• Keine Ergebnisse gefunden

Alpha-rootProcessesforDerivativespricing Balakrishna,BS MunichPersonalRePEcArchive

N/A
N/A
Protected

Academic year: 2022

Aktie "Alpha-rootProcessesforDerivativespricing Balakrishna,BS MunichPersonalRePEcArchive"

Copied!
15
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Munich Personal RePEc Archive

Alpha-root Processes for Derivatives pricing

Balakrishna, B S

11 January 2010

Online at https://mpra.ub.uni-muenchen.de/21396/

MPRA Paper No. 21396, posted 16 Mar 2010 01:19 UTC

(2)

Alpha-Root Processes for Derivatives Pricing

B. S. BALAKRISHNA

βˆ—

January 11, 2010, Revised: February 16, 2010.

Abstract

A class of mean reverting positive stochastic processes driven by π›Όβˆ’stable distri- butions, referred to here as π›Όβˆ’root processes in analogy to the square root process (Cox-Ingersoll-Ross process), is a subclass of affine processes, in particular continuous state branching processes with immigration (CBI processes). Being affine, they provide semi-analytical results for the implied term structures as well as for the characteristic exponents for their associated distributions. Their use has not been appreciated in the field perhaps due to lack of an efficient numerical algorithm to supplement their semi-analytical results. The present article introduces a convenient formulation of such processes, CBI processes in general, in the form of pure-jump processes of infinite activ- ity. The formulation admits an efficient simulation algorithm that enables an extensive investigation of their properties.

Stochastic processes are the building blocks of modeling discipline. Though Brownian motion has been largely successful in this regard, there are certain areas where more advanced processes could be helpful. This is especially so in mathematical finance wherein alternate processes have been utilized, in particular to provide an explanation to parameter smiles, such as volatility smiles or correlation smiles. Among other approaches, a class of stochastic processes called π›Όβˆ’stable LΒ΄evy processes have been used for this purpose with encouraging results. Because applicable𝛼usually lies between 1 and 2, and the associated stable processes can have negative values, their use has been largely limited to their exponentials as stochastic variables of interest. This makes them analytically intractable for many objects of interest, such as term structures of discount factors in interest rate modeling or survival probabilities in credit risk modeling.

It is known that the Cox-Ingersoll-Ross process, also known as the square-root process, though confined only to the positive real axis, admits analytical results for term structure modeling. It belongs to a class of affine processes, the spot rate in interest rate modeling being related affinely to the short rate. It is driven by Brownian motion which in the language of stable processes has 𝛼 = 2. A natural question then arises as to whether there exist π›Όβˆ’root processes driven by π›Όβˆ’stable distributions, and whether they too exhibit the affine property. As it turns out, the answer to this question is pleasantly in the affirmative. π›Όβˆ’root processes thus provide a natural and appealing approach to affine jump diffusion processes

βˆ—Email: balak bs@yahoo.co.in

(3)

by incorporating jumps into the diffusion component itself to turn it into anπ›Όβˆ’root process, rather than extending the process to include a jump component.

The class of affine processes is a well-studied branch of mathematics, and has been char- acterized in generality by Duffie, Filipovic and Schachermayer [2003]. However, this class being rather large, identification of specific affine processes for their usefulness is important in itself. Being a subclass of affine processes, in particular continuous state branching pro- cesses with immigration (CBI processes), and a natural extension of the square-root process, π›Όβˆ’root processes have caught the attention of researchers in the field. For instance, they are briefly touched upon by Carr and Wu [2004] as an activity process for generating random time. Their use has not been appreciated in the field perhaps due to lack of an efficient nu- merical algorithm to supplement their semi-analytical results. The present article introduces a convenient formulation of such processes, CBI processes in general, in the form of pure- jump processes of infinite activity. The formulation admits an efficient simulation algorithm that enables an extensive investigation of their properties. The algorithm is also adaptable to the case of standard mean reverting processes (Ornstein-Uhlenbeck-Type processes) driven byπ›Όβˆ’stable processes, or LΒ΄evy processes of infinite activity.

Section 1 introduces the π›Όβˆ’root process, CBI process in general, in the form of a mean- reverting pure-jump process of infinite activity and presents semi-analytical solutions for the implied term structures and the Laplace exponents. Section 2 presents closed form expressions for the Laplace exponents in some special cases. Section 3 presents an efficient Monte Carlo simulation algorithm that enables a numerical investigation of the process.

Section 4 discusses the simulation results. Semi-analytical solutions are derived in Appendix A. The results of a numerical investigation are presented in Figures 1-10.

1 Alpha-Root Process

Let us start with the following pure jump process for a positive stochastic variable π‘Ÿ(𝑑), π‘‘π‘Ÿ(𝑑) = [πœ™(𝑑)βˆ’π‘šπ‘Ÿ(𝑑)]𝑑𝑑+

