Abstract

This paper provides new explicit results for some boundary crossing distributions in a multidimensional geometric Brownian motion framework when the boundary is a piecewise constant function of time. Among their various possible applications, they enable accurate and efficient analytical valuation of a large number of option contracts traded in the financial markets belonging to the classes of barrier and look-back options.

1. Introduction

The joint law of the maximum (or the minimum) of a real-valued Brownian motion and its endpoint over a finite time interval is a central result in the study of Brownian motion, particularly with regard to the many applications of the theory in finance, medical imaging, robotics, and biology. It can be obtained as a consequence of the “reflection principle,” which derives from the strong Markov property of Brownian motion (Freedman [1]). Application of Girsanov’s theorem easily generalizes this seminal result to the case of a geometric Brownian motion (GBM), a frequently encountered diffusion process that is the building block of financial engineering. Alternatively, the law can be derived by a partial differential equation approach, using Kolmogorov’s equation for the transition density function of a diffusion process. The distribution of the first passage time by a one-dimensional GBM to a one-sided or a two-sided straight boundary then follows. A few cases where the boundary is curved have been handled (Barba Escribá [2]; Salminen [3]; Kunitomo and Ikeda [4]). For a comprehensive source of formulae, one may refer to Borodin and Salminen [5].

Various extensions to these results are often needed to solve practical engineering problems. In particular, one may look for joint distributions of the highest or lowest points hit over several time intervals, and one may deal with a multidimensional GBM. There are few known explicit results in these more general settings, partly because of the laborious analytical calculations involved, but also due to numerical obstacles: the rapidly increasing dimension of the boundary crossing problem leads to analytical solutions that are expressed in terms of functions that are hard to compute with accuracy. As far as one-sided boundaries are concerned, formulae have been published for the joint law of a sequence of maxima or minima of a one-dimensional GBM over several time intervals (Guillaume [6]), as well as for the joint law of the maxima or minima of a two-dimensional GBM over one time interval (Iyengar [7]; He et al. [8]). A formula for the joint law of the exit times of a one-dimensional GBM from two successive two-sided boundaries is also known (Guillaume [9]).

This paper focuses on a sequence of two one-sided straight boundaries conditional on two correlated GBMs, while the value of a third correlated GBM is taken into account at the endpoint of the time interval. The state space is thus three-dimensional. The choice of this particular distribution is motivated both by its usefulness in financial engineering applications and by the fact that it leads to tractable analytical solutions that can be computed with great accuracy and efficiency.

Section 2 presents the two main formulae of this paper. Section 3 deals with applications of these formulae and discusses their numerical implementation.

2. Main Formulae

This section contains the two main formulae of the paper. Let {𝑆1(𝑡),𝑡0} and {𝑆2(𝑡),𝑡0} be two geometric Brownian motions with constant drift coefficients, 𝛼1 and 𝛼2, respectively, under a probability measure , constant positive diffusion coefficients 𝜎1 and 𝜎2, respectively, and constant correlation coefficient 𝜌12. In other words, the processes 𝑆1(𝑡) and 𝑆2(𝑡) evolve in time according to the following dynamics: 𝑑𝑆1(𝑡)=𝛼1𝑆1(𝑡)𝑑𝑡+𝜎1𝑆1(𝑡)𝑑𝐵1(𝑡),𝑑𝑆2(𝑡)=𝛼2𝑆2(𝑡)𝑑𝑡+𝜎2𝑆2(𝑡)𝑑𝐵2(𝑡),(2.1) where 𝐵1(𝑡) and 𝐵2(𝑡) are standard real-valued Brownian motions and 𝑑[𝐵1,𝐵2](𝑡)=𝜌12𝑑𝑡. Let 𝐻1, 𝐻2, 𝐾1, 𝐾2, 𝐾3 be positive real numbers and 𝑃[](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2) defined as one of the four following cumulative distribution functions, where 𝑡2𝑡1𝑡0=0: (1)𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,(2.2)(2)𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,(2.3)(3)𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,(2.4)(4)𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3.(2.5)

Let us introduce the following notations:𝑘1𝐾=ln1𝑆1𝑡0,𝑘2𝐾=ln2𝑆2𝑡0,𝑘3𝐾=ln3𝑆2𝑡0,1𝐻=ln1𝑆1𝑡0,2𝐻=ln2𝑆2𝑡0,𝜇1=𝛼1𝜎212,𝜇2=𝛼2𝜎222.(2.6) The next proposition provides an exact formula for the four above-mentioned cumulative distribution functions.

Proposition 2.1. Let 𝑁3[,,;𝜃12,𝜃13,𝜃23] denote the joint trivariate cumulative distribution function of three standard normal random variables 𝑋1, 𝑋2, 𝑋3, where 𝜃𝑎𝑏 is the correlation coefficient between 𝑋𝑎 and 𝑋𝑏, (𝑎,𝑏){1,2,3}.
Then, for the up-and-up and the down-and-down distributions, one has P[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑁3𝜆𝑘1𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2;𝜌12,𝑡1𝑡2𝜌12,𝑡1𝑡2exp2𝜇22𝜎22𝑁3𝜆𝑘1𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝑘,𝜆322𝜇2𝑡2𝜎2𝑡2;𝜌12,𝜌12𝑡1𝑡2,𝑡1𝑡2exp2𝜇11𝜎21×𝑁3𝜆𝑘121𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2𝜌1221𝜎1𝑡2;𝜌12,𝜌12𝑡1𝑡2,𝑡1𝑡2+exp2𝜇1𝜎214𝜇2𝜌12𝜎1𝜎21+2𝜇22𝜎22×𝑁3𝜆𝑘121𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1,𝜆𝑘322𝜇2𝑡2𝜎2𝑡2+𝜌1221𝜎1𝑡2;𝜌12,𝜌12𝑡1𝑡2,𝑡1𝑡2,(2.7) where 𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2,𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2.(2.8) And, for the up-and-down and the down-and-up distributions, one has 𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑁3𝜆𝑘1𝜇1𝑡1𝜎1𝑡1,𝜆𝑘2+𝜇2𝑡1𝜎2𝑡1,𝜆𝑘3+𝜇2𝑡2𝜎2𝑡2;𝜌12,𝑡1𝑡2𝜌12,𝑡1𝑡2exp2𝜇22𝜎22×𝑁3𝜆𝑘1𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2,𝜆𝑘2𝜇2𝑡1𝜎2𝑡1,𝜆𝑘3+22+𝜇2𝑡2𝜎2𝑡2;𝜌12,𝑡1𝑡2𝜌12,𝑡1𝑡2exp2𝜇11𝜎21×𝑁3𝜆𝑘121𝜇1𝑡1𝜎1𝑡1,𝜆𝑘2+𝜇2𝑡1𝜎2𝑡1+𝜌1221𝜎1𝑡1,𝜆𝑘3+𝜇2𝑡2𝜎2𝑡2+𝜌1221𝜎1𝑡2;𝜌12,𝑡1𝑡2𝜌12,𝑡1𝑡2+exp2𝜇1𝜎214𝜇2𝜌12𝜎1𝜎21+2𝜇22𝜎22×𝑁3𝜆𝑘121𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2,𝜆𝑘2𝜇2𝑡1𝜎2𝑡1+𝜌1221𝜎1𝑡1,𝜆𝑘3+22+𝜇2𝑡2𝜎2𝑡2𝜌1221𝜎1𝑡2;𝜌12,𝑡1𝑡2𝜌12,𝑡1𝑡2,(2.9) where 𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2,𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2=𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2.(2.10)

