+ All Categories
Home > Documents > Research Article A General Multidimensional Monte Carlo ...claim describing the net payo whose...

Research Article A General Multidimensional Monte Carlo ...claim describing the net payo whose...

Date post: 12-Feb-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
22
Research Article A General Multidimensional Monte Carlo Approach for Dynamic Hedging under Stochastic Volatility Daniel Bonetti, 1 Dorival LeΓ£o, 2 Alberto Ohashi, 3 and VinΓ­cius Siqueira 2 1 Departamento de Sistemas de Computacc ¸˜ ao, Universidade de S˜ ao Paulo, 13560-970 S˜ ao Carlos, SP, Brazil 2 Departamento de MatemΒ΄ atica Aplicada e EstatΒ΄ Δ±stica, Universidade de S˜ ao Paulo, 13560-970 S˜ ao Carlos, SP, Brazil 3 Departamento de MatemΒ΄ atica, Universidade Federal da ParaΒ΄ Δ±ba, 13560-970 Jo˜ ao Pessoa, PB, Brazil Correspondence should be addressed to Alberto Ohashi; [email protected] Received 22 August 2014; Accepted 14 December 2014 Academic Editor: Enzo Orsingher Copyright Β© 2015 Daniel Bonetti et al. is is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. We propose a feasible and constructive methodology which allows us to compute pure hedging strategies with respect to arbitrary square-integrable claims in incomplete markets. In contrast to previous works based on PDE and BSDE methods, the main merit of our approach is the flexibility of quadratic hedging in full generality without a priori smoothness assumptions on the payoff. In particular, the methodology can be applied to multidimensional quadratic hedging-type strategies for fully path-dependent options with stochastic volatility and discontinuous payoffs. In order to demonstrate that our methodology is indeed applicable, we provide a Monte Carlo study on generalized FΒ¨ ollmer-Schweizer decompositions, locally risk minimizing, and mean variance hedging strategies for vanilla and path-dependent options written on local volatility and stochastic volatility models. 1. Introduction 1.1. Background and Motivation. Let (, F, P) be a financial market composed by a continuous F-semimartingale which represents a discounted risky asset price process, F ={F ;0≀ ≀ } is a filtration which encodes the information flow in the market on a finite horizon [0, ], P is a physical probability measure, and M is the set of equivalent local martingale measures. Let be an F -measurable contingent claim describing the net payoff whose trader is faced at time . In order to hedge this claim, the trader has to choose a dynamic portfolio strategy. Under the assumption of an arbitrage-free market, the classical Galtchouk-Kunita-Watanabe (henceforth abbrevi- ated as GKW) decomposition yields = E Q [] + ∫ 0 ,Q β„“ β„“ + ,Q under Q ∈ M , (1) where ,Q is a Q-local martingale which is strongly orthog- onal to and ,Q is an adapted process. e GKW decomposition plays a crucial role in deter- mining optimal hedging strategies in a general Brownian- based market model subject to stochastic volatility . For instance, if is a one-dimensional ItΛ† o risky asset price process which is adapted to the information generated by a two- dimensional Brownian motion = ( (1) , (2) ), then there exists a two-dimensional adapted process ,Q := ( ,1 , ,2 ) such that = E Q [] + ∫ 0 ,Q , (2) which also realizes ,Q = ,1 [ ] βˆ’1 , ,Q =∫ 0 ,2 (2) ; 0 ≀ ≀ . (3) In the complete market case, there exists a unique Q ∈ M and, in this case, ,Q =0, E Q [], is the unique fair price and the hedging replicating strategy is fully described by the process ,Q . In a general stochastic volatility framework, there are infinitely many GKW orthogonal decompositions Hindawi Publishing Corporation International Journal of Stochastic Analysis Volume 2015, Article ID 863165, 21 pages http://dx.doi.org/10.1155/2015/863165
Transcript
  • Research ArticleA General Multidimensional Monte Carlo Approach forDynamic Hedging under Stochastic Volatility

    Daniel Bonetti,1 Dorival LeΓ£o,2 Alberto Ohashi,3 and VinΓ­cius Siqueira2

    1Departamento de Sistemas de ComputaccΜ§aΜƒo, Universidade de SaΜƒo Paulo, 13560-970 SaΜƒo Carlos, SP, Brazil2Departamento de Matemática Aplicada e Estat́ıstica, Universidade de SaΜƒo Paulo, 13560-970 SaΜƒo Carlos, SP, Brazil3Departamento de Matemática, Universidade Federal da Paraı́ba, 13560-970 JoaΜƒo Pessoa, PB, Brazil

    Correspondence should be addressed to Alberto Ohashi; [email protected]

    Received 22 August 2014; Accepted 14 December 2014

    Academic Editor: Enzo Orsingher

    Copyright Β© 2015 Daniel Bonetti et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

    We propose a feasible and constructive methodology which allows us to compute pure hedging strategies with respect to arbitrarysquare-integrable claims in incomplete markets. In contrast to previous works based on PDE and BSDE methods, the main meritof our approach is the flexibility of quadratic hedging in full generality without a priori smoothness assumptions on the payoff.In particular, the methodology can be applied to multidimensional quadratic hedging-type strategies for fully path-dependentoptions with stochastic volatility and discontinuous payoffs. In order to demonstrate that our methodology is indeed applicable,we provide a Monte Carlo study on generalized Föllmer-Schweizer decompositions, locally risk minimizing, and mean variancehedging strategies for vanilla and path-dependent options written on local volatility and stochastic volatility models.

    1. Introduction

    1.1. Background and Motivation. Let (𝑆, F,P) be a financialmarket composed by a continuous F-semimartingale 𝑆whichrepresents a discounted risky asset price process, F = {F

    𝑑; 0 ≀

    𝑑 ≀ 𝑇} is a filtration which encodes the information flowin the market on a finite horizon [0, 𝑇], P is a physicalprobability measure, and M𝑒 is the set of equivalent localmartingale measures. Let𝐻 be an F

    𝑇-measurable contingent

    claim describing the net payoff whose trader is faced at time𝑇. In order to hedge this claim, the trader has to choose adynamic portfolio strategy.

    Under the assumption of an arbitrage-free market, theclassical Galtchouk-Kunita-Watanabe (henceforth abbrevi-ated as GKW) decomposition yields

    𝐻 = EQ [𝐻] + βˆ«π‘‡

    0 πœƒπ»,Qβ„“

    𝑑𝑆ℓ+ 𝐿𝐻,Q

    𝑇

    under Q ∈ M𝑒, (1)

    where 𝐿𝐻,Q is aQ-local martingale which is strongly orthog-onal to 𝑆 and πœƒπ»,Q is an adapted process.

    The GKW decomposition plays a crucial role in deter-mining optimal hedging strategies in a general Brownian-based market model subject to stochastic volatility 𝜎. Forinstance, if 𝑆 is a one-dimensional ItoΜ‚ risky asset price processwhich is adapted to the information generated by a two-dimensional Brownian motion π‘Š = (π‘Š(1),π‘Š(2)), thenthere exists a two-dimensional adapted process πœ™π»,Q :=(πœ™π»,1, πœ™π»,2) such that

    𝐻 = EQ [𝐻] + βˆ«π‘‡

    0

    πœ™π»,Q𝑑

    π‘‘π‘Šπ‘‘, (2)

    which also realizes

    πœƒπ»,Q𝑑

    = πœ™π»,1

    𝑑[π‘†π‘‘πœŽπ‘‘]βˆ’1

    , 𝐿𝐻,Q𝑑

    = βˆ«π‘‘

    0

    πœ™π»,2

    π‘‘π‘Š(2)

    𝑠; 0≀ 𝑑≀ 𝑇.

    (3)

    In the complete market case, there exists a unique Q ∈M𝑒 and, in this case, 𝐿𝐻,Q = 0, EQ[𝐻], is the unique fairprice and the hedging replicating strategy is fully described bythe process πœƒπ»,Q. In a general stochastic volatility framework,there are infinitely many GKW orthogonal decompositions

    Hindawi Publishing CorporationInternational Journal of Stochastic AnalysisVolume 2015, Article ID 863165, 21 pageshttp://dx.doi.org/10.1155/2015/863165

  • 2 International Journal of Stochastic Analysis

    parameterized by the set M𝑒 and hence one can ask if itis possible to determine the notion of non-self-financingoptimal hedging strategies solely based on the quantities (3).This type of question was firstly answered by Föllmer andSondermann [1] and later on extended by Schweizer [2] andFöllmer and Schweizer [3] through the existence of the so-called Föllmer-Schweizer decomposition which turns out tobe equivalent to the existence of locally risk minimizinghedging strategies. The GKW decomposition under the so-called minimal martingale measure constitutes the startingpoint to get locally risk minimizing strategies provided thatone is able to check some square-integrability properties ofthe components in (1) under the physical measure. See, forexample, [4, 5] for details and other references therein. Seealso, for example, [6], where Fölmer-Schweizer decomposi-tions can be retrieved by solving linear backward stochasticdifferential equations (BSDEs). Orthogonal decompositionswithout square-integrability properties can also be definedin terms of the the so-called generalized Föllmer-Schweizerdecomposition (see, e.g., [7]).