∫ ∞

𝑧=0

β„Ž(𝑧/π‘Ÿ(𝑑))𝑑𝑀(𝑑𝑧, 𝑑). (1) Here𝑑𝑀(𝑑𝑧, 𝑑) = 𝑑𝑁(𝑑𝑧, 𝑑)βˆ’π‘‘π‘§π‘‘π‘‘where𝑁(𝑑𝑧, 𝑑)s are independent Poisson processes. Pro- cess𝑁(𝑑𝑧, 𝑑) is of intensity𝑑𝑧 and is associated with the interval (𝑧, 𝑧+𝑑𝑧). If𝑁(𝑑𝑧, 𝑑) jumps up by one at time 𝑑, 𝑑𝑁(𝑑𝑧, 𝑑) causes π‘Ÿ(𝑑) to jump up by β„Ž(𝑧/π‘Ÿ(π‘‘βˆ’)) where π‘‘βˆ’ is just prior to 𝑑. We may refer to β„Ž(π‘₯) as the jump function. It is taken to be nonnegative, integrable from π‘₯ = 0, going to zero as its argument π‘₯ β†’ ∞. 𝑀(𝑑𝑧, 𝑑) is the compensated Poisson process (a Martingale). Parameter π‘š is the mean reversion rate. Drift πœ™(𝑑) is assumed to be positive. Note that the total intensity of the Poisson processes is infinite and hence the stochastic process is of infinite activity (however, effective intensity depends on the jump function and is not necessarily infinite).

An attractive feature of the above process is that it is an affine model, just as the well- known square-root process is. Note that process (1) is written in somewhat an unconventional way. It is usual to regard jump β„Ž as an independent variable with the Poisson intensity 𝑑𝑧 = (𝑑𝑧/π‘‘β„Ž)π‘‘β„Ž giving rise to an intensity density βˆ£π‘‘π‘§/π‘‘β„Žβˆ£ called the LΒ΄evy density. Working with the jump function β„Ž(π‘₯) has provided us with a convenient formulation of an affine

(4)

process, in particular a CBI process (for constant πœ™(𝑑), that could also have square-root diffusion and π‘Ÿβˆ’independent nonnegative jump component), in the form of a stochastic differential equation that forms the basis of a simulation to be discussed later.

Being an affine model, process (1) admits semi-analytical results for the implied term structures as well as for the characteristic exponents for their associated distributions. The following result is derived in Appendix A,

E𝑑 {

exp [

βˆ’

∫ 𝑇

𝑑

𝑑𝑠𝑒(𝑇 βˆ’π‘ )π‘Ÿ(𝑠) ]}

= exp [

βˆ’

∫ 𝑇

𝑑

π‘‘π‘ πœ™(𝑠)𝐡(𝑇 βˆ’π‘ )βˆ’π΅(𝑇 βˆ’π‘‘)π‘Ÿ(𝑑) ]

, (2) where 𝑒(𝜏) is some deterministic function and 𝐡(𝜏), satisfying𝐡(0) = 0, is a solution of

𝑑𝐡(𝜏)

π‘‘πœ +π‘šπ΅(𝜏) = 𝑒(𝜏) +

∫ ∞

0

𝑑π‘₯{1βˆ’β„Ž(π‘₯)𝐡(𝜏)βˆ’exp [βˆ’β„Ž(π‘₯)𝐡(𝜏)]}. (3) Result (2) features the affine property, the expression within square brackets being related affinely toπ‘Ÿ(𝑑). For term structure modeling, one is interested in solving the above equation with 𝑒(𝜏) = 1. If interested in the Laplace transform E𝑑{exp [βˆ’π‘’π‘Ÿ(𝑇)]} of the probability density function of π‘Ÿ(𝑇), or its negative logarithm known as the Laplace exponent, the equation is solved in the absence of 𝑒(𝜏), but under the initial condition𝐡(0) =𝑒.

The above result is for a general jump function β„Ž(π‘₯). For β„Ž(π‘₯) = π‘Žπ‘₯βˆ’1/𝛼, 1 < 𝛼 < 2, we have β„Ž(𝑧/π‘Ÿ)βˆπ‘Ÿ1/𝛼 and (1) may be referred to as anπ›Όβˆ’root process. Equation for𝐡(𝜏) then becomes

𝑑𝐡(𝜏)

π‘‘πœ +π‘šπ΅(𝜏) = 𝑒(𝜏)βˆ’πœŽπ›Ό[𝐡(𝜏)]𝛼, 1< 𝛼 <2,

= 𝑒(𝜏)βˆ’πœŽπ΅(𝜏) ln𝐡(𝜏), 𝛼= 1. (4) where 𝜎 is π‘Ž[𝛼Γ(βˆ’π›Ό)]1/𝛼 for 1 < 𝛼 < 2 and is π‘Ž for 𝛼 = 1. Equation for 𝛼 = 1 is also presented above, though it needs to be treated as a special case. For the Laplace exponent, the above can be solved with 𝑒(𝜏) = 0 and 𝐡(0) =𝑒 to obtain

𝐡(𝜏) = π‘’βˆ’π‘šπœ {

π‘’βˆ’(π›Όβˆ’1)+πœŽπ›Ό π‘š

[1βˆ’π‘’βˆ’(π›Όβˆ’1)π‘šπœ]

}βˆ’1/(π›Όβˆ’1)

, 1< 𝛼 <2,

= exp[ π‘’βˆ’πœŽπœ(

ln𝑒+ π‘š 𝜎

)βˆ’π‘š 𝜎

], 𝛼 = 1. (5)

Case 𝛼 <1 turns out to be inconsistent. These results have a limit as 𝛼→2 (given a fixed 𝜎) to correspond to the case of the square-root process. Closed form expressions for the Laplace exponent can be obtained in some special cases as discussed in the next section.

