosity solution of (2.15) on with boundary values for or Also, for or By Theorem 5.2, we find that on In particular, for and Since is arbitrary, on

Following Pham [7], we denote by the set of functions defined on uniformly continuous in , uniformly in . We have already shown that the function satisfies Hence, . Using the two previous theorems, we can show the uniqueness in .

Theorem 5.4. is the unique viscosity solution of (2.15)-(2.16) in .

Proof. Let be another viscosity solution of (2.15)-(2.16). Let Then there exists such that and for In particular, for all , all and Furthermore, by Theorem 5.3, we obtain that and by definition. By the comparison principle of Theorem 5.2, we find that for all Hence, for all Since is arbitrary, we obtain the desired result.

To prove the uniqueness of the solution, we first modified the optimal stopping problem by defining it on in order to avoid the degeneracy of the infinitesimal generator. This suggests that one may be able to use known numerical schemes for nondegenerate IPDEs for Lévy processes to design a numerical scheme for the value function . When is close to , we obtain an approximation for the value function . The reader is referred to the work of Levendorskiĭ et al. [6] and references therein for possible numerical implementation techniques and difficulties associated to them. On the other hand, it is likely that the method of proof of the uniqueness property could be used for a larger class of stochastic volatility models, including the models of Heston [12] and Hull and White [13]. Indeed, all these models have the non-Lipschitz term in the differential equation for the log-returns. There remain open problems which we leave to future research.

Acknowledgment

Some parts of this paper were done at HEC Montréal, and the author would like to thank Bruno Rémillard and an anonymous referee for their help. This work was supported in part by the NSERC, the IFM2 and the Fonds québécois de la recherche sur la nature et les technologies.