    In contrast to the local risk minimization approach, onecan insist on working with self-financing hedging strategieswhich give rise to the so-called mean variance hedgingmethodology. In this approach, the spirit is to minimizethe expectation of the squared hedging error over all initialendowments π‘₯ and all suitable admissible strategies πœ‘ ∈ Θ:

    infπœ‘βˆˆΞ˜,π‘₯∈R

    EP

    𝐻 βˆ’ π‘₯ βˆ’ ∫

    𝑇

    0

    πœ‘π‘‘π‘‘π‘†π‘‘

    2

    . (4)

    The nature of the optimization problem (4) suggests to workwith the subset M𝑒

    2:= {Q ∈ M𝑒; 𝑑Q/𝑑P ∈ 𝐿2(P)}.

    Rheinlander and Schweizer [9], Gourieroux et al. [10], andSchweizer [11] show that ifM𝑒

    2̸= 0 and𝐻 ∈ 𝐿2(P), then the

    optimal quadratic hedging strategy exists and it is given by(EPΜƒ[𝐻], πœ‚

    P̃), where

    πœ‚P̃𝑑:= πœƒ

    𝐻,P̃𝑑

    βˆ’πœπ‘‘

    𝑍𝑑

    (𝑉𝐻,PΜƒπ‘‘βˆ’

    βˆ’ EPΜƒ [𝐻] βˆ’ βˆ«π‘‘

    0

    πœ‚P̃ℓ𝑑𝑆ℓ) ;

    0 ≀ 𝑑 ≀ 𝑇.

    (5)

    Here πœƒπ»,PΜƒ is computed in terms of PΜƒ; the so-called varianceoptimal martingale measure, 𝜁, realizes

    𝑍𝑑:= EPΜƒ [

    𝑑PΜƒ

    𝑑P| F𝑑] = 𝑍

    0+ ∫

    𝑑

    0

    πœβ„“π‘‘π‘†β„“; 0 ≀ 𝑑 ≀ 𝑇, (6)

    and 𝑉.𝐻,PΜƒ := EPΜƒ[𝐻 | Fβ‹…] is the value option price processunder PΜƒ. See also Černý and Kallsen [12] for the generalsemimartingale case and the works [13–15] for other utility-based hedging strategies based on GKW decompositions.

    Concrete representations for the pure hedging strategies{πœƒπ»,Q; Q = PΜ‚, PΜƒ} can in principle be obtained by comput-

    ing cross-quadratic variations 𝑑[𝑉𝐻,Q, 𝑆]𝑑/𝑑[𝑆, 𝑆]

    𝑑for Q ∈

    {P̃, P̂}. For instance, in the classical vanilla case, pure hedgingstrategies can be computed by means of the Feynman-Kac

    theorem (see, e.g., [4]). In the path-dependent case, theobtention of concrete computationally efficient representa-tions for πœƒπ»,Q is a rather difficult problem. Feynman-Kac-type arguments for fully path-dependent options mixed withstochastic volatility typically face not-well-posed problemson the whole trading period; highly degenerate PDEs arisein this context as well. Generically speaking, one has to workwith non-Markovian versions of the Feynman-Kac theoremin order to get robust dynamic hedging strategies for fullypath-dependent options written on stochastic volatility riskyasset price processes.

    In the mean variance case, the only quantity in (5)not related to GKW decomposition is 𝑍 which can inprinciple be expressed in terms of the so-called fundamentalrepresentation equations given by Hobson [16] and Biagini etal. [17] in the stochastic volatility case. For instance, Hobsonderives closed form expressions for 𝜁 and also for any typeof π‘ž-optimal measure in the Heston model [18]. Recently,semiexplicit formulas for vanilla options based on generalcharacterizations of the variance-optimal hedge in Černýand Kallsen [12] have been also proposed in the literaturewhich allow for a feasible numerical implementation in affinemodels. See Kallsen and Vierthauer [19] and Černý andKallsen [20] for some results in this direction.

    A different approach based on linear BSDEs can alsobe used in order to get useful characterizations for theoptimal hedging strategies. In this case, concrete numericalschemes for BSDEs play a key role in applications. In theMarkovian case, there are several efficient methods. See, forexample, Delong [6] and other references therein. In thenon-Markovian case, when the terminal value is allowed todepend on the whole history of a forward diffusion, thedifficulty is notorious. One fundamental issue is the imple-mentation of feasible approximations for the β€œmartingaleintegrand” of BSDEs. To the best of our knowledge, allthe existing numerical methods require a priori regularityconditions on the final condition. See, for example, [6, 21–23] and other references therein. Recently, Briand and Labart[24] use Malliavin calculus methods to compute conditionalexpectations based on Wiener chaos expansions under someregularity conditions. See also the recent results announcedby Gobet and Turkedjiev [25, 26] by using regression meth-ods.

    1.2. Contribution of the Current Paper. Themain contributionof this paper is the obtention of flexible and computationallyefficient multidimensional non-Markovian representationsfor generic option price processes which allow for a con-crete computation of the associated GKW decomposition(πœƒπ»,Q, 𝐿𝐻,Q) forQ-square-integrable payoffs𝐻withQ ∈ M𝑒.

    We provide a Monte Carlo methodology able to computeoptimal quadratic hedging strategies with respect to generalsquare-integrable claims in a multidimensional Brownian-based market model. In contrast to previous works (see, e.g.,[6] and other references therein), the main contribution ofthis paper is the formulation of a concrete numerical schemefor quadratic hedging (local risk minimization) under fullgenerality, where only square-integrability assumption is

  • International Journal of Stochastic Analysis 3

    imposed. As far as the mean variance hedging is concerned,we are able to compute pure optimal hedging strategiesπœƒπ»,PΜƒ for arbitrary square-integrable payoffs. Hence, ourmethodology also applies to this case provided that one isable to compute the fundamental representation equations inHobson [16] and Biagini et al. [17] which is the case for theclassical Heston model.

    The starting point of this paper is based on weakapproximations developed by LeaΜƒo and Ohashi [27] forone-dimensional Brownian functionals. They introduced aone-dimensional space-filtration discretization scheme con-structed from suitable waiting times which measure theinstants when the Brownian motion hits some a priori levels.In the present work, we extend [27] in one direction: weprovide a feasible numerical scheme for multidimensionalQ-GKW decompositions under rather weak integrabilityconditions for a given Q ∈ M𝑒. In order to apply ourmethodology for hedging, we analyze the convergence ofour approximating hedging strategies to the respective valueprocesses in a Brownian-based incompletemarket setup.Thisallows us to perform quadratic hedging for generic square-integrable payoffs written on stochastic volatility models.Thenumerical scheme of this work can also be viewed as part of amore general theory concerning a weak version of functionalItoΜ‚ calculus (see [28, 29]) as introduced by Ohashi et al.[30]. We implement the multidimensional weak derivativeoperators defined in [30] in the pure martingale case to solvehedging problems in generic stochastic volatility models.