Drift πœ™(𝑑) has been assumed to be positive. This ensures that the origin is inaccessible, that the probability density of π‘Ÿ(𝑇) as π‘Ÿ(𝑇) β†’ 0 goes to zero. This can be examined, as usual in Laplace transforms, by looking at the 𝑒 β†’ ∞ limit of 𝑒E𝑑{exp [βˆ’π‘’π‘Ÿ(𝑇)]}. The leading contribution comes from the integral in (2) near 𝑠=𝑇,

𝑒E𝑑{exp [βˆ’π‘’π‘Ÿ(𝑇)]} βˆΌπ‘’exp [

βˆ’πœ™(𝑇) 𝑒2βˆ’π›Ό (2βˆ’π›Ό)πœŽπ›Ό

]

, as 𝑒→ ∞. (6)

(5)

Given πœ™(𝑑) > 0, this goes to zero as 𝑒 β†’ ∞. For 𝛼 = 2, one obtains the well-known requirementπœ™(𝑑)> 𝜎2 (volatility of the square-root process is 𝜎√

2 in our scale convention).

As for𝛼= 1, 𝐡(𝜏)β†’ ∞as𝑒→ ∞ for all𝜏 so that the above quantity goes to zero for any πœ™(𝑑)β‰₯0 (in this case, πœ™(𝑑) can be zero).

The π›Όβˆ’root process can be viewed as being driven by an π›Όβˆ’stable LΒ΄evy process. This is analogous to the square root process being driven by the Brownian motion. To see this, consider small 𝜏 = 𝑇 βˆ’ 𝑑 when 𝐡(𝜏) ≃ (1βˆ’π‘šπœ)𝑒 βˆ’ πœŽπ›Όπœ 𝑒𝛼 and the Laplace exponent approximates to

[π‘Ÿ(𝑑) + (πœ™(𝑑)βˆ’π‘šπ‘Ÿ(𝑑))𝜏]π‘’βˆ’πœŽπ›Όπ‘Ÿ(𝑑)𝜏 𝑒𝛼. (7) The 𝑒𝛼 term is the Laplace exponent of a stable distribution of index𝛼 and skew parameter one (maximally skewed to the right) with zero location, the term linear in𝑒arising from the deterministic part of theπ‘Ÿβˆ’process. Its scale parameter is𝜎(π‘Ÿ(𝑑)𝜏)1/𝛼(times [βˆ’cos(πœ‹π›Ό/2)]1/𝛼 to be exact), as expected with theπ›Όβˆ’root of π‘Ÿ(𝑑) attached (similar analysis can be done for 𝛼= 1). Given the above infinitesimal result, one can indeed recover the full Laplace exponent using the law of iterated expectations. Note that infinitesimally, theπ›Όβˆ’root process can be viewed as being driven by a time-scaled stable process, 𝜏 getting effectively scaled by π‘Ÿ(𝑑).

This is a stochastic scaling of time, scaling by the stochastic process π‘Ÿ(𝑑) itself. This gives us an alternate view of process (1) for generalβ„Ž(π‘₯) as well, providing a relationship between CBI processes and LΒ΄evy processes (known as Lamperti representation).

The expression for term structure in (2) involves convolution of πœ™(𝑠) and 𝐡(𝑠) (consider 𝑑 = 0). When modeling term structure models, say for interest rates or credit spreads, one approach is to imply the drift πœ™(𝑑) from the given data on discount factors or survival probabilities as the case may be. If this deconvolving procedure is not convenient, one may consider the well-known approach in affine modeling of working with a constantπœ™, but with the stochastic variable π‘Ÿ(𝑑) related to the variable of interest by a deterministic shift that is implied from the given data (see Brigo and Alfonsi (2005) for such an approach with the square root process).

2 Laplace Exponents

The Laplace exponent of the distribution of π‘Ÿ(𝑇) can be obtained given the solution (5) for 𝐡(𝜏). For constant drift πœ™(𝑑) =πœ™ and for 1< 𝛼 <2, this gives for the exponent

πœˆπœ™ π‘šπ‘žπœˆ

∫ 1+𝑝𝑒1/𝜈

1

𝑑π‘₯π‘₯βˆ’πœˆ(

1 +π‘žπ‘’1/𝜈 βˆ’π‘₯)πœˆβˆ’1

+π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

(1 +𝑝𝑒1/𝜈)𝜈 , (8) where𝜈 = 1/(π›Όβˆ’1), π‘ž=πœŽπ›Ό/π‘š and 𝑝=π‘ž(1βˆ’π‘’βˆ’(π›Όβˆ’1)π‘š(π‘‡βˆ’π‘‘)). The integral can be expressed in terms of incomplete beta functions. For small 𝑒, the exponent has the expansion

[πœ™

π‘š(1βˆ’π‘ ) +π‘Ÿ(𝑑)𝑠 ]

π‘’βˆ’ { πœ™

π‘šπ›Ό[π‘ž(1βˆ’π‘ )βˆ’π‘πœˆπ‘ ] +π‘πœˆπ‘Ÿ(𝑑)𝑠 }

𝑒𝛼, (9)

where𝑠 =π‘’βˆ’π‘š(π‘‡βˆ’π‘‘). This gives the mean, and the scale parameter for the largeπ‘Ÿ(𝑇) behavior (nonanalytic π‘’π›Όβˆ’behavior as 𝑒→0 indicates that theπ‘Ÿ(𝑇)β†’ ∞ tail is similar to that of a

(6)

stable distribution of index 𝛼). Closed form expression for the exponent can be obtained if π‘š= 0, that reads

πœ™π‘’2βˆ’π›Ό (2βˆ’π›Ό)πœŽπ›Ό

[1βˆ’(

1 +π‘π‘’π›Όβˆ’1)βˆ’(2βˆ’π›Ό)/(π›Όβˆ’1)]