Corollary of Proposition 2.1
The four following joint cumulative distribution functions, that will be useful in Section 3, are deduced from Proposition 2.1:(1)sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3=𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,(2.11)(2)inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3=𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,(2.12)(3)inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3=𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,(2.13)(4)sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3=𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3.(2.14)

Proof of Proposition 2.1 is provided in the Appendices A, B, and C.

The numerical implementation of Proposition 2.1 and its corollary is easy using Genz’s algorithm for the computation of trivariate normal cumulative distribution functions (Genz [10]).

In the next Proposition, we introduce a third correlated geometric Brownian motion that will serve as the endpoint of the joint distribution, and we show that this can still be analytically valued. Let {𝑆3(𝑡),𝑡0} be a geometric Brownian motion driven by the following dynamics under the initial probability measure :𝑑𝑆3(𝑡)=𝛼3𝑆3(𝑡)𝑑𝑡+𝜎3𝑆3(𝑡)𝑑𝐵3(𝑡),(2.15) where 𝛼3 and 𝜎3 are real constants (𝜎3>0), 𝐵3(𝑡) is a real-valued standard Brownian motion and correlations are as follows:𝑑𝐵1,𝐵3(𝑡)=𝜌13𝐵𝑑𝑡,𝑑2,𝐵3(𝑡)=𝜌23𝑑𝑡.(2.16) Let 𝐾4 be a positive real number and 𝑃[](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3) defined as one of the four following joint cumulative distribution functions, where 𝑡3𝑡2: (1)𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4,(2.17)(2)𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4,(2.18)(3)𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4,(2.19)(4)𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4.(2.20)

Let us introduce the following new notations:𝑘4𝐾=ln4𝑆3𝑡0,𝜇3=𝛼3𝜎232.(2.21) The next proposition provides an exact formula for the four above-mentioned joint cumulative distribution functions.

Proposition 2.2. Let the real function Φ[𝑏1,𝑏2,𝑏3,𝑏4;𝜃12,𝜃13,𝜃14,𝜃23,𝜃34], where {𝑏1,𝑏2,𝑏3,𝑏4}4 and each of the real numbers 𝜃12,𝜃13,𝜃14,𝜃23,𝜃34 is included in ]1,1[, be defined as follows: Φ𝑏1,𝑏2,𝑏3,𝑏4;𝜃12,𝜃13,𝜃14,𝜃23,𝜃34=𝑏1𝑥1=𝑏2𝑥2=𝑏3𝑥3=1𝜙21𝜙328𝜋3𝑥exp222𝑥1𝜃12𝑥222𝜙221𝑥3𝜃23𝑥222𝜙232𝑏×𝑁4𝜃14𝑥1𝜃341𝑥3𝜃13𝑥1/1𝜃213𝜙413𝑑𝑥3𝑑𝑥2𝑑𝑥1,(2.22) where 𝑁[] is the standard normal cumulative distribution function and the following definitions apply: 𝜙21=1𝜃212,𝜙32=1𝜃223,𝜃341=𝜃34𝜃13𝜃14𝜙31,𝜙413=1𝜃214𝜃2341.(2.23) Then, for the up-and-up and the down-and-down distributions, one has P[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3𝜆𝑘=Φ1𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2𝑘,𝜆4𝜇3𝑡3𝜎3𝑡3;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23exp2𝜇22𝜎22𝜆𝑘×Φ1𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑡1𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝑘,𝜆322𝜇2𝑡2𝜎2𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3𝜃34𝜃13𝜃141𝜃21322𝜎2𝑡2+𝜃13𝜌122𝜇2𝑡1𝜎2𝑡1+𝜃14𝜌122𝜇2𝑡1𝜎2𝑡1;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23exp2𝜇11𝜎21𝜆𝑘×Φ121𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2𝜌1221𝜎1𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3𝜃1421𝜎1𝑡1𝜃34𝜃13𝜃141𝜃213𝜌1221𝜎1𝑡2𝜃1321𝜎1𝑡1;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23+exp2𝜇1𝜎214𝜇2𝜌12𝜎1𝜎21+2𝜇22𝜎22𝜆𝑘×Φ121𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑡1𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1,𝜆𝑘322𝜇2𝑡2𝜎2𝑡2+𝜌1221𝜎1𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3+𝜃14𝜌122𝜇2𝑡1𝜎221𝜎1𝑡1𝜃34𝜃13𝜃141𝜃21322𝜎2𝑡2𝜌1221𝜎1𝑡2+𝜃13𝜌122𝜇2𝑡1𝜎221𝜎1𝑡1;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23,(2.24) where 𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3=𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3,𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3=𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3.(2.25) And, for the up-and-down and the down-and-up distributions, one has 𝑃[.]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3𝜆=Φ𝑘1+𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2𝑘,𝜆4𝜇3𝑡3𝜎3𝑡3;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23exp2𝜇22𝜎22𝜆×Φ𝑘1+𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑡1𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝑘,𝜆322𝜇2𝑡2𝜎2𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3𝜃34𝜃13𝜃141𝜃21322𝜎2𝑡2+𝜃13𝜌122𝜇2𝑡1𝜎2+𝜃14𝜌122𝜇2𝑡1𝜎2;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23exp2𝜇11𝜎21𝜆×Φ𝑘1+21+𝜇1𝑡1𝜎1𝑡1𝑘,𝜆2𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1𝑘,𝜆3𝜇2𝑡2𝜎2𝑡2𝜌1221𝜎1𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3𝜃1421𝜎1𝑡1𝜃34𝜃13𝜃141𝜃213𝜌1221𝜎1𝑡2𝜃1321𝜎1𝑡1;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23+exp2𝜇1𝜎214𝜇2𝜌12𝜎1𝜎21+2𝜇22𝜎22𝜆×Φ𝑘1+21+𝜇1𝑡1𝜎1𝑡1+𝜌122𝜇2𝑡1𝜎2𝑘,𝜆2+𝜇2𝑡1𝜎2𝑡1𝜌1221𝜎1𝑡1,𝜆𝑘322𝜇2𝑡2𝜎2𝑡2+𝜌1221𝜎1𝑡2,𝜆𝑘4𝜇3𝑡3𝜎3𝑡3+𝜃14𝜌122𝜇2𝑡1𝜎221𝜎1𝑡1𝜃34𝜃13𝜃141𝜃21322𝜎2𝑡2𝜌1221𝜎1𝑡2+𝜃13𝜌122𝜇2𝑡1𝜎221𝜎1𝑡1;𝜃12=𝜌12,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃23=𝑡1𝑡2,𝜃34=𝑡2𝑡3𝜌23,(2.26) where 𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3=𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3,𝜆=1if𝑃[]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3=𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3.(2.27)