    In this paper, themultidimensional numerical scheme formartingale representations lies in the exact simulation of ani.i.d sequence of increments of hitting times

    π‘‡π‘˜,𝑗

    𝑛:= inf {𝑑 > π‘‡π‘˜,𝑗

    π‘›βˆ’1:π‘Š(𝑗)

    π‘‘βˆ’π‘Š

    (𝑗)

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    = πœ–π‘˜} ; 𝑛 β‰₯ 1, (7)

    where π‘‡π‘˜,𝑗0

    := 0 for 1 ≀ 𝑗 ≀ 𝑝 and πœ–π‘˜β†“ 0 as π‘˜ β†’ ∞. The

    fundamental object which allows us to obtain a numericalscheme for πœƒπ»,Q is the following ratio:

    EQ[[

    [

    EQ [𝐻 | Fπ‘˜

    π‘‡π‘˜,𝑗

    1

    ] βˆ’ EQ [𝐻 | Fπ‘˜

    πœπ‘˜,𝑗

    1

    ]

    π‘Š(𝑗)

    π‘‡π‘˜,𝑗

    1

    ]]

    ]

    , 𝑗 = 1, . . . , 𝑑,

    (8)

    for 𝑑 ≀ 𝑝, where πœπ‘˜,𝑗1

    := max{π‘‡π‘˜π‘›; π‘‡π‘˜π‘›< 𝑇

    π‘˜,𝑗

    1}, π‘‡π‘˜0:= 0, and

    π‘‡π‘˜

    𝑛:= inf1≀𝑗≀𝑝

    π‘šβ‰₯1

    {π‘‡π‘˜,𝑗

    π‘š; π‘‡π‘˜,𝑗

    π‘šβ‰₯ 𝑇

    π‘˜

    π‘›βˆ’1} , (9)

    for 𝑛 β‰₯ 1. Here, there are 𝑑 asset price processes drivenby a 𝑝-dimensional Brownian motion (π‘Š(1), . . . ,π‘Š(𝑝)). Byapproximating the payoff 𝐻 in terms of functionals of therandom walks

    π΄π‘˜,𝑗

    𝑑:= π‘Š

    (𝑗)

    π‘‡π‘˜,𝑗

    𝑛

    on {π‘‡π‘˜,𝑗𝑛

    ≀ 𝑑 < π‘‡π‘˜,𝑗

    𝑛+1} , (10)

    we will take advantage of the discrete structure of the sigma-algebras in (37) to evaluate (8) by standard Monte Carlo

    methods. The information set contained in (Fπ‘˜π‘‡π‘˜,𝑗

    1

    ,Fπ‘˜πœπ‘˜,𝑗

    1

    ) isperfectly implementable by using the algorithm proposedby Burq and Jones [8]. We leave the implementation ofsimulation-regression method for a further study.

    In order to demonstrate that our methodology is indeedapplicable, we provide a Monte Carlo study on generalizedFöllmer-Schweizer decompositions, locally risk minimizingand mean variance hedging strategies for vanilla and path-dependent options written on local volatility and stochas-tic volatility models. The numerical experiments suggestthat pure hedging strategies based on generalized Föllmer-Schweizer decompositions mitigate very well the cost ofhedging of a path-dependent option even if there is no guar-antee of the existence of locally risk minimizing strategies.We also compare hedging errors arising from optimal meanvariance hedging strategies for one-touch options written ona Heston model with nonzero correlation.

    Lastly, we want to emphasize the fact that it is our chiefgoal is to provide a feasible numerical method which worksin full generality. In this case, the price we pay is to workwith weak convergence results instead of 𝐿𝑝 or uniformconvergence in probability. We leave a more refined analysison error estimates and rates of convergence underMarkovianassumptions to a future research.

    The remainder of this paper is structured as follows.In Section 2, we fix the notation and we describe the basicunderlying market model. In Section 3, we provide the basicelements of the Monte Carlo methodology proposed in thispaper. In Section 4, we formulate dynamic hedging strategiesstarting from a given GKW decomposition and we translateour results to well-known quadratic hedging strategies. TheMonteCarlo algorithmand the numerical study are describedin Sections 5 and 6, respectively.TheAppendix presentsmorerefined approximations when the martingale representationsadmit additional hypotheses.

    2. Preliminaries

    Throughout this paper, we assume that we are in the usualBrownian market model with finite time horizon 0 <𝑇 < ∞ equipped with the stochastic basis (Ξ©, F,P)generated by a standard 𝑝-dimensional Brownian motion𝐡 = {(𝐡

    (1)

    𝑑, . . . , 𝐡

    (𝑝)

    𝑑); 0 ≀ 𝑑 < ∞} starting from 0. The

    filtration F := (F𝑑)0≀𝑑≀𝑇

    is the P-augmentation of the naturalfiltration generated by 𝐡. For a given π‘š-dimensional vector𝐽 = (𝐽

    1, . . . , 𝐽

    π‘š), we denote by diag(𝐽) the π‘š Γ— π‘š diagonal

    matrix whose β„“th diagonal term is 𝐽ℓ. In this paper, for

    all unexplained terminology concerning general theory ofprocesses, we refer to Dellacherie and Meyer [31].

    In view of stochastic volatility models, let us split 𝐡 intotwo multidimensional Brownian motions as follows: 𝐡𝑆 :=(𝐡(1), . . . , 𝐡(𝑑)) and𝐡𝐼 := (𝐡(𝑑+1), . . . , 𝐡(𝑝)). In this section, themarket consists of 𝑑+1 assets (𝑑 ≀ 𝑝): one riskless asset givenby

    𝑑𝑆0

    𝑑= π‘Ÿπ‘‘π‘†0

    𝑑𝑑𝑑, 𝑆

    0

    0= 1; 0 ≀ 𝑑 ≀ 𝑇, (11)

  • 4 International Journal of Stochastic Analysis

    and a 𝑑-dimensional vector of risky assets 𝑆 := (𝑆1, . . . , 𝑆𝑑)which satisfies the following stochastic differential equation:

    𝑑𝑆𝑑= diag (𝑆

    𝑑) (𝑏

    𝑑𝑑𝑑 + 𝜎

    𝑑𝑑𝐡𝑆

    𝑑) , 𝑆

    0= π‘₯ ∈ R

    𝑑;

    0 ≀ 𝑑 ≀ 𝑇.

    (12)

    Here, the real-valued interest rate process π‘Ÿ = {π‘Ÿπ‘‘; 0 ≀

    𝑑 ≀ 𝑇}, the vector of mean rates of return 𝑏 := {𝑏𝑑=

    (𝑏1𝑑, . . . , 𝑏𝑑

    𝑑); 0 ≀ 𝑑 ≀ 𝑇}, and the volatility matrix 𝜎 :=

    {πœŽπ‘‘= (𝜎

    𝑖𝑗

    𝑑); 1 ≀ 𝑖 ≀ 𝑑, 1 ≀ 𝑗 ≀ 𝑑, 0 ≀ 𝑑 ≀ 𝑇}

    are assumed to be predictable and they satisfy the standardassumptions in such way that both 𝑆0 and 𝑆 are well-definedpositive semimartingales. We also assume that the volatilitymatrix 𝜎 is nonsingular for almost all (𝑑, πœ”) ∈ [0, 𝑇] Γ— Ξ©. Thediscounted price 𝑆 := {𝑆

    𝑖:= 𝑆

    𝑖

    /𝑆0; 𝑖 = 1, . . . , 𝑑} follows

    𝑑𝑆𝑑= diag (𝑆

    𝑑) [(𝑏

    π‘‘βˆ’ π‘Ÿπ‘‘1𝑑) 𝑑𝑑 + 𝜎

    𝑑𝑑𝐡𝑆

    𝑑] ; 𝑆

    0= π‘₯ ∈ R

    𝑑,

    0 ≀ 𝑑 ≀ 𝑇,

    (13)

    where 1𝑑is a d-dimensional vector with every component

    equal to 1. The market price of risk is given by

    πœ“π‘‘:= 𝜎

    βˆ’1

    𝑑[π‘π‘‘βˆ’ π‘Ÿπ‘‘1𝑑] , 0 ≀ 𝑑 ≀ 𝑇, (14)

    where we assume

    βˆ«π‘‡

    0

    πœ“π‘’2

    R𝑑𝑑𝑒 < ∞ a.s. (15)

    In the sequel,M𝑒 denotes the set of P-equivalent prob-ability measures Q such that, respectively, Radon-Nikodymderivative process is aP-martingale and the discounted price𝑆 is a Q-local martingale. Throughout this paper, we assumethatM𝑒 ΜΈ= 0. In our setup, it is well known thatM𝑒 is givenby the subset of probability measures with Radon-Nikodymderivatives of the form

    𝑑Q

    𝑑P:= exp [βˆ’βˆ«

    𝑇

    0

    πœ“π‘’π‘‘π΅π‘†

    π‘’βˆ’ ∫

    𝑇

    0

    ]𝑒𝑑𝐡𝐼

    𝑒

    βˆ’1

    2βˆ«π‘‡

    0

    {πœ“π‘’

    2

    R𝑑+]𝑒

    2

    Rπ‘βˆ’π‘‘} 𝑑𝑒] ,

    (16)

    for some Rπ‘βˆ’π‘‘-valued adapted process ] such thatβˆ«π‘‡

    0β€–]𝑑‖2Rπ‘βˆ’π‘‘

    𝑑𝑑 < ∞ a.s.