+ π‘Ÿ(𝑑)𝑒

(1 +π‘π‘’π›Όβˆ’1)1/(π›Όβˆ’1), (10) where 𝑝 = (π›Όβˆ’1)πœŽπ›Ό(𝑇 βˆ’π‘‘). If π‘š βˆ•= 0, closed form expressions can be obtained for some special values of𝛼. For the limiting case of 𝛼= 2, we obtain the well-known result

πœ™

𝜎2ln(1 +𝑝𝑒) + π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

1 +𝑝𝑒 , (11)

where 𝑝 = (𝜎2/π‘š)(1βˆ’π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)). This is the exponent of the non-central chi-square distri- bution (volatility of the square-root process is𝜎√

2). For 𝛼= 3/2, one obtains 2πœ™

π‘šπ‘ž2 {π‘βˆš

𝑒(1 +π‘žβˆš 𝑒) 1 +π‘βˆš

𝑒 βˆ’ln(

1 +π‘βˆš 𝑒)

}

+π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒 (1 +π‘βˆš

𝑒)2 , (12)

where π‘ž=𝜎3/2/π‘š and 𝑝=π‘ž(1βˆ’π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)/2). For 𝛼= 4/3, the exponent is 3πœ™

π‘šπ‘ž3

{𝑝𝑒1/3(1 +π‘žπ‘’1/3) 𝑃(𝑒)

[π‘ž 2𝑒1/3

(

1 + π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)/3 𝑃(𝑒)

)

βˆ’1 ]

+ ln (𝑃(𝑒)) }

+ π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

(𝑃(𝑒))3 , (13) where π‘ž=𝜎4/3/π‘š and 𝑝=π‘ž(1βˆ’π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)/3) and 𝑃(𝑒) = 1 +𝑝𝑒1/3. Another integrable case is 𝛼= 5/3 that gives

3πœ™ π‘šπ‘žβˆšπ‘ž

{βˆšπ‘žπ‘’1/3𝑅(𝑒)

√1 +𝑝𝑒2/3 βˆ’Sinβˆ’1

[βˆšπ‘žπ‘’1/3𝑅(𝑒) 1 +π‘žπ‘’2/3

]}

+π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

(1 +𝑝𝑒2/3)3/2. (14) Here π‘ž=𝜎5/3/π‘š, 𝑝=π‘ž(1βˆ’π‘’βˆ’2π‘š(π‘‡βˆ’π‘‘)/3) and 𝑅(𝑒) =√

1 +𝑝𝑒2/3βˆ’π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)/3. Closed form expressions can be obtained more generally for 𝛼= 1 + 2/π‘˜ whereπ‘˜ β‰₯2 is an integer.

Closed form expressions for the exponent can also be obtained for certain time-dependent drifts. For instance, consider a time-dependence of the form πœ™(𝑑) = πœ™π‘’βˆ’πœ…π‘šπ‘‘ given some constant πœ…. The exponent in integral form then reads

πœˆπœ™π‘’βˆ’πœ…π‘šπ‘‡π‘’πœ… π‘šπ‘ž(1βˆ’πœ…)𝜈

∫ 1+𝑝𝑒1/𝜈

1

𝑑π‘₯π‘₯βˆ’πœˆ(

1 +π‘žπ‘’1/πœˆβˆ’π‘₯)(1βˆ’πœ…)πœˆβˆ’1

+π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

(1 +𝑝𝑒1/𝜈)𝜈 , (15) where as before 𝜈 = 1/(𝛼 βˆ’ 1), π‘ž = πœŽπ›Ό/π‘š and 𝑝 = π‘ž(1βˆ’ π‘’βˆ’(π›Όβˆ’1)π‘š(π‘‡βˆ’π‘‘)). Closed form expression can be obtained for πœ…= 2βˆ’π›Ό,

πœ™π‘’βˆ’(2βˆ’π›Ό)π‘šπ‘‡π‘’(2βˆ’π›Ό) (2βˆ’π›Ό)πœŽπ›Ό

[1βˆ’(

1 +π‘π‘’π›Όβˆ’1)βˆ’(2βˆ’π›Ό)/(π›Όβˆ’1)]

+ π‘Ÿ(𝑑)π‘’βˆ’π‘š(π‘‡βˆ’π‘‘)𝑒

(1 +π‘π‘’π›Όβˆ’1)1/(π›Όβˆ’1). (16) Closed form expressions can also be obtained for some other choices of πœ…, for instance when (1βˆ’πœ…)𝜈 is a positive integer.

(7)

3 Monte Carlo Simulation

Process (1) is of infinite activity as presented. The integral over 𝑧 needs to be cut off at the higher end to render the total intensity of the Poisson processes finite for simulation purpose.