Corollary of Proposition 2.2
The four following cumulative distribution functions, that will be useful in Section 3, are deduced from Proposition 2.2:(1)sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4=𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑃[up-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,(2.28)(2)inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4=𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑃[down-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,(2.29)(3)inf0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4=𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑃[down-and-up]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,(2.30)(4)sup0𝑡𝑡1𝑆1(𝑡)𝐻1,𝑆1𝑡1𝐾1,𝑆2𝑡1𝐾2,inf𝑡1𝑡𝑡2𝑆2(𝑡)𝐻2,𝑆2𝑡2𝐾3,𝑆3𝑡3𝐾4=𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑃[up-and-down]𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4.(2.31)

Proof of Proposition 2.2 is provided in the Appendices A, B, and C.

Practical use of Proposition 2.2 and its corollary in engineering applications depend on the accuracy and the efficiency with which the function Φ can be numerically implemented. This question is dealt with in Section 3.

3. Applications and Numerical Implementation

This Section deals with applications of the results provided in Section 1 in financial engineering. Indeed, Propositions 2.1 and 2.2 can be used as the building blocks for the valuation and the risk management of a large number of option contracts. The main class of instruments under consideration will be barrier options. In their standard form, the latter are contracts whereby the holder is entitled (but not obligated) to buy (call option) or sell (put option) an asset at a prespecified future date (the option expiry) at a prespecified price (the strike price), on condition that the asset price has not (knock-out type) or has (knock-in type) crossed a specific upward level (called the up-barrier) or downward level (called the down-barrier) at any given time from valuation date to expiry. In standard option pricing theory, asset prices are modelled as geometric Brownian motions, so that boundary crossing distributions such as those provided in Section 1 are of immediate use in financial engineering.

It must be emphasized that the market for barrier options is huge. They are the most actively traded class of nonstandard options (usually referred to as exotic options). Many variations on the standard barrier option payoff have been designed to match investors’ demand more closely. One of them is the partial-time barrier option, which specifies that the barrier level is monitored during only a fraction of the option lifespan. The basic contracts were studied by Heynen and Kat [11], and more general contracts were valued by Armstrong [12] and Guillaume [13]. Another variation is the step barrier option (Guillaume [6, 9]), whereby the barrier evolves as a step function of time. It is also common to encounter outside barrier options, whose payoff is a function of two asset prices: one of them is compared with the strike price at expiry to determine the moneyness of the option, while the other one is monitored up until expiry to check whether the barrier level has been crossed. Outside barrier options were originally valued by Heynen and Kat [14], and they were further studied by Kwok et al. [15] and Wong and Kwok [16].

These different features: partial time barrier, step barrier, and outside barrier, often combine to allow for increased flexibility. But then, performing analytical valuation becomes more and more involved and practitioners have to turn to numerical methods that are slow and relatively inaccurate. This is when the results provided in Section 1 become valuable. Indeed, they enable to price a large number of barrier option contracts sharing the three above-mentioned innovative features (partial time, step, and outside barrier), that is, options based on two or even three assets that may knock out depending on whether the underlying assets move within a sequence of prespecified ranges of prices over all or part of the option lifespan. More precisely, Proposition 2.1 and its corollary are the building blocks for the valuation of sequential two-asset knock-out calls and puts whose four pay-off structures are defined as follows:(1)(𝐾𝑆𝑡(2)2)+𝕀{sup10𝑡𝑡𝑆𝑡(1)<𝐻1,sup𝑡12𝑡𝑡𝑆𝑡(2)<𝐻2} for a two-asset up-and-up knock-out put (substitute (𝑆𝑡(2)2𝐾)+ for (𝐾𝑆𝑡(2)2)+ for an up-and-up knock-out call),(2)±(𝑆𝑡(2)2𝐾)+𝕀{inf10𝑡𝑡𝑆𝑡(1)>𝐻1,inf𝑡12𝑡𝑡𝑆𝑡(2)>𝐻2} for a two-asset down-and-down knock-out call or put, (3)±(𝑆𝑡(2)2𝐾)+𝕀{sup10𝑡𝑡𝑆𝑡(1)<𝐻1,inf𝑡12𝑡𝑡𝑆𝑡(2)>𝐻2} for a two-asset up-and-down knock-out call or put,(4)±(𝐾𝑆𝑡(2)2)+𝕀{inf10𝑡𝑡𝑆𝑡(1)>𝐻1,sup𝑡12𝑡𝑡𝑆𝑡(2)<𝐻2} for a two-asset down-and-up knock-out put or call,