    Example 1. The typical example studied in the literature is thefollowing one-dimensional stochastic volatility model:

    𝑑𝑆𝑑= π‘†π‘‘πœ‡ (𝑑, 𝑆

    𝑑, πœŽπ‘‘) 𝑑𝑑 + 𝑆

    π‘‘πœŽπ‘‘π‘‘π‘Œ(1)

    𝑑,

    π‘‘πœŽ2

    𝑑= π‘Ž (𝑑, 𝑆

    𝑑, πœŽπ‘‘) 𝑑𝑑 + 𝑏 (𝑑, 𝑆

    𝑑, πœŽπ‘‘) π‘‘π‘Œ

    (2)

    𝑑; 0 ≀ 𝑑 ≀ 𝑇,

    (17)

    where π‘Œ(1) and π‘Œ(2) are correlated Brownian motions withcorrelation 𝜌 ∈ [βˆ’1, 1] and πœ‡, π‘Ž, and 𝑏 are suitablefunctions such that (𝑆, 𝜎2) is a well-defined two-dimensional

    Markov process. All continuous stochastic volatility modelscommonly used in practice fit into specification (17). Inthis case, 𝑝 = 2 > 𝑑 = 1 and we recall that themarket is incomplete where the set M𝑒 is infinity. Thedynamic hedging procedure turns out to be quite challengingdue to extrinsic randomness generated by the nontradeablevolatility, specially with respect to to exotic options.

    2.1. GKWDecomposition. In the sequel, we takeQ ∈ M𝑒 andwe setπ‘Šπ‘† := (π‘Š(1), . . . ,π‘Š(𝑑)) andπ‘ŠπΌ := (π‘Š(𝑑+1), . . . ,π‘Š(𝑝)),where

    π‘Š(𝑗)

    𝑑:=

    {{{{

    {{{{

    {

    𝐡(𝑗)

    𝑑+ ∫

    𝑑

    0

    πœ“π‘—π‘’π‘‘π‘’, 𝑗 = 1, . . . , 𝑑

    𝐡(𝑗)

    𝑑+ ∫

    𝑑

    0

    ]𝑗𝑒𝑑𝑒, 𝑗 = 𝑑 + 1, . . . , 𝑝; 0 ≀ 𝑑 ≀ 𝑇,

    (18)

    is a standard 𝑝-dimensional Brownian motion under themeasure Q and filtration F := {F

    𝑑; 0 ≀ 𝑑 ≀ 𝑇} generated by

    π‘Š = (π‘Š(1), . . . ,π‘Š(𝑝)). In what follows, we fix a discountedcontingent claim 𝐻. Recall that the filtration F is containedin F, but it is not necessarily equal. In the remainder of thispaper, we assume the following hypothesis.

    (M)The contingent claim𝐻 is alsoF𝑇-measurable.

    Remark 2. Assumption (M) is essential for the approachtaken in this work because the whole algorithm is basedon the information generated by the Brownian motion π‘Š(defined under themeasureQ and filtration F). As long as theshort rate is deterministic, this hypothesis is satisfied for anystochastic volatility model of form (17) and a payoffΞ¦(𝑆

    𝑑; 0 ≀

    𝑑 ≀ 𝑇) where Ξ¦ : C𝑇→ R is a Borel map and C

    𝑇is the

    usual space of continuous paths on [0, 𝑇]. Hence, (M) holdsfor a very large class of examples founded in practice.

    For a given Q-square-integrable claim 𝐻, the Brownianmartingale representation (computed in terms of (F ,Q))yields

    𝐻 = EQ [𝐻] + βˆ«π‘‡

    0

    πœ™π»,Q𝑒

    π‘‘π‘Šπ‘’, (19)

    where πœ™π»,Q := (πœ™π»,Q,1, . . . , πœ™π»,Q,𝑝) is a 𝑝-dimensional F-predictable process. In what follows, we set πœ™π»,Q,𝑆 := (πœ™π»,Q,1,. . . , πœ™π»,Q,𝑑), πœ™π»,Q,𝐼 := (πœ™π»,Q,𝑑+1, . . . , πœ™π»,Q,𝑝), and

    𝐿𝐻,Q𝑑

    := βˆ«π‘‘

    0

    πœ™π»,Q,𝐼𝑒

    π‘‘π‘ŠπΌ

    𝑒, οΏ½Μ‚οΏ½

    𝑑:= EQ [𝐻 | F𝑑] ;

    0 ≀ 𝑑 ≀ 𝑇.

    (20)

    The discounted stock price process has the following Q-dynamics:

    𝑑𝑆𝑑= diag (𝑆

    𝑑) πœŽπ‘‘π‘‘π‘Š

    𝑆

    𝑑, 𝑆

    0= π‘₯, 0 ≀ 𝑑 ≀ 𝑇, (21)

  • International Journal of Stochastic Analysis 5

    and therefore the Q-GKW decomposition for the pair oflocally square-integrable local martingales (οΏ½Μ‚οΏ½, 𝑆) is given by

    �̂�𝑑= EQ [𝐻] + ∫

    𝑑

    0

    πœ™π»,Q,𝑆𝑒

    π‘‘π‘Šπ‘†

    𝑒+ 𝐿𝐻,Q𝑑

    = EQ [𝐻] + βˆ«π‘‘

    0

    πœƒπ»,Q𝑒

    𝑑𝑆𝑒+ 𝐿𝐻,Q𝑑

    ; 0 ≀ 𝑑 ≀ 𝑇,

    (22)

    where

    πœƒπ»,Q

    := πœ™π»,Q,𝑆

    [diag(𝑆)𝜎]βˆ’1 . (23)

    The 𝑝-dimensional process πœ™π»,Q which constitutes (20) and(23) plays a major role in several types of hedging strategiesin incomplete markets and it will be our main object of study.

    Remark 3. If we set ]𝑗 = 0 for 𝑗 = 𝑑 + 1, . . . , 𝑝 and the cor-respondent density process is a martingale, then the resultingminimalmartingalemeasure PΜ‚ yields a GKWdecompositionwhere 𝐿𝐻,PΜ‚ is still a P-local martingale orthogonal to themartingale component of 𝑆 under P. In this case, it is alsonatural to implement a pure hedging strategy based onπœƒπ»,PΜ‚ regardless of the existence of the Föllmer-Schweizerdecomposition. If this is the case, this hedging strategy can bebased on the generalized Föllmer-Schweizer decomposition(see, e.g., Th. 9 in [7]).

    3. The Random Skeleton and WeakApproximations for GKW Decompositions

    In this section, we provide the fundamentals of the numericalalgorithm of this paper for the obtention of hedging strategiesin complete and incomplete markets.

    3.1. The Multidimensional Random Skeleton. At first, we fixonce and for all Q ∈ M𝑒 and a Q-square-integrablecontingent claim 𝐻 satisfying (M). In the remainder of thissection, we are going to fix aQ-Brownianmotionπ‘Š andwitha slight abuse of notation all Q-expectations will be denotedby E. The choice of Q ∈ M𝑒 is dictated by the pricing andhedging method used by the trader.

    In the sequel, [β‹…, β‹…] denotes the usual quadratic variationbetween semimartingales and the usual jump of a process isdenoted by Ξ”π‘Œ

    𝑑= π‘Œ

    π‘‘βˆ’ π‘Œ

    π‘‘βˆ’where π‘Œ

    π‘‘βˆ’is the left-hand limit

    of a cadlag process π‘Œ. For a pair (π‘Ž, 𝑏) ∈ R2, we denote π‘Ž βˆ¨π‘ := max{π‘Ž, 𝑏} and π‘Ž ∧ 𝑏 := min{π‘Ž, 𝑏}. Moreover, for anytwo stopping times 𝑆 and 𝐽, we denote the stochastic intervalsβŸ¦π‘†, 𝐽⟦ := {(πœ”, 𝑑); 𝑆(πœ”) ≀ 𝑑 < 𝐽(πœ”)}, βŸ¦π‘†βŸ§ := {(πœ”, 𝑑); 𝑆(πœ”) = 𝑑}and so on.Throughout this article, Leb denotes the Lebesguemeasure on the interval [0, 𝑇].

    For a fixed positive integer π‘˜ and for each 𝑗 = 1, 2, . . . , 𝑝we define π‘‡π‘˜,𝑗

    0:= 0 a.s. and

    π‘‡π‘˜,𝑗

    𝑛:= inf {π‘‡π‘˜,𝑗

    π‘›βˆ’1< 𝑑 < ∞;

    π‘Š(𝑗)

    π‘‘βˆ’π‘Š

    (𝑗)

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    = 2βˆ’π‘˜} ,

    𝑛 β‰₯ 1,

    (24)

    where π‘Š := (π‘Š(1), . . . ,π‘Š(𝑝)) is the 𝑝-dimensional Q-Brownian motion as defined in (18).