This can be done by forcingβ„Ž(π‘₯) = 0 for π‘₯ > 𝑋 given a sufficiently large𝑋. Process (1) can now be viewed as being driven by a compound Poisson process of stochastic total intensity π‘Ÿ(𝑑)𝑋. It can be simulated starting with a more convenient form,

𝑑[π‘Ÿ(𝑑)βˆ’π‘π‘‹(𝑑)] =βˆ’π‘šπ‘‹[π‘Ÿ(𝑑)βˆ’π‘π‘‹(𝑑)]𝑑𝑑+

∫ π‘Ÿ(𝑑)𝑋

𝑧=0

β„Ž(𝑧/π‘Ÿ(𝑑))𝑑𝑁(𝑑𝑧, 𝑑). (17) Hereπ‘šπ‘‹ =π‘š+βˆ«π‘‹

0 𝑑π‘₯β„Ž(π‘₯) and𝑐𝑋 is introduced viaπœ™(𝑑) =𝑑𝑐𝑋(𝑑)/𝑑𝑑+π‘šπ‘‹π‘π‘‹(𝑑). Since πœ™(𝑑) is taken to be positive, 𝑐𝑋(𝑑) solves to be positive. 𝑐𝑋(0) can be conveniently chosen, say as π‘Ÿ(0) or πœ™(0)/π‘šπ‘‹. The algorithm reads as follows.

1. Set π‘‘π‘œ = 0 and π‘Ÿ=π‘Ÿ(0).

2. Draw an independent exponentially distributed unit mean random number 𝑣. Set𝑑to the next event arrival time π‘‘π‘œ+𝑣/𝑍 where𝑍 =π‘Ÿπ‘‹, or the time horizon whichever is earlier.

3. Update π‘Ÿ to π‘Ÿβˆ’ given by

π‘Ÿβˆ’= (π‘Ÿβˆ’π‘π‘‹(π‘‘π‘œ))π‘’βˆ’π‘šπ‘‹(π‘‘βˆ’π‘‘π‘œ)+𝑐𝑋(𝑑). (18) 4. If 𝑑 is the time-horizon, go to step 6.

5. Draw an independent uniformly distributed random number π‘€βˆˆ[0,1]. Update π‘Ÿβˆ’ to π‘Ÿ =π‘Ÿβˆ’+β„Ž(π‘₯), whereπ‘₯=𝑀𝑍/π‘Ÿβˆ’. (19) Note thatβ„Ž(π‘₯) = 0 if π‘₯ > 𝑋. Set π‘‘π‘œ =𝑑 and go to step 2.

6. Collect this sample or value a derivative. For the next scenario, go to step 1.

7. From all the samples thus obtained, determine the distribution, or average the values to obtain a price for the derivative.

An attractive feature of the algorithm is that it does not involve discretization of time.

Some improvements are possible to ensure that𝑍 β‰₯π‘Ÿ(𝑑)𝑋 in between Poisson events if𝑐𝑋(𝑑) increases with𝑑 and can makeπ‘Ÿβˆ’ larger thanπ‘Ÿ before the next event arrival time. Note that, since jumps are nonnegative, π‘Ÿ(𝑑) never goes below 𝑐𝑋(𝑑) (consider 𝑐𝑋(0) = π‘Ÿ(0)). Hence, because 𝑐𝑋(𝑑)> π‘βˆž(𝑑) for any finite𝑋 (and 𝑑 > 0), to sample π‘Ÿ(𝑑) close to its lower bound of π‘βˆž(𝑑), 𝑋 will have to be very large. For the π›Όβˆ’root process,π‘βˆž(𝑑) is zero and there will always be some region left unsampled near zero for any finite 𝑋. This deficiency is corrected in the updated algorithm discussed below.

For β„Ž(π‘₯) = π‘Žπ‘₯βˆ’1/𝛼, 1 < 𝛼 < 2, there is an issue of convergence. The π‘₯βˆ’integral in (3), limited to π‘₯≀𝑋, can be approximated as

βˆ’π›ΌΞ“(βˆ’π›Ό)(π‘Žπ΅)𝛼+ 𝛼

2(2βˆ’π›Ό)(π‘Žπ΅)2𝑋1βˆ’2/π›Όβˆ’ 𝛼

6(3βˆ’π›Ό)(π‘Žπ΅)3𝑋1βˆ’3/𝛼+π’ͺ(

(π‘Žπ΅)4𝑋1βˆ’4/𝛼) .

Note that, as 𝛼→2, the second term tends to be of the same order as the leading contribu- tion. This makes our Monte Carlo not useful near𝛼= 2. Fortunately, there is an interesting solution. Consider extending process (1) to include another set of Poisson processes. If

(8)

identical to the first, but with its jump functionβ„Ž(𝑦) = π‘π‘¦βˆ’1/πœ” for some parameters𝑏,πœ” and cutoff π‘Œ, this adds a π‘¦βˆ’integral to (3) that can be approximated as above. Note that the sign of the second term in its expansion can be made negative by choosing πœ” >2, orπœ” large enough to keep (𝑏𝐡)πœ” term farther away. Any such πœ” could be chosen, in fact,πœ” =∞ turns out to be a good choice. For πœ” =∞, β„Ž(𝑦) = 𝑏 for π‘¦β‰€π‘Œ and zero otherwise, and the added process is effectively just one Poisson process. Its π‘¦βˆ’integral is (1βˆ’π‘π΅βˆ’π‘’βˆ’π‘π΅)π‘Œ that can be expanded in powers of𝑏𝐡. Parameter𝑏 can be chosen so as to cancel the troubling term.

The π‘₯ and π‘¦βˆ’integrals then together get approximated to βˆ’π›ΌΞ“(βˆ’π›Ό)(π‘Žπ΅)𝛼.