where:(i)𝕀{} is the indicator function, (ii){𝑆1(𝑡),𝑡0} and {𝑆2(𝑡),𝑡0} are the price processes of two risky assets,(iii)𝐾 is the strike price and 𝑡2 is the option expiry, (iv)𝐻1 is a knock-out barrier monitored with respect to {𝑆1(𝑡),𝑡0} over the time interval [𝑡0,𝑡1],(v)𝐻2 is a knock-out barrier monitored with respect to {𝑆2(𝑡),𝑡0} over the time interval [𝑡1,𝑡2].

The no-arbitrage price of these eight different option contracts is now provided by the following proposition.

Proposition 3.1. Let two geometric Brownian motions {𝑆1(𝑡),𝑡0} and {𝑆2(𝑡),𝑡0} be the price processes of two risky assets with constant volatilities 𝜎1 and 𝜎2, respectively, and constant correlation coefficient 𝜌12. Let 𝛿1 and 𝛿2 be two constant pay-out rates on assets 𝑆1 and 𝑆2, respectively, and 𝑟 be the constant risk-free interest rate. Let 𝜐1=𝑟𝛿1+𝜎1𝜎2𝜌12 and 𝜐2=𝑟𝛿2+𝜎22.
The value, at time 𝑡0=0, of a two-asset sequential knock-out option with strike price 𝐾, with knock-out barriers 𝐻1 and 𝐻2 monitored with respect to 𝑆1 over the time interval [𝑡0,𝑡1] and 𝑆2 over the time interval [𝑡1,𝑡2], respectively, and with maturity 𝑡2, 𝑡2𝑡1𝑡0, is given by 𝑉𝑆1𝑡0,𝑆2𝑡0,𝑡1,𝑡2,𝐾,𝐻1,𝐻2=±𝐾exp𝑟𝑡2𝑃1𝑆2𝑡0exp𝛿2𝑡2𝑃2,(3.1) where the following specifications hold. (1) If the option is an up-and-up put, ±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1 and 𝛼2=𝑟𝛿2 in 𝑃[up-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1, and 𝑃2 is obtained by substituting 𝛼1=𝜐1 and 𝛼2=𝜐2 in 𝑃[up-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1. (2) If the option is a down-and-down call,±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1 and 𝛼2=𝑟𝛿2 in 𝑃[down-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1, and 𝑃2 is obtained by substituting 𝛼1=𝜐1 and 𝛼2=𝜐2 in 𝑃[down-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1. (3) If the option is an up-and-down call,±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1 and 𝛼2=𝑟𝛿2 in 𝑃[up-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1 and 𝑃2 is obtained by substituting 𝛼1=𝜐1 and 𝛼2=𝜐2 in 𝑃[up-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1. (4) If the option is a down-and-up put, ±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1 and 𝛼2=𝑟𝛿2 in 𝑃[down-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1 and 𝑃2 is obtained by substituting 𝛼1=𝜐1 and 𝛼2=𝜐2 in 𝑃[down-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑡1,𝑡2) in Proposition 2.1. By using Corollary of Proposition 2.1, one can value up-and-up and down-and-up call options, as well as down-and-down and up-and-down put options.

A sketch of proof of Proposition 3.1 is provided in the Appendices A, B, and C.

Just as Proposition 2.1 enables to value two-asset knock-out calls and puts, Proposition 2.2 can be used to value three-asset knock-out calls and puts. More specifically, if {𝑆3(𝑡),𝑡0} is the price process of a third risky asset and 𝑡3 is the new option expiry, with 𝑡3𝑡2𝑡1𝑡0, then the four following pay-off structures can be analytically tackled:(1)±(𝐾𝑆𝑡(3)3)+𝕀{sup10𝑡𝑡𝑆𝑡(1)<𝐻1,sup𝑡12𝑡𝑡𝑆𝑡(2)<𝐻2} for a three-asset up-and-up knock-out put or call,(2)±(𝑆𝑡(3)3𝐾)+𝕀{inf10𝑡𝑡𝑆𝑡(1)>𝐻1,inf𝑡12𝑡𝑡𝑆𝑡(2)>𝐻2} for a three-asset down-and-down knock-out call or put, (3)±(𝑆𝑡(3)3𝐾)+𝕀{sup10𝑡𝑡𝑆𝑡(1)<𝐻1,inf𝑡12𝑡𝑡𝑆𝑡(2)>𝐻2} for a three-asset up-and-down knock-out call or put,(4)±(𝐾𝑆𝑡(3)3)+𝕀{inf10𝑡𝑡𝑆𝑡(1)>𝐻1,sup𝑡12𝑡𝑡𝑆𝑡(2)<𝐻2} for a three-asset down-and-up knock-out put or call.

The no-arbitrage prices of the eight types of option under consideration are now provided by the following proposition.