    For each 𝑗 ∈ {1, . . . , 𝑝}, the familyTπ‘˜,𝑗 := {π‘‡π‘˜,𝑗𝑛; 𝑛 β‰₯ 0} is

    a sequence of F-stopping times where the increments {π‘‡π‘˜,𝑗𝑛

    βˆ’

    π‘‡π‘˜,𝑗

    π‘›βˆ’1; 𝑛 β‰₯ 1} are an i.i.d sequence with the same distribution

    asπ‘‡π‘˜,𝑗1. In the sequel, we defineπ΄π‘˜ := (π΄π‘˜,1, . . . , π΄π‘˜,𝑝) as the𝑝-

    dimensional step process given in a component-wise mannerby

    π΄π‘˜,𝑗

    𝑑:=

    ∞

    βˆ‘π‘›=1

    2βˆ’π‘˜πœ‚π‘˜,𝑗

    𝑛1{π‘‡π‘˜,𝑗

    𝑛 ≀𝑑}; 0 ≀ 𝑑 ≀ 𝑇, (25)

    where

    πœ‚π‘˜,𝑗

    𝑛:=

    {{{{{

    {{{{{

    {

    1; if π‘Š(𝑗)π‘‡π‘˜,𝑗

    𝑛

    βˆ’π‘Š(𝑗)

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    = 2βˆ’π‘˜, π‘‡π‘˜,𝑗𝑛

    < ∞

    βˆ’1; if π‘Š(𝑗)π‘‡π‘˜,𝑗

    𝑛

    βˆ’π‘Š(𝑗)

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    = βˆ’2βˆ’π‘˜, π‘‡π‘˜,𝑗𝑛

    < ∞

    0; if π‘‡π‘˜,𝑗𝑛

    = ∞

    (26)

    for π‘˜, 𝑛 β‰₯ 1, and 𝑗 = 1, . . . , 𝑝. We split π΄π‘˜ into (𝐴𝑆,π‘˜, 𝐴𝐼,π‘˜)where 𝐴𝑆,π‘˜ is the 𝑑-dimensional process constituted by thefirst 𝑑 components of π΄π‘˜ and 𝐴𝐼,π‘˜ and the remainder of the𝑝 βˆ’ 𝑑-dimensional process. Let Fπ‘˜,𝑗 := {Fπ‘˜,𝑗

    𝑑: 0 ≀ 𝑑 ≀ 𝑇}

    be the natural filtration generated by {π΄π‘˜,𝑗𝑑; 0 ≀ 𝑑 ≀ 𝑇}. One

    should notice that Fπ‘˜,𝑗 is a discrete-type filtration in the sensethat

    Fπ‘˜,𝑗

    𝑑=

    ∞

    ⋁ℓ=0

    (Fπ‘˜,𝑗

    π‘‡π‘˜,𝑗

    β„“

    ∩ {π‘‡π‘˜,𝑗

    ℓ≀ 𝑑 < 𝑇

    π‘˜,𝑗

    β„“+1}) , 0 ≀ 𝑑 ≀ 𝑇, (27)

    where Fπ‘˜,𝑗0

    = {Ξ©, 0} and Fπ‘˜,π‘—π‘‡π‘˜,𝑗

    π‘š

    = 𝜎(π‘‡π‘˜,𝑗

    1, . . . , π‘‡π‘˜,𝑗

    π‘š, πœ‚π‘˜,𝑗

    1, . . . ,

    πœ‚π‘˜,π‘—π‘š) forπ‘š β‰₯ 1 and 𝑗 = 1, . . . , 𝑝. In (27),∨denotes the smallest

    sigma-algebra generated by the union. One can easily checkthatFπ‘˜,𝑗

    π‘‡π‘˜,𝑗

    π‘š

    = 𝜎(π΄π‘˜,𝑗

    π‘ βˆ§π‘‡π‘˜,𝑗

    π‘š

    ; 𝑠 β‰₯ 0) and hence

    Fπ‘˜,𝑗

    π‘‡π‘˜,𝑗

    π‘š

    = Fπ‘˜,𝑗

    𝑑a.s on {π‘‡π‘˜,𝑗

    π‘šβ‰€ 𝑑 < 𝑇

    π‘˜,𝑗

    π‘š+1} . (28)

    With a slight abuse of notation, we write Fπ‘˜,𝑗𝑑to denote its

    Q-augmentation satisfying the usual conditions.Let us now introduce the multidimensional filtration

    generated by π΄π‘˜. Let us consider Fπ‘˜ := {Fπ‘˜π‘‘; 0 ≀ 𝑑 ≀ 𝑇}

    where Fπ‘˜π‘‘:= Fπ‘˜,1

    π‘‘βŠ— Fπ‘˜,2

    π‘‘βŠ— β‹… β‹… β‹… βŠ— F

    π‘˜,𝑝

    𝑑for 0 ≀ 𝑑 ≀ 𝑇. Let

    Tπ‘˜ := {π‘‡π‘˜π‘š; π‘š β‰₯ 0} be the order statistics obtained from the

    family of random variables {π‘‡π‘˜,𝑗ℓ; β„“ β‰₯ 0; 𝑗 = 1, . . . , 𝑝}.That is,

    we set π‘‡π‘˜0:= 0,

    π‘‡π‘˜

    1:= inf1≀𝑗≀𝑝

    {π‘‡π‘˜,𝑗

    1} , 𝑇

    π‘˜

    𝑛:= inf1≀𝑗≀𝑝

    π‘šβ‰₯1

    {π‘‡π‘˜,𝑗

    π‘š; π‘‡π‘˜,𝑗

    π‘šβ‰₯ 𝑇

    π‘˜

    π‘›βˆ’1} (29)

    for 𝑛 β‰₯ 1. In this case, Tπ‘˜ is the partition generated byall stopping times defined in (24). The finite-dimensionaldistribution of π‘Š(𝑗) is absolutely continuous for each 𝑗 =1, . . . , 𝑝 and therefore the elements of Tπ‘˜ are almost surelydistinct for every π‘˜ β‰₯ 1. The following result is an immediateconsequence of our construction.

  • 6 International Journal of Stochastic Analysis

    Lemma 4. For every π‘˜ β‰₯ 1, the set Tπ‘˜ is a sequence of Fπ‘˜-stopping times such that

    Fπ‘˜

    𝑑= F

    π‘˜

    π‘‡π‘˜π‘›

    π‘Ž.𝑠 π‘œπ‘› {π‘‡π‘˜

    𝑛≀ 𝑑 < 𝑇

    π‘˜

    𝑛+1} , (30)

    for each 𝑛 β‰₯ 0 and π‘˜ β‰₯ 1.

    ItoΜ‚ representation theorem yields

    E [𝐻 | F𝑑] = E [𝐻] + ∫

    𝑑

    0

    πœ™π»

    π‘’π‘‘π‘Š

    𝑒; 0 ≀ 𝑑 ≀ 𝑇, (31)

    where πœ™π» is a 𝑝-dimensional F-predictable process such that

    Eβˆ«π‘‡

    0

    πœ™π»

    𝑑

    2

    R𝑝𝑑𝑑 < ∞. (32)

    The payoff 𝐻 induces the Q-square-integrable F-martingale𝑋𝑑:= E[𝐻 | F

    𝑑], 0 ≀ 𝑑 ≀ 𝑇. We now embed the process 𝑋

    into the filtration Fπ‘˜ by means of the following operator:

    (π›Ώπ‘˜π‘‹)𝑑:=

    ∞

    βˆ‘π‘š=0

    E [π‘‹π‘‡π‘˜π‘š| F

    π‘˜

    π‘‡π‘˜π‘š

    ] 1{π‘‡π‘˜π‘šβ‰€π‘‘

  • International Journal of Stochastic Analysis 7

    3.2. Weak Approximation for the Hedging Process. Based on(20), (22), and (23), let us denote

    πœƒπ»

    𝑑:= πœ™

    𝐻,𝑆

    𝑑[diag(𝑆

    𝑑)πœŽπ‘‘]βˆ’1

    , 𝐿𝐻

    𝑑:= E [𝐻] + ∫

    𝑑

    0

    πœ™π»,𝐼

    β„“π‘‘π‘Š

    𝐼

    β„“;

    0 ≀ 𝑑 ≀ 𝑇.

    (43)

    In order to shorten notation, we do not write (πœ™π»,Q,𝑆, πœ™π»,Q,𝐼)in (43). The main goal of this section is the obtention ofbounded variation martingale weak approximations for bothgain and cost processes, given, respectively, by

    βˆ«π‘‘

    0

    πœƒπ»

    𝑒𝑑𝑆𝑒, 𝐿

    𝐻

    𝑑; 0 ≀ 𝑑 ≀ 𝑇. (44)

    We assume the trader has some knowledge of the underlyingvolatility so that the obtention of πœ™π»,𝑆 will be sufficientto recover πœƒπ». The typical example we have in mind isgeneralized Föllmer-Schweizer decompositions and locallyrisk minimizing and mean variance strategies as explainedin the Introduction. The scheme will be very constructive insuch way that all the elements of our approximation will beamenable to a feasible numerical analysis. Under very mildintegrability conditions, theweak approximations for the gainprocess will be translated into the physical measure.