However, convergence is still not satisfactory, and the issue of the unsampled region near zero remains. Hence, consider extending process (1) with one more Poisson process with its jump function β„Ž(𝑦) =βˆ’π‘for π‘¦β‰€π‘Œ and zero otherwise. It is now possible to choose𝑏 and𝑐 to cancel both the (π‘Žπ΅)2 and (π‘Žπ΅)3 terms. The equations for 𝑏 and 𝑐 turn out to be cubic that can be solved to obtain

𝑏=π‘Žπ‘ž(𝑠+𝑑)π‘‹βˆ’1/𝛼, 𝑐=π‘Žπ‘ž(π‘ βˆ’π‘‘)π‘‹βˆ’1/𝛼, (20) where

𝑠=√

1/2βˆ’π‘‘2, 𝑑 = cos((πœ‹+πœƒ)/3), πœƒ= cosβˆ’1(𝑝/π‘ž3), 𝑝= 𝛼𝑋

(3βˆ’π›Ό)π‘Œ, π‘ž =

√ 𝛼𝑋

(2βˆ’π›Ό)π‘Œ . (21) As long as π‘Œ /𝑋 ≀𝛼(3βˆ’π›Ό)2/(2βˆ’π›Ό)3, this gives a solution 𝑏 β‰₯ 𝑐β‰₯ 0. To keep the higher order terms introduced by the added processes small, π‘Œ should not be too small relative to 𝑋. The next correction term is then of π’ͺ(𝑋1βˆ’4/𝛼). The region near zero now gets sampled because of negative jumps introduced. As one gets closer toπ‘Ÿ= 0, the total Poisson intensity becomes small, and hence the likelihood of getting into negative π‘Ÿβˆ’values is small.

Changes to Monte Carlo are straightforward. There is an additional positive contribution (π‘βˆ’π‘)π‘Œ to π‘šπ‘‹. Total Poisson intensity is now 𝑍𝑋 + 2π‘π‘Œ where 𝑍𝑋 =π‘Ÿπ‘‹ and π‘π‘Œ = π‘Ÿπ‘Œ. Further in step 5, the original process is chosen with probability 𝑋/(𝑋+ 2π‘Œ) and the two added processes with probabilities π‘Œ /(𝑋 + 2π‘Œ) each, and an appropriate jump is added to π‘Ÿβˆ’ (based on π‘₯ =𝑀𝑍𝑋/π‘Ÿβˆ’ or 𝑦 = π‘€π‘π‘Œ/π‘Ÿβˆ’). If π‘Ÿ does end up negative after adding βˆ’π‘ in step 5, it is set to zero (or infinitesimally small). For the present simulation results, π‘Œ is chosen to be equal to 𝑋. To improve efficiency, Sobol sequences are used to generate each of the independent random numbers.

As 𝛼 β†’ 2, π‘Ž = 𝜎[𝛼Γ(βˆ’π›Ό)]βˆ’1/𝛼 tends to zero for a given 𝜎, but 𝑏 and 𝑐 tend to a nonzero value ensuring that the square root process is simulated appropriately in the limit as trinomial branching. For processes with generic jump functions, there can be a similar issue of convergence depending on the behavior ofβ„Ž(π‘₯) asπ‘₯β†’ ∞, and a similar approach to convergence can be attempted. The algorithm is also adaptable to the case of standard mean reverting π›Όβˆ’stable LΒ΄evy processes, or more general LΒ΄evy processes (Ornstein-Uhlenbeck- Type processes) of infinite activity withπ‘Ÿβˆ’independent jump functions, extended to include negative jumps if desired. The analysis of section 1 can be carried through to obtain the well- known results. Theπ‘₯βˆ’integral then appears in an equation containing the drift term and, for simulation purpose, the process can be rewritten with a cutoff introducing an appropriately redefined drift term if necessary. In this context, the issue of convergence was addressed in Asmussen and RosiΒ΄nski (2001) with the addition of a Brownian component that effectively cancels out the (π‘Žπ΅)2 term.

(9)

4 Simulation Results

Results of the Monte Carlo simulation for constant driftπœ™ and a choice of other parameters are presented Figures 1-10 (𝑑 is set to zero). Figures 1-5 present the dependence of the probability distribution of π‘Ÿ(𝑇) at 𝑇 = 5 on 𝛼, 𝜎, π‘š and πœ™. Figure 6 shows the dependence on𝑇 itself. As can be seen from Figure 7, 𝑋 need not be too large. To understand the order of magnitude of 𝑋, note that the total intensity of Poisson processes starts off at 3π‘Ÿ(0)𝑋 that is about 10 for π‘Ÿ(0) = 0.03 and 𝑋 = 100, and corresponds to a time-step of about 0.1.

To confirm the accuracy, the Laplace exponent is computed and displayed in Figure 8 for 𝛼= 3/2,5/3 and 2 for which closed form expressions are available from section 2.

An usual approach to understanding the distribution of a positive random variable is to compare it to a lognormal one. This can be done by computing the implied Black-Scholes volatility for a call or a put option onπ‘Ÿ(𝑇) at various strikes, ignoring discounting and setting the underlying to E0(π‘Ÿ(𝑇)). The resulting volatility smile is plotted in Figure 9 for different values of 𝛼. Figure 10 shows its dependence on 𝑇. The smile features are encouraging and further study is needed to confirm their applicability.