Proposition 3.2. Let three geometric Brownian motions {𝑆1(𝑡),𝑡0}, {𝑆2(𝑡),𝑡0}, and {𝑆3(𝑡),𝑡0} be the price processes of three risky assets with constant volatilities 𝜎1, 𝜎2, and 𝜎3, respectively, and constant pairwise correlation coefficients 𝜌𝑎𝑏, (𝑎,𝑏){1,2,3}, at any fixed time 𝑡0. Let 𝛿1, 𝛿2, and 𝛿3 be three constant pay-out rates on assets 𝑆1, 𝑆2, and 𝑆3, respectively, and 𝑟 be the constant risk-free interest rate. Let 𝜐1=𝑟𝛿1+𝜎1𝜎3𝜌13, 𝜐2=𝑟𝛿2+𝜎2𝜎3𝜌23, and 𝜐3=𝑟𝛿3+𝜎23.
The value, at time 𝑡0=0, of a three-asset sequential knock-out option with strike price 𝐾, with knock-out barriers 𝐻1 and 𝐻2 monitored with respect to 𝑆1 over the time interval [𝑡0,𝑡1] and 𝑆2 over the time interval [𝑡1,𝑡2], respectively, with maturity 𝑡3, 𝑡3𝑡2𝑡1𝑡0, is given by 𝑉𝑆1𝑡0,𝑆2𝑡0,𝑆3𝑡0,𝑡1,𝑡2,𝑡3,𝐾,𝐻1,𝐻2=±𝐾exp𝑟𝑡3𝑃1𝑆3𝑡0exp𝛿3𝑡3𝑃2,(3.2) where the following specifications hold. (1) If the option is an up-and-up put, ±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1, 𝛼2=𝑟𝛿2 and 𝛼3=𝑟𝛿3 in 𝑃[up-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2, and 𝑃2 is obtained by substituting 𝛼1=𝜐1, 𝛼2=𝜐2 and 𝛼3=𝜐3 in 𝑃[up-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2. (2) If the option is a down-and-down call,±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1, 𝛼2=𝑟𝛿2 and 𝛼3=𝑟𝛿3 in 𝑃[down-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2, and 𝑃2 is obtained by substituting 𝛼1=𝜐1, 𝛼2=𝜐2 and 𝛼3=𝜐3 in 𝑃[down-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2. (3) If the option is an up-and-down call, ±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1, 𝛼2=𝑟𝛿2 and 𝛼3=𝑟𝛿3 in 𝑃[up-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2 and 𝑃2 is obtained by substituting 𝛼1=𝜐1, 𝛼2=𝜐2 and 𝛼3=𝜐3 in 𝑃[up-and-down](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2. (4) If the option is a down-and-up put,±=1, and 𝑃1 is obtained by substituting 𝛼1=𝑟𝛿1, 𝛼2=𝑟𝛿2 and 𝛼3=𝑟𝛿3 in 𝑃[down-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2 and 𝑃2 is obtained by substituting 𝛼1=𝜐1, 𝛼2=𝜐2 and 𝛼3=𝜐3 in 𝑃[down-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2,𝐾,𝑡1,𝑡2,𝑡3) in Proposition 2.2.By using Corollary of Proposition 2.2, one can value up-and-up and down-and-up call options, as well as down-and-down and up-and-down put options.

A sketch of proof of Proposition 3.2 is provided in the Appendices A, B, and C.

It is easy to notice that Proposition 3.2 nests Proposition 3.1 under the following three conditions:(1)𝜐1=𝑟𝛿1+𝜎1𝜎2𝜌12 and 𝜐2=𝑟𝛿2+𝜎22,(2)take 𝑃[](𝐻1,𝐻2,𝐻1,𝐻2,𝐾,𝑈,𝑡1,𝑡2,𝑡3) with 𝑈 if you are dealing with a put and with 𝑈 if you are dealing with a call,(3)compute ±[𝐾exp(𝑟𝑡2)𝑃1𝑆2(𝑡0)exp(𝛿2𝑡2)𝑃2].

It is useful, though, to state Proposition 3.1, as it uses more standard functions that can be numerically computed with known high accuracy and thus can serve as a benchmark to test the reliability of the numerical implementation of Proposition 3.2.

It can also be noticed that Proposition 3.2 enables to value sequential two-asset knock-out options having the property that the moneyness of the option at expiry is tested with respect to asset 𝑆1 instead of asset 𝑆2 as in Proposition 3.1. This is achieved by implementing Proposition 3.2 with the following specifications: (1)𝜐1=𝑟𝛿1+𝜎21 and 𝜐2=𝑟𝛿2+𝜎1𝜎2𝜌12,(2)𝑡3𝑡2, 𝜎3𝜎1, 𝜌131 and 𝜌23𝜌12, (3)compute ±[𝐾exp(𝑟𝑡2)𝑃1𝑆1(𝑡0)exp(𝛿1𝑡1)𝑃2].

Similarly, Proposition 3.2 can be used to value sequential two-asset knock-out options with expiry 𝑡3, so that, overall, this formula alone suffices to compute 40 different types of sequential multiasset knock-out options.

Proposition 3.2 implies the numerical implementation of Proposition 2.2. The function Φ in Proposition 2.2 involves three-dimensional numerical integration. Given the fact that the integrand is very smooth, a classical Gauss-Legendre quadrature can be used, along with a simple adaptive rule increasing the number of points in case the required accuracy is not reached after one iteration, starting with an 8-point quadrature. To test the efficiency and the accuracy of this numerical implementation, four series of tests have been implemented. In the first three categories of tests, the option parameters are specified so that the option can be valued by a known analytical formula involving numerical integration of lower dimension. More specifically, tests of category 1 reduce a three-asset up-and-up knock-out put to a two-asset up-and-up knock-out put by letting 𝐾4 in 𝑃[up-and-up](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3); the benchmark is then Proposition 3.1, that involves the computation of trivariate normal cumulative distribution functions. The latter can be obtained with high precision by Genz’s algorithm (Genz [10]), as mentioned earlier.

Tests of category 2 reduce the three-asset up-and-up knock-out put to an outside knock-out put by letting 𝐾1=𝐻1, 𝐻2, 𝐾3, 𝐾4, and 𝑡3𝑡2𝑡1 in 𝑃[up-and-up](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3); the benchmark, then, is the formula for an outside barrier option as given by Heynen and Kat [14], that involves the computation of bivariate normal cumulative distribution functions. Again, the latter can be obtained with high precision using Genz [10].

Tests of category 3 reduce the three-asset up-and-up knock-out put to a single asset up-and-out put by letting 𝐻2, 𝐾2=𝐾1, 𝐾3, 𝐾4, 𝑡3𝑡2𝑡1, 𝜎2=𝜎1 and 𝜌121 in 𝑃[up-and-up](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3); the benchmark, then, is the formula for a basic barrier option as given, for example, by Haug [17], that involves the computation of univariate normal cumulative distribution functions.