    TheWeak Topology. In order to obtain approximation resultsunder full generality, it is important to consider a topologywhich is flexible to deal with nonsmooth hedging strategiesπœƒπ» for possibly non-Markovian payoffs 𝐻 and at the sametime justifiesMonte Carlo procedures. In the sequel, wemakeuse of the weak topology 𝜎(𝐡𝑝,π‘€π‘ž) of the Banach space𝐡𝑝(F) constituted by F-optional processes π‘Œ such that

    Eπ‘Œβˆ—

    𝑇

    𝑝

    < ∞, (45)

    where π‘Œβˆ—π‘‡:= sup

    0≀𝑑≀𝑇|π‘Œπ‘‘| and 1 ≀ 𝑝, π‘ž < ∞ such that 1/𝑝 +

    1/π‘ž = 1.The subspace of the square-integrable F-martingaleswill be denoted by 𝐻2(F). It will be also useful to work with𝜎(𝐡1, Ξ›βˆž)-topology given in [27]. For more details aboutthese topologies, we refer to the works [27, 31, 32]. It turns outthat 𝜎(𝐡2,𝑀2) and 𝜎(𝐡1, Ξ›βˆž) are very natural notions to dealwith generic square-integrable random variables as describedin [27].

    In the sequel, we recall the following notion of covariationintroduced in [27, 30].

    Definition 7. Let {π‘Œπ‘˜; π‘˜ β‰₯ 1} be a sequence of square-integrable Fπ‘˜-martingales. One says that {π‘Œπ‘˜; π‘˜ β‰₯ 1} has 𝛿-covariation with respect to jth component of π΄π‘˜ if the limit

    limπ‘˜β†’βˆž

    [π‘Œπ‘˜, π΄π‘˜,𝑗]𝑑

    (46)

    exists weakly in 𝐿1(Q) for every 𝑑 ∈ [0, 𝑇].

    The covariation notion in Definition 7 slightly differsfrom [27, 30] because {π‘Œπ‘˜; π‘˜ β‰₯ 1} is not necessarily a

    sequence of pure jump Fπ‘˜-adapted process. In fact, since weare in the puremartingale case, we will relax such assumptionas demonstrated by the following Lemma.

    Lemma 8. Let {π‘Œπ‘˜,𝑗 = βˆ«β‹…0π»π‘˜,π‘—π‘ π‘‘π΄π‘˜,𝑗; π‘˜ β‰₯ 1, 𝑗 = 1, . . . , 𝑝} be

    a sequence of stochastic integrals and π‘Œπ‘˜ := βˆ‘π‘π‘—=1

    π‘Œπ‘˜,𝑗. Assumethat

    supπ‘˜β‰₯1

    E [π‘Œπ‘˜, π‘Œπ‘˜]𝑇< ∞. (47)

    Then π‘Œπ‘— := limπ‘˜β†’βˆž

    π‘Œπ‘˜,𝑗 exists weakly in 𝐡2(F) for each 𝑗 =1, . . . , 𝑝 with π‘Œπ‘— ∈ 𝐻2(F) if and only if {π‘Œπ‘˜; π‘˜ β‰₯ 1} admits𝛿-covariation with respect to jth component of π΄π‘˜. In this case,

    limπ‘˜β†’βˆž

    [π‘Œπ‘˜, π΄π‘˜,𝑗]𝑑

    = limπ‘˜β†’βˆž

    [π‘Œπ‘˜,𝑗, π΄π‘˜,𝑗]𝑑

    = [π‘Œπ‘—,π‘Š(𝑗)]𝑑

    π‘€π‘’π‘Žπ‘˜π‘™π‘¦ 𝑖𝑛 𝐿1(Q) ; 𝑑 ∈ [0, 𝑇] ,

    (48)

    for 𝑗 = 1, . . . , 𝑝.

    Proof. Let π‘π‘˜π‘‘:= E[π‘Œπ‘˜

    𝑑| F

    𝑑], 0 ≀ 𝑑 ≀ 𝑇, be a sequence

    of F-square-integrable martingales. Similar to Lemma 4.2in [30] or Lemma 3.2 in [27], one can easily check thatassumption (47) implies that {π‘Œπ‘˜; π‘˜ β‰₯ 1} is 𝐡2(F)-weaklyrelatively sequentially compact where all limit points areF-square-integrable martingales. Moreover, since π‘Œπ‘˜ is asquare-integrable Fπ‘˜-martingale, we will repeat the sameargument given in Lemma 3.5 in [27] to safely state that

    limπ‘˜β†’βˆž

    [π‘Œπ‘˜π‘– , 𝐴

    π‘˜π‘– ,𝑗]𝑑= [𝑍,π‘Š

    𝑗]𝑑

    weakly in 𝐿1 (Q) ; 0 ≀ 𝑑 ≀ 𝑇,(49)

    for any 𝐡2(F)-weakly convergent subsequence wherelimπ‘–β†’βˆž

    π‘Œπ‘˜π‘– = 𝑍 ∈ H2(F). The multidimensional version ofthe Brownian motion martingale representation theoremallows us to conclude the proof.

    In the sequel, wemake use of the following notion of weakfunctional derivative introduced in [27, 30].

    Definition 9. Let 𝐻 be a Q-square-integrable contingentclaim satisfying (M) and one sets𝑋

    𝑑= E[𝐻 | F

    𝑑], 0 ≀ 𝑑 ≀ 𝑇.

    We say that𝑋 is weakly differentiable if

    D𝑗𝑋 := lim

    π‘˜β†’βˆž

    Dπ‘˜,𝑗𝑋 exists weakly in 𝐿2 (Q Γ— Leb) (50)

    for each 𝑗 = 1, . . . , 𝑝. In this case, we set D𝑋 :=(D1𝑋, . . . ,D𝑝𝑋).

    In Leão and Ohashi [27] and Ohashi et al. [30], theauthors introduce this notion of differential calculus whichproves to be a weak version of the pathwise functional Itôcalculus developed by Dupire [28] and further studied by

  • 8 International Journal of Stochastic Analysis

    Cont and Fournié [29]. We refer the reader to these worksfor further details. The following result is an immediateconsequence of Proposition 3.1 in [30]. See alsoTh. 4.1 in [27]for the one-dimensional case.

    Lemma 10. Let 𝐻 be a Q-square-integrable contingent claimsatisfying (M). Then the F-martingale 𝑋

    𝑑= E[𝐻 | F

    𝑑], 0 ≀

    𝑑 ≀ 𝑇, is weakly differentiable and

    D𝑋 = (πœ™π»,1

    , . . . , πœ™π»,𝑝

    ) . (51)

    In particular,

    limπ‘˜β†’βˆž

    𝑑

    βˆ‘π‘—=1

    βˆ«β‹…

    0

    Dπ‘˜,π‘—π‘‹π‘ π‘‘π΄π‘˜,𝑗

    𝑠=

    𝑑

    βˆ‘π‘—=1

    βˆ«β‹…

    0

    πœ™π»,𝑗

    π‘’π‘‘π‘Š

    (𝑗)

    𝑒= ∫

    β‹…

    0

    πœƒπ»

    𝑒𝑑𝑆𝑒

    π‘€π‘’π‘Žπ‘˜π‘™π‘¦ 𝑖𝑛 𝐡2(F) .

    (52)

    The result in Lemma 10 in not sufficient to implementdynamic hedging strategies based on Dπ‘˜,𝑗𝑋, 𝑗 = 1, . . . , 𝑝.In order to ensure that our hedging strategies are nonantic-ipative, we need to study the limiting behavior of Dπ‘˜,𝑗𝑋 asπ‘˜ β†’ ∞. It turns out that they share the same asymptoticbehavior as follows. In the sequel, ∫Dπ‘˜,𝑗𝑋

    π‘ π‘‘π΄π‘˜,𝑗𝑠denotes the

    usual stochastic integral with respect to the square-integrableFπ‘˜-martingale π΄π‘˜,𝑗.