A Semi-Analytics

Because affine processes have been well-studied, analytics of an π›Όβˆ’root process can be written down as a special case. However, for our purpose, it is simpler and more illuminating to derive the same starting with the pure-jump process

π‘‘π‘Ÿ(𝑑) = [πœ™(𝑑)βˆ’π‘šπ‘‹π‘Ÿ(𝑑)]𝑑𝑑+

∫ ∞

𝑧=0

β„Žπ‘‹(𝑧/π‘Ÿ(𝑑))𝑑𝑁(𝑑𝑧, 𝑑). (22) Here β„Žπ‘‹(π‘₯) = 0 for π‘₯ > 𝑋 given a large 𝑋 and β„Žπ‘‹(π‘₯) = β„Ž(π‘₯) for π‘₯ ≀ 𝑋. This effectively cuts off the integral over𝑧 at the higher end ensuring that the total intensity of the Poisson processes is finite. The object of interest is the following expectation value

𝐹𝑇(π‘Ÿ(𝑑), 𝑑)≑E𝑑

{ exp

[

βˆ’

∫ 𝑇

𝑑

𝑑𝑠𝑒𝑇(𝑠)π‘Ÿ(𝑠) ]}

. (23)

Its differential can be written down using Ito’s calculus leading to

βˆ‚πΉπ‘‡

βˆ‚π‘‘ + (πœ™βˆ’π‘šπ‘‹π‘Ÿ)βˆ‚πΉπ‘‡

βˆ‚π‘Ÿ βˆ’π‘’π‘‡π‘ŸπΉπ‘‡ +π‘Ÿ

∫ ∞

0

𝑑π‘₯[𝐹𝑇(π‘Ÿ+β„Žπ‘‹(π‘₯), 𝑑)βˆ’πΉπ‘‡(π‘Ÿ, 𝑑)] = 0. (24) Integration variable 𝑧 is scaled to π‘₯=𝑧/π‘Ÿ(𝑑). The above can be solved with the ansatz

𝐹𝑇(π‘Ÿ(𝑑), 𝑑) = exp [βˆ’π΄π‘‡(𝑑)βˆ’π΅π‘‡(𝑑)π‘Ÿ(𝑑)]. (25) Equating coefficients of 𝐹𝑇 independent ofπ‘Ÿ and those linear in π‘Ÿ separately gives

𝑑𝐴𝑇(𝑑)

𝑑𝑑 +πœ™(𝑑)𝐡𝑇(𝑑) = 0, 𝑑𝐡𝑇(𝑑)

𝑑𝑑 βˆ’π‘šπ‘‹π΅π‘‡(𝑑) +𝑒𝑇(𝑑) +

∫ 𝑋

0

𝑑π‘₯{1βˆ’exp [βˆ’β„Ž(π‘₯)𝐡𝑇(𝑑)]} = 0. (26)

(10)

Consider now 𝑒𝑇(𝑑) = 𝑒(𝜏) as a function of 𝜏 = 𝑇 βˆ’π‘‘ only. Then 𝐡𝑇(𝑑) = 𝐡(𝜏) is also a function of 𝜏 only, satisfying 𝐡(0) = 0 and the differential equation

𝑑𝐡(𝜏)

π‘‘πœ +π‘šπ‘‹π΅(𝜏) =𝑒(𝜏) +

∫ 𝑋

0

𝑑π‘₯{1βˆ’exp [βˆ’β„Ž(π‘₯)𝐡(𝜏)]}. (27) With β„Ž(π‘₯) assumed to go to zero as π‘₯β†’ ∞, the integrand above goes to zero as β„Ž(π‘₯)𝐡(𝜏), so that the above equation tends to be independent of 𝑋 for large 𝑋 if π‘‘π‘šπ‘‹/𝑑𝑋 =β„Ž(𝑋).

A choice for π‘šπ‘‹ with such a large 𝑋 behavior is π‘šπ‘‹ =π‘š+βˆ«π‘‹

0 𝑑π‘₯β„Ž(π‘₯) (assuming β„Ž(π‘₯) is integrable from π‘₯= 0). Equation for 𝐡(𝜏) then reads as in (3) in the limit 𝑋 β†’ ∞. Given a solution for𝐡(𝜏) satisfying 𝐡(0) = 0, and 𝐴𝑇(𝑑) expressed as an integral of𝐡(𝜏), solution for 𝐹𝑇(π‘Ÿ(𝑑), 𝑑) is as given in equation (2).

If 𝑒(𝜏) = 𝑒𝛿(πœβˆ’0+) where 𝛿(𝜏 βˆ’0+) is a Dirac-delta function sitting just above 𝜏 = 0, one obtains the Laplace transform E𝑑{exp [βˆ’π‘’π‘Ÿ(𝑇)]} of the probability density function of π‘Ÿ(𝑇), or its negative logarithm known as the Laplace exponent. For this, equation (3) is solved for 𝐡(𝜏) in the absence of 𝑒(𝜏), but under the initial condition𝐡(0) =𝑒.

For β„Ž(π‘₯) =π‘Žπ‘₯βˆ’1/𝛼, 1< 𝛼 < 2, equation for 𝐡(𝜏) reads as in (4). The π‘₯βˆ’integral in (3) isβˆ’π›ΌΞ“(βˆ’π›Ό)π‘Žπ›Ό[𝐡(𝜏)]𝛼. Note that βˆ«π‘‹

0 𝑑π‘₯β„Ž(π‘₯) =π‘Žπ›Όπ‘‹(π›Όβˆ’1)/𝛼/(π›Όβˆ’1) diverges as𝑋 β†’ ∞, but gets absorbed intoπ‘šπ‘‹. The𝛼 = 1 case is special. Theπ‘₯βˆ’integral in (27) isβˆ’π‘Žπ΅(𝜏) ln𝐡(𝜏) up to terms linear in 𝐡(𝜏) that are taken care of by π‘šπ‘‹ =π‘š+π‘Žln(𝑋/π‘Ž) +π‘Ž(1βˆ’π›Ύ) where 𝛾 is the Euler’s constant.