In the last set of tests, the three-asset up-and-up knock-out put cannot be valued as a simpler contract, so that no reduction of the dimension of numerical integration is possible. Our goal, then, is to find a robust numerical valuation method and study its convergence pattern with the results obtained by implementing Proposition 3.2. Neither using trees nor numerically solving a partial differential equation stands as the most suitable methods in this case, in view of the double difficulty of having three correlated diffusion processes and path-dependent features in the option payoff. On another hand, crude Monte Carlo simulation would be very poor both in terms of accuracy and efficiency, due to the required time discretization. That is why the most suitable numerical method in this case is probably a Brownian Bridge Monte Carlo approximation (Glasserman [18]), as it involves only 6 random draws of the asset prices for each simulation and avoids a discretization bias. This is indeed the method commonly used by market practitioners when they are faced with a pricing problem involving both multiasset and path-dependent features.

For each category of tests, a total of 10,000 different option prices have been computed with randomly drawn volatility, correlation, and expiry parameters. Let us now sum up the main findings of these numerical experiments. First, the average computational time when applying Proposition 3.2 on an ordinary personal computer (P7350 2 GHz) is 0.6 seconds, which is fast. The longest recorded computational time was 1.4 second. Speed could obviously be greatly increased by involving more hardware resources. In absolute terms, Proposition 3.2 can thus be considered as an efficient way of computing option prices. In relative terms, Proposition 3.2 is even extremely efficient. Indeed, 10,000,000 simulations are required to obtain a modest 104 digits convergence of the Brownian Bridge Monte Carlo approximation with Proposition 3.2, which amounts to a computational time of over 4 minutes! Such slowness becomes a serious issue in real-time trading or when large portfolios of options are valued. As far as accuracy is concerned, the results obtained using Proposition 3.2 always matched to at least 108 digits the analytical benchmarks in almost all the numerical experiments conducted. The very rare exceptions were observed when the absolute value of one of the correlation coefficients was above 99.7%. It must be noticed that in this couple of cases, the correlation matrix was found to be almost singular. Besides, risky assets with such levels of correlation are never encountered in the markets, so that it does not seem worthwhile to us to search for more complicated integration rules that would be better suited to these negligible subregions of integration. For all practical purposes, the numerical accuracy of our simple implementation of Proposition 3.2 can thus be deemed as very satisfactory. The results did not vary significantly between tests of categories 1, 2, and 3, which means that an increase in the number of option parameters that are given limit values does not entail instability in the quadrature rule.

To close this section, it is worth mentioning that the results provided in Section 1 can also apply to other popular derivatives contracts. Among them is the look-back option, which allows investors to sell at the highest and buy at the lowest over a time interval, thus optimizing market exit and market entry on a given asset. In its basic form, a fixed strike look-back call pays off (sup0𝑡𝑇𝑆(𝑡)𝐾)+ at expiry 𝑇. Many variations on this standard structure have been designed. One of them is the double look-back call, written on two assets 𝑆1 and 𝑆2, which provides investors with the following payoff at expiry 𝑇: (𝑎1sup0𝑡𝑇𝑆1(𝑡)𝑎2sup0𝑡𝑇𝑆2(𝑡)𝐾)+, where 𝑎1 and 𝑎2 are two positive weights. The valuation of this contract has been studied by He et al. [8]. A sequential double look-back call is a variation on the double look-back call. Its payoff at expiry 𝑡2 writes (𝑎1sup0𝑡𝑡1𝑆1(𝑡)𝑎2sup𝑡1𝑡𝑡2𝑆2(𝑡)𝐾)+. Drawing on Proposition 2.1, the no-arbitrage price of this option is equal toexp𝑟𝑡21=02=𝑎1𝑆1𝑡0exp1𝑎2𝑆2𝑡0exp2𝐾+𝜕2𝑃[up-and-up]𝐻1,𝐻2,𝐻1,𝐻2,𝐻2𝜕1𝜕2𝑑2𝑑1,(3.3) where 𝛼1=𝑟𝛿1 and 𝛼2=𝑟𝛿2 in Proposition 2.1.

The partial derivative in (3.3) can be determined by analytical differentiation of 𝑃[up-and-up](𝐻1,𝐻2,𝐻1,𝐻2,𝐻2), but this is relatively cumbersome. Given the simplicity of the computation of Proposition 2.1, it is faster and easier to compute those partial derivatives numerically. One can notice that the valuation of sequential double look-back options is only semianalytical since it still requires numerical double integration. For a given degree of the quadrature rule, it will not be as accurate as the valuation of two-asset and three-asset sequential knock-out options because of the inaccuracies arising from the numerical differentiation of Proposition 2.1.

4. Conclusion

This paper provides new explicit solutions for some sequential boundary crossing problems in a multidimensional geometric Brownian motion framework. Among their various possible applications, they form the basis for accurate and fast valuation of a large number of option contracts traded in the markets. The methods presented here could be used to solve higher-dimensional problems, especially the introduction of a third boundary monitored with respect to the third process in the model. However, the resulting formulae would become quite cumbersome and, more importantly, the quality of the numerical integration involved should have to be compared with a Brownian Bridge Monte Carlo approximation. Another possible extension would be to study the analytical tractability of introducing curved boundaries, but it should be pointed out that this form of boundary is rarely encountered in the options markets.

Appendices

A. Proof of Proposition 2.1

Proof is outlined for 𝑃[up-and-up](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝑡1,𝑡2) only. The other distributions are obtained in a similar manner.