    Theorem 11. Let𝐻 be aQ-square-integrable contingent claimsatisfying (M). Then

    limπ‘˜β†’βˆž

    𝑑

    βˆ‘π‘—=1

    βˆ«β‹…

    0

    Dπ‘˜,π‘—π‘‹π‘ π‘‘π΄π‘˜,𝑗

    𝑠=

    𝑑

    βˆ‘π‘—=1

    βˆ«β‹…

    0

    πœ™π»,𝑗

    π‘’π‘‘π‘Š

    (𝑗)

    𝑒= ∫

    β‹…

    0

    πœƒπ»

    𝑒𝑑𝑆𝑒,

    𝐿𝐻= limπ‘˜β†’βˆž

    𝑝

    βˆ‘π‘—=𝑑+1

    βˆ«β‹…

    0

    Dπ‘˜,π‘—π‘‹π‘ π‘‘π΄π‘˜,𝑗

    𝑠

    (53)

    weakly in 𝐡2(F). In particular,

    limπ‘˜β†’βˆž

    Dπ‘˜,𝑗𝑋 = πœ™π»,𝑗, (54)

    weakly in 𝐿2(Q Γ— 𝐿𝑒𝑏) for each 𝑗 = 1, . . . , 𝑝.

    Proof. We divide the proof into two steps. Throughout thisproof 𝐢 is a generic constant which may defer from line toline.

    Step 1. In the sequel, let π‘œ,π‘˜(β‹…) and 𝑝,π‘˜(β‹…) be the optional andpredictable projections with respect to Fπ‘˜, respectively. See,for example, [31, 33] for further details. Let us consider theFπ‘˜-martingales given by

    π‘€π‘˜

    𝑑:=

    𝑝

    βˆ‘π‘—=1

    π‘€π‘˜,𝑗

    𝑑; 0 ≀ 𝑑 ≀ 𝑇, (55)

    where

    π‘€π‘˜,𝑗

    𝑑:= ∫

    𝑑

    0

    Dπ‘˜,π‘—π‘‹π‘ π‘‘π΄π‘˜,𝑗

    𝑠; 0 ≀ 𝑑 ≀ 𝑇, 𝑗 = 1, . . . , 𝑝. (56)

    We claim that supπ‘˜β‰₯1

    E[π‘€π‘˜,π‘€π‘˜]𝑇

    < ∞. By the verydefinition,

    {(𝑑, πœ”) ∈ [0, 𝑇] Γ— Ξ©; Ξ” [π΄π‘˜,𝑗, π΄π‘˜,𝑗]𝑑(πœ”) ΜΈ= 0}

    =

    ∞

    ⋃𝑛=1

    βŸ¦π‘‡π‘˜,𝑗

    𝑛, π‘‡π‘˜,𝑗

    π‘›βŸ§ .

    (57)

    Therefore, Jensen inequality yields

    E [π‘€π‘˜,π‘€π‘˜]𝑇= E

    𝑝

    βˆ‘π‘—=1

    βˆ«π‘‡

    0

    Dπ‘˜,𝑗𝑋

    𝑠

    2

    𝑑 [π΄π‘˜,𝑗, π΄π‘˜,𝑗]𝑠

    ≀

    𝑝

    βˆ‘π‘—=1

    E∞

    βˆ‘π‘›=1

    E [(Dπ‘˜,π‘—π‘‹π‘‡π‘˜,𝑗

    𝑛

    )2

    | Fπ‘˜

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    ]

    Γ— 2βˆ’2π‘˜

    l{π‘‡π‘˜,𝑗

    𝑛 ≀𝑇}=: 𝐽

    π‘˜.

    (58)

    We will write π½π‘˜ in a slightly different manner as follows. Inthe sequel, for each 𝑑 ∈ (0, 𝑇], we set πœπ‘˜,𝑗

    π‘‘βˆ’:= max{π‘‡π‘˜,𝑗

    𝑛; π‘‡π‘˜,𝑗𝑛

    ≀

    𝑑} and πœπ‘˜,𝑗𝑑+

    := min{π‘‡π‘˜,𝑗𝑛; π‘‡π‘˜,𝑗𝑛

    > 𝑑}. Then, we will write

    π½π‘˜= E

    {

    {

    {

    𝑝

    βˆ‘π‘—=1

    ∞

    βˆ‘π‘›=1

    E [(Dπ‘˜,π‘—π‘‹π‘‡π‘˜,𝑗

    𝑛

    )2

    | Fπ‘˜

    π‘‡π‘˜,𝑗

    π‘›βˆ’1

    ] 2βˆ’2π‘˜

    1{π‘‡π‘˜,𝑗

    π‘›βˆ’1≀𝑇}

    βˆ’

    𝑝

    βˆ‘π‘—=1

    E [(Dπ‘˜,π‘—π‘‹πœπ‘˜,𝑗

    𝑇+

    )2

    | Fπ‘˜

    πœπ‘˜,𝑗

    π‘‡βˆ’

    ] 2βˆ’2π‘˜

    1{πœπ‘˜,𝑗

    π‘‡βˆ’β‰€π‘‡

  • International Journal of Stochastic Analysis 9

    In order to prove (61), let us check that

    limπ‘˜β†’βˆž

    Eβˆ«π‘‘

    0

    𝑝,π‘˜(𝑔)𝑠(Dπ‘˜,𝑗𝑋

    π‘ βˆ’ D

    π‘˜,𝑗𝑋𝑠) 𝑑 [𝐴

    π‘˜,𝑗, π΄π‘˜,𝑗]𝑠= 0,

    (63)

    limπ‘˜β†’βˆž

    Eβˆ«π‘‘

    0

    (π‘œ,π‘˜(𝑔)π‘ βˆ’π‘,π‘˜(𝑔)𝑠) (Dπ‘˜,𝑗𝑋

    π‘ βˆ’ D

    π‘˜,𝑗𝑋𝑠)

    Γ— 𝑑 [π΄π‘˜,𝑗, π΄π‘˜,𝑗]𝑠= 0.

    (64)

    The same trick we did in (59) together with (57) yields

    Eβˆ«π‘‘

    0

    𝑝,π‘˜(𝑔)𝑠(Dπ‘˜,𝑗𝑋

    π‘ βˆ’ D

    π‘˜,𝑗𝑋𝑠) 𝑑 [𝐴

    π‘˜,𝑗, π΄π‘˜,𝑗]𝑠

    = E [𝑝,π‘˜(𝑔)πœπ‘˜,𝑗

    𝑑+

    Dπ‘˜,π‘—π‘‹πœπ‘˜,𝑗

    𝑑+

    ] 2βˆ’2π‘˜

    1{πœπ‘˜,𝑗

    π‘‘βˆ’β‰€π‘‘

  • 10 International Journal of Stochastic Analysis

    Corollary 12. For a given Q ∈ M𝑒, let 𝐻 be a Q-square-integrable claim satisfying (𝑀). Let

    𝐻 = E [𝐻] + βˆ«π‘‡

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑+ 𝐿𝐻

    𝑇(72)

    be the correspondent GKW decomposition under Q. If 𝑑P/𝑑Q ∈ 𝐿1(P) and

    EP sup0≀𝑑≀𝑇

    βˆ«π‘‘

    0

    πœƒπ»

    𝑒𝑑𝑆𝑒

    < ∞, (73)

    then∞

    βˆ‘π‘›=1

    πœƒπ‘˜,𝐻

    π‘‡π‘˜,1

    π‘›βˆ’1

    (π‘†π‘‡π‘˜,1𝑛

    βˆ’ π‘†π‘‡π‘˜,1

    π‘›βˆ’1

    ) 1{π‘‡π‘˜,1𝑛 ≀⋅}

    β†’ βˆ«β‹…

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑

    π‘Žπ‘  π‘˜ β†’ ∞,

    (74)

    in the 𝜎(𝐡1, Ξ›βˆž)-topology under P.

    Proof. We have E|𝑑P/𝑑Q|2 = EP|𝑑P/𝑑Q|2(𝑑Q/𝑑P) = EP(𝑑P/𝑑Q) < ∞. To shorten notation, let π‘Œπ‘˜

    𝑑:= ∫

    𝑑

    0Dπ‘˜,1π‘ π‘‹π‘‘π΄π‘˜,1

    𝑠

    and π‘Œπ‘‘:= ∫

    𝑑

    0πœƒπ»β„“π‘‘π‘†β„“for 0 ≀ 𝑑 ≀ 𝑇. Let 𝐺 be an arbitrary

    F-stopping time bounded by 𝑇 and let 𝑔 ∈ 𝐿∞(P) be anessentiallyP-bounded random variable andF

    𝐺-measurable.