One may wonder whether an π›Όβˆ’root process can be defined for 𝛼 <1 as well. After all, the π‘₯βˆ’integral in (27) is then finite as 𝑋 β†’ ∞ and is βˆ’π›ΌΞ“(βˆ’π›Ό)π‘Žπ›Ό[𝐡(𝜏)]𝛼. However, the integral dominates the π‘šπ‘‹π΅(𝜏) term as 𝐡(𝜏) β†’ 0, and solving for the Laplace exponent with 𝑒(𝜏) = 0 and 𝐡(0) =𝑒 yields a 𝐡(𝜏) that does not go to zero as 𝑒→0.

References

[1] S. Asmussen and J. RosiΒ΄nski (2001), β€œApproximations of small jumps of LΒ΄evy processes with a view towards simulation”, Journal of Applied Probability 38, 482-493.

[2] J. Bertoin (2000), β€œSubordinators, LΒ΄evy Processes with no negative jumps and branch- ing processes”, MaPhySto Lecture Notes, Series 8.

[3] D. Brigo and A. Alfonsi (2005), β€œCredit default swap calibration and derivatives pricing with the SSRD stochastic intensity model ”, Finance and Stochastics 9, 29-42.

[4] P. Carr and L. Wu (2004), β€œTime-changed LΒ΄evy processes and option pricing”, Journal of Financial Economics 71, 113-141.

[5] D. Duffie, D. Filipovic and W. Schachermayer (2003), β€œAffine processes and applications in finance”, The Annals of Applied Probability 13, 984-1053.

[6] P. Tankov (2009), β€œPricing and hedging in exponential LΒ΄evy models: review of recent results”, Lecture notes, available from http://people.math.jussieu.fr/∼tankov/.

(11)

Figure 1: Plots of the probability density functions for 𝛼 = 1.65,1.80 and 1.95. Other parameters chosen are 𝑇 = 5, 𝜎 = 0.04, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is one million and cutoff 𝑋 is 100.

Figure 2: Plots of the probability density functions for 𝛼 = 1.20,1.35 and 1.50. Other parameters chosen are 𝑇 = 5, 𝜎 = 0.04, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

(12)

Figure 3: Plots of the probability density functions for 𝜎 = 0.03,0.04 and 0.05. Other parameters chosen are 𝑇 = 5, 𝛼 = 1.80, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

Figure 4: Plots of the probability density functions for π‘š = 0.05,0.0 and βˆ’0.05. Other parameters chosen are 𝑇 = 5, 𝛼 = 1.80, 𝜎 = 0.04, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

(13)

Figure 5: Plots of the probability density functions for πœ™ = 0.003,0.006 and 0.009. Other parameters chosen are 𝑇 = 5, 𝛼 = 1.80, 𝜎 = 0.04, π‘š = 0.01 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

Figure 6: Plots of the probability density functions for 𝑇 = 3,5 and 10. Other parameters chosen are𝛼= 1.80, 𝜎 = 0.04, π‘š = 0.01, πœ™ = 0.006 andπ‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

(14)

Figure 7: Plots of the probability density functions for cutoff 𝑋 = 20,100 and 500. Other parameters chosen are 𝑇 = 5, 𝛼 = 1.80, 𝜎 = 0.04, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03.

Number of Monte Carlo scenarios is 100,000.

Figure 8: Plots of the Laplace exponents computed analytically and numerically for 𝛼 = 3/2,5/3 and 2. Other parameters chosen are 𝑇 = 5, 𝜎 = 0.04, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff 𝑋 is 100.

(15)

Figure 9: Plots of the volatility smiles for 𝛼= 1.65,1.80 and 1.95. Other parameters chosen are 𝑇 = 5, 𝜎= 0.04, π‘š = 0.01, πœ™ = 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff𝑋 = 100.

Figure 10: Plots of the volatility smiles for 𝑇 = 3,5 and 10. Other parameters chosen are 𝛼 = 1.80, 𝜎 = 0.04, π‘š= 0.01, πœ™= 0.006 and π‘Ÿ(0) = 0.03. Number of Monte Carlo scenarios is 100,000 and cutoff𝑋 = 100.

Referenzen

Γ„HNLICHE DOKUMENTE

Using similar methods, we improve the best known smoothed upper bound for the popular k-means method to n O(k) , once again independent of the

SΒ΄ aez and the second author [13] introduced for the study of standard mean curvature flow can be used to obtain a new approach for mean curvature flow with obstacles that avoids

Intended for terminal users whose data is primarily textual (as opposed to formatted data entry), the Entry Assist feature provides a number of useability

The choice h(t) = p(t) is implemented by setting h(t) to be the piecewise constant hazard rate curve obtained by matching the index default swap spreads outside of the Monte

Figure 1 shows the joint default probability distribution as a function of the number of defaults over a time period of 5, 7 and 10 years using the model parameters from Table

Balochistan University of Information Technology, Engineering and Management Sciences, Quetta, Pakistan, Sardar Bahadur Khan Women University, Quetta, Pakistan. 14

It is a tractable dynamical model, computationally structured similar to the one-factor Gaussian copula model, providing easy calibration to individual hazard rate curves and

Because the LΒ΄evy subordinator model is structured similar to the Gaussian copula model, efficient pricing techniques of the latter can be directly employed in the present case1. One