The first step is to take the logarithms of 𝑆1(𝑡1), 𝑆2(𝑡1), and 𝑆2(𝑡2), in order to move from lognormal random variables to normal random variables. The sought cumulative distribution becomessup0𝑡𝑡1𝑋1(𝑡)1,𝑋1𝑡1𝑘1,𝑋2𝑡1𝑘2,sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑘3,(A.1) where𝑋1𝑆(𝑡)=ln1(𝑡)𝑆1𝑡0,𝑋2𝑆(𝑡)=ln2(𝑡)𝑆2𝑡0.(A.2) For a fixed 𝑡0, 𝑋1(𝑡) and 𝑋2(𝑡) are absolutely continuous random variables, hence one can express (A.1) as the following triple integral:𝑘1𝑥1=𝑘2𝑥2=𝑘3𝑥3=sup0𝑡𝑡1𝑋1(𝑡)1,𝑋1𝑡1𝑑𝑥1,𝑋2𝑡1𝑑𝑥2,sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑑𝑥3𝑑𝑥3𝑑𝑥2𝑑𝑥1.(A.3) By conditioning and using the Markov property of processes 𝑋1(𝑡) and 𝑋2(𝑡), (A.3) becomes𝑘1𝑥1=𝑘2𝑥2=𝑘3𝑥3=𝑃1𝑥1𝑃2𝑥1,𝑥2𝑃3𝑥2,𝑥3𝑑𝑥3𝑑𝑥2𝑑𝑥1,(A.4) where𝑃1𝑥1=sup0𝑡𝑡1𝑋1(𝑡)1,𝑋1𝑡1𝑑𝑥1,𝑃2𝑥1,𝑥2𝑋=2𝑡1𝑑𝑥2𝑋1𝑡1𝑑𝑥1,𝑃3𝑥2,𝑥3=sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑑𝑥3𝑋2𝑡1𝑑𝑥2.(A.5) The functions 𝑃1(𝑥1) and 𝑃2(𝑥1,𝑥2) are classical results. To obtain the function 𝑃3(𝑥2,𝑥3), the following formula is used (Guillaume [6]):𝑋2𝑡1𝑥2,sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑥3=𝑁2𝑥2𝜇2𝑡1𝜎2𝑡1,𝑥3𝜇2𝑡2𝜎2𝑡2;𝑡1𝑡2exp2𝜇2𝜎222𝑁2𝑥2+𝜇2𝑡1𝜎2𝑡1,𝑥322𝜇2𝑡2𝜎2𝑡2;𝑡1𝑡2,(A.6) where 𝑁2[,;𝜌] is the bivariate standard normal cumulative distribution function with correlation coefficient 𝜌. Differentiating (A.6) twice with respect to variables 𝑥2 and 𝑥3, and applying the definition of conditional probability yields𝑃3𝑥2,𝑥3=𝑥exp(1/2)3𝑥2𝜇2𝑡2𝑡1/𝜎2𝑡2𝑡12𝜎2𝑡2𝜋2𝑡1exp2𝜇2/𝜎222𝑥2𝑥(1/2)3+𝑥222𝜇2𝑡2𝑡1/𝜎2𝑡2𝑡12𝜎2𝑡2𝜋2𝑡1.(A.7) Then, the most cumbersome part of the proof consists in performing the necessary calculations to solve (A.4) in closed form and find the sum of trivariate standard normal cumulative distribution functions given by Proposition 2.1.

B. Proof of Proposition 2.2

Proof is outlined for 𝑃[up-and-up](𝐻1,𝐻2,𝐾1,𝐾2,𝐾3,𝐾4,𝑡1,𝑡2,𝑡3) only. The other distributions are obtained in a similar manner.

The following lemma is instrumental in proving Proposition 2.2.

Lemma B.1. Let 𝑋1, 𝑋2, and 𝑋3 be three standard normal random variables defined on a probability space with measure . Denote by 𝜃𝑎𝑏 the correlation coefficient between 𝑋𝑎 and 𝑋𝑏, (𝑎,𝑏){1,2,3}. Then, the joint density function of (𝑋1,𝑋2,𝑋3) is given by 𝑋1𝑑𝑥1,𝑋2𝑑𝑥2,𝑋3𝑑𝑥3𝑥=exp21212𝜙221𝑥2𝜃12𝑥1212𝜙2312𝑥3𝜃13𝑥1𝜃231𝑥2𝜃12𝑥1𝜙212×(2𝜋)3/2𝜙21𝜙3121,(B.1) where 𝜙2|1=1𝜃212 is the standard deviation of 𝑋2 conditional on 𝑋1, 𝜃23|1=(𝜃23𝜃12𝜃13)/𝜙2|1 is the partial correlation between 𝑋2 and 𝑋3 conditional on 𝑋1, and 𝜙3|12=1𝜃213𝜃223|1 is the standard deviation of 𝑋3 conditional on 𝑋1 and on 𝑋2.

Proof of Lemma B.1 is simple. It is known that 𝑋2 is equal in law to 𝜃12𝑋1+𝜙2|1𝑌2, where 𝑌2 is a standard normal random variable uncorrelated with 𝑋1. By applying Cholesky’s decomposition to the correlation matrix associated with (𝑋1,𝑋2,𝑋3), one also finds that 𝑋3 admits the following representation:𝑋3=𝜃13𝑋1+𝜃231𝑌2+𝜙312𝑌3𝑋3=𝜃13𝑋1+𝜃231𝑋2𝜃12𝑋1𝜙21+𝜙312𝑌3,(B.2) where 𝑌3 is a standard normal random variable uncorrelated with 𝑋1 and 𝑌2.

Thus, given 𝑋1, 𝑋2, and 𝑋3 is normally distributed with mean 𝜃13𝑋1+𝜃23|1((𝑋2𝜃12𝑋1)/𝜙2|1) and standard deviation 𝜙3|12. Hence, the density function of 𝑋3 conditional on 𝑋1 and 𝑋2 writes𝑋3𝑑𝑥3𝑋1𝑑𝑥1,𝑋2𝑑𝑥21=exp2𝜙2312𝑥3𝜃13𝑥1𝜃231𝑥2𝜃12𝑥1𝜙212𝜙3122𝜋1,(B.3) which suffices to prove Lemma B.1.

Next, we proceed to outline proof of Proposition 2.2. By taking the logarithms of the asset prices at the times 𝑡1, 𝑡2, 𝑡3, and integrating with respect to these variables, the sought cumulative distribution function becomes 𝑘1𝑥1=𝑘2𝑥2=𝑘3𝑥3=𝑘4𝑥4=sup0𝑡𝑡1𝑋1(𝑡)1,𝑋1𝑡1𝑑𝑥1,𝑋2𝑡1𝑑𝑥2,sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑑𝑥3,𝑋3𝑡3𝑑𝑥4𝑑𝑥4𝑑𝑥3𝑑𝑥2𝑑𝑥1,(B.4) where 𝑋3(𝑡)=ln(𝑆3(𝑡)/𝑆3(𝑡0)) and the other terms are defined in the proof of Proposition 2.1. Then, by conditioning and using the Markov property of the processes 𝑋1(𝑡), 𝑋2(𝑡) and 𝑋3(𝑡), (B.4) becomes𝑘1𝑥1=𝑘2𝑥2=𝑘3𝑥3=𝑘4𝑥4=𝑃1𝑥1𝑃2𝑥1,𝑥2𝑃3𝑥2,𝑥3𝑃4𝑥1,𝑥3,𝑥4𝑑𝑥4𝑑𝑥3𝑑𝑥2𝑑𝑥1,(B.5) where𝑃1𝑥1=sup0𝑡𝑡1𝑋1(𝑡)1,𝑋1𝑡1𝑑𝑥1,𝑃2𝑥1,𝑥2𝑋=2𝑡1𝑑𝑥2𝑋1𝑡1𝑑𝑥1,𝑃3𝑥2,𝑥3=sup𝑡1𝑡𝑡2𝑋2(𝑡)2,𝑋2𝑡2𝑑𝑥3𝑋2𝑡1𝑑𝑥2,𝑃4𝑥1,𝑥3,𝑥4𝑋=3𝑡3𝑑𝑥4𝑋2𝑡2𝑑𝑥3,𝑋1𝑡1𝑑𝑥1.(B.6) The functions 𝑃1(𝑥1), 𝑃2(𝑥1,𝑥2), and 𝑃3(𝑥2,𝑥3) are dealt with in the proof of Proposition 2.1. The function 𝑃4(𝑥1,𝑥3,𝑥4) is now derived. Since the processes {𝑆1(𝑡),𝑡0}, {𝑆2(𝑡),𝑡0} and {𝑆3(𝑡),𝑡0} are geometric Brownian motions with constant drift and diffusion coefficients, the random variables 𝑋1(𝑡1), 𝑋2(𝑡2), and 𝑋3(𝑡3) are normally distributed. Elementary calculations yield their cross-correlations. Applying Cholesky’s decomposition to 𝑋2(𝑡2) and 𝑋3(𝑡3), one can see that any linear combination 𝑎1𝑋1(𝑡1)+𝑎2𝑋2(𝑡2)+𝑎3𝑋3(𝑡3) for any set of real numbers {𝑎1,𝑎2,𝑎3} is a sum of three uncorrelated normal random variables; hence, it is normally distributed. Therefore, using Fréchet’s characterization of multivariate normality (Fréchet [19]), the joint density function 𝑃4(𝑥1,𝑥3,𝑥4) is trivariate normal, and Lemma B.1 yields 𝑃4𝑥1,𝑥3,𝑥41=exp2𝜙2413𝜑3𝑥4𝜃14𝜑1𝑥1𝜃341𝜑2𝑥3𝜃13𝜑1𝑥1𝜙312×𝜙413𝜙31𝜎1𝜎2𝜎3𝑡1𝑡2𝑡38𝜋31,(B.7) where the notations are defined as follows:𝜑1𝑥1=𝑥1𝜇1𝑡1𝜎1𝑡1,𝜑2𝑥3=𝑥3𝜇2𝑡2𝜎2𝑡2,𝜑3𝑥4=𝑥4𝜇3𝑡3𝜎3𝑡3,𝜃13=𝑡1𝑡2𝜌12,𝜃14=𝑡1𝑡3𝜌13,𝜃34=𝑡2𝑡3𝜌23,𝜙31=1𝜃213,𝜃341=𝜃34𝜃13𝜃14𝜙31,𝜙413=1𝜃214𝜃2341.(B.8) Then, the most cumbersome part of the proof consists in performing the necessary calculations to solve (B.5) in closed form and find the sum of functions given by Proposition 2.2.

C. Sketch of Proof of Propositions 3.1 and 3.2

Once endowed with Proposition 2.1, proof of Proposition 3.1 is straightforward, as an application of option pricing theory (Harrison and Pliska [20]) and Girsanov’s theorem in two dimensions. The substitutions 𝛼1=𝜐1 and 𝛼2=𝜐2 in Proposition 2.1 derive from the introduction of a new measure, which is equivalent to the risk-neutral measure, under which the market numeraire is asset 𝑆2. If we denote by {𝐵1(𝑡),𝑡0} and {𝐵2(𝑡),𝑡0} the two correlated Brownian motions driving {𝑆1(𝑡),𝑡0} and {𝑆2(𝑡),𝑡0} under the risk-neutral measure, then the processes {𝐵1(𝑡)𝜎2𝜌12𝑡,𝑡0} and {𝐵2(𝑡)𝜎21𝜌212𝑡,𝑡0} are independent standard Brownian motions under the new measure.

Similarly, once endowed with Proposition 2.2, proof of Proposition 3.2 is straightforward, as an application of option pricing theory and Girsanov’s theorem in three dimensions. The substitutions 𝛼1=𝜐1, 𝛼2=𝜐2, and 𝛼3=𝜐3 in Proposition 2.2 derive from the introduction of a new measure, which is equivalent to the risk-neutral measure, under which the market numeraire is asset 𝑆3. If we denote by {𝐵1(𝑡),𝑡0}, {𝐵2(𝑡),𝑡0}, and {𝐵3(𝑡),𝑡0} the three correlated Brownian motions driving {𝑆1(𝑡),𝑡0}, {𝑆2(𝑡),𝑡0}, and {𝑆3(𝑡),𝑡0} under the risk-neutral measure, then the three following processes:𝐵1(𝑡)𝜎3𝜌13,𝐵𝑡,𝑡02(𝑡)𝜎3𝜌23𝜌12𝜌131𝜌212,𝐵𝑡,𝑡03(𝑡)𝜎3𝑡1𝜌213𝜌23𝜌12𝜌1321𝜌212,𝑡0(C.1) are independent standard Brownian motions under the new measure.