    Let 𝐽 ∈ 𝑀2 be a continuous linear functional given by thepurely discontinuous F-optional bounded variation process

    𝐽𝑑:= 𝑔E [

    𝑑P

    𝑑Q| F

    𝐺] 1{𝐺≀𝑑}

    ; 0 ≀ 𝑑 ≀ 𝑇, (75)

    where the duality action (β‹…, β‹…) is given by (𝐽,𝑁) = Eβˆ«π‘‡0𝑁𝑠𝑑𝐽𝑠,

    𝑁 ∈ 𝐡2(F). See Section 3.1 in [27] for more details. ThenTheorem 11 and the fact that 𝑑P/𝑑Q ∈ 𝐿2(Q) yield

    EPπ‘”π‘Œπ‘˜

    𝐺= Eπ‘Œ

    π‘˜

    𝐺𝑔𝑑P

    𝑑Q= (𝐽, π‘Œ

    π‘˜) β†’ (𝐽, π‘Œ)

    = Eπ‘ŒπΊπ‘”π‘‘P

    𝑑Q= EPπ‘”π‘ŒπΊ

    (76)

    as π‘˜ β†’ ∞. By the very definition,

    βˆ«π‘‘

    0

    Dπ‘˜,1π‘‹π‘ π‘‘π΄π‘˜,1

    𝑠

    =

    ∞

    βˆ‘π‘›=1

    E [Dπ‘˜,1π‘‹π‘‡π‘˜,1𝑛

    | Fπ‘˜

    π‘‡π‘˜,1

    π‘›βˆ’1

    ] Ξ”π΄π‘˜,1

    π‘‡π‘˜,1𝑛

    1{π‘‡π‘˜,1𝑛 ≀𝑑}

    =

    ∞

    βˆ‘π‘›=1

    πœƒπ‘˜,𝐻

    π‘‡π‘˜,1

    π‘›βˆ’1

    πœŽπ‘‡π‘˜,1

    π‘›βˆ’1

    π‘†π‘‡π‘˜,1

    π‘›βˆ’1

    (π‘Š(1)

    π‘‡π‘˜,1𝑛

    βˆ’π‘Š(1)

    π‘‡π‘˜,1

    π‘›βˆ’1

    ) 1{π‘‡π‘˜,1𝑛 ≀𝑑}

    =

    ∞

    βˆ‘π‘›=1

    πœƒπ‘˜,𝐻

    π‘‡π‘˜,1

    π‘›βˆ’1

    (π‘†π‘‡π‘˜,1𝑛

    βˆ’ π‘†π‘‡π‘˜,1

    π‘›βˆ’1

    ) 1{π‘‡π‘˜,1𝑛 ≀𝑑}

    ; 0 ≀ 𝑑 ≀ 𝑇.

    (77)

    Then from the definition of the 𝜎(𝐡1, Ξ›βˆž)-topology based onthe physical measure P, we will conclude the proof.

    Remark 13. Corollary 12 provides a nonantecipative Riem-man-sum approximation for the gain process βˆ«β‹…

    0πœƒπ»π‘‘π‘‘π‘†π‘‘in a

    multidimensional filtration setting where no path regularityof the pure hedging strategy πœƒπ» is imposed. The price we payis a weak-type convergence instead of uniform convergencein probability. However, from the financial point of view thistype of convergence is sufficient for the implementation ofMonte Carlo methods in hedging. More importantly, we willsee that πœƒπ‘˜,𝐻 can be fairly simulated and hence the resultingMonte Carlo hedging strategy can be calibrated from marketdata.

    Remark 14. If one is interested only in convergence at theterminal time 0 < 𝑇 < ∞, then assumption (73) can beweakened to EP| ∫

    𝑇

    0πœƒπ»

    𝑑𝑑𝑆𝑑| < ∞. Assumption EP(𝑑P/𝑑Q) <

    ∞ is essential to change theQ-convergence into the physicalmeasure P. One should notice that the associated densityprocess is no longer a P-local-martingale and in generalsuch integrability assumption must be checked case by case.Such assumption holds locally for every underlying ItoΜ‚ riskyasset price process. Our numerical results suggest that thisproperty behaves well for a variety of spot price models.

    Of course, in practice both the spot prices and the tradingdates are not observable at the stopping times so we need totranslate our results to a given deterministic set of rebalancinghedging dates.

    4.1. Hedging Strategies. In this section, we provide a dynamichedging strategy based on a refined set of hedging datesΞ  :=0 = 𝑠

    0< β‹… β‹… β‹… < 𝑠

    π‘žβˆ’1< π‘ π‘ž= 𝑇 for a fixed integer π‘ž. For this,

    we need to introduce some objects. For a given π‘ π‘–βˆˆ Ξ , we set

    π‘Š(𝑗)

    𝑠𝑖 ,𝑑:= π‘Š

    (𝑗)

    𝑠𝑖+π‘‘βˆ’ π‘Š(𝑗)

    𝑠𝑖, 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖, for 𝑗 = 1, 2. Of course,

    by the strong Markov property of the Brownian motion, weknow thatπ‘Š(𝑗)

    𝑠𝑖 ,β‹…is an (F𝑗

    𝑠𝑖 ,𝑑)0β‰€π‘‘β‰€π‘‡βˆ’π‘ π‘–

    -Brownianmotion for each𝑗 = 1, 2 and is independent ofF𝑗

    𝑠𝑖, whereF𝑗

    𝑠𝑖 ,𝑑:= F

    𝑗

    𝑠𝑖+𝑑for

    0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠𝑖. Similar to Section 3.1, we set π‘‡π‘˜,1

    𝑠𝑖 ,0:= 0 and

    π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛:= inf {𝑑 > π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βˆ’1;π‘Š(𝑗)

    𝑠𝑖 ,π‘‘βˆ’π‘Š

    (𝑗)

    𝑠𝑖 ,π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βˆ’1

    = 2βˆ’π‘˜} ;

    𝑛 β‰₯ 1, 𝑗 = 1, 2.

    (78)

    For a given π‘˜ β‰₯ 1 and 𝑗 = 1, 2, we defineHπ‘˜,𝑗𝑠𝑖 ,𝑛as the sigma-

    algebra generated by {π‘‡π‘˜,𝑗𝑠𝑖 ,β„“; 1 ≀ β„“ ≀ 𝑛} andπ‘Š(𝑗)

    𝑠𝑖 ,π‘‡π‘˜,𝑗

    𝑠𝑖 ,β„“

    βˆ’π‘Š(𝑗)

    𝑠𝑖 ,π‘‡π‘˜,𝑗

    𝑠𝑖 ,β„“βˆ’1

    ,

    1 ≀ β„“ ≀ 𝑛. We then define the following discrete jumpingfiltration:

    Fπ‘˜,𝑗

    𝑠𝑖 ,𝑑:= H

    π‘˜,𝑗

    𝑠𝑖 ,𝑛a.s on {π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛≀ 𝑑 < 𝑇

    π‘˜,𝑗

    𝑠𝑖 ,𝑛+1} . (79)

    In order to deal with fully path-dependent options, it isconvenient to introduce the following augmented filtration:

    Gπ‘˜,𝑗

    𝑠𝑖 ,𝑑:= F

    𝑗

    π‘ π‘–βˆ¨F

    π‘˜,𝑗

    𝑠𝑖 ,𝑑; 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖, (80)

    for 𝑗 = 1, 2. The bidimensional information flows are definedbyF

    𝑠𝑖 ,𝑑:= F1

    𝑠𝑖 ,π‘‘βŠ—F2

    𝑠𝑖 ,𝑑and Gπ‘˜

    𝑠𝑖 ,𝑑:= Gπ‘˜,1

    𝑠𝑖 ,π‘‘βŠ— Gπ‘˜,2

    𝑠𝑖 ,𝑑for 0 ≀ 𝑑 ≀

  • International Journal of Stochastic Analysis 11

    𝑇 βˆ’ 𝑠𝑖. We set Gπ‘˜

    𝑠𝑖:= {Gπ‘˜

    𝑠𝑖 ,𝑑; 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖}. We will assume

    that they satisfy the usual conditions. The piecewise constantmartingale projection π΄π‘˜,𝑗

    𝑠𝑖based onπ‘Š(𝑗)

    𝑠𝑖is given by

    π΄π‘˜,𝑗

    𝑠𝑖 ,𝑑:= E [π‘Š

    (𝑗)

    𝑠𝑖 ,π‘‡βˆ’π‘ π‘–| Gπ‘˜,𝑗

    𝑠𝑖 ,𝑑] ; 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖. (81)

    We set {π‘‡π‘˜π‘ π‘– ,𝑛; 𝑛 β‰₯ 0} as the order statistic generated by the

    stopping times {π‘‡π‘˜,𝑗𝑠𝑖 ,𝑛; 𝑗 = 1, 2, 𝑛 β‰₯ 0} similar to (29).

    If 𝐻 ∈ 𝐿2(Q) and 𝑋𝑑= E[𝐻 | F

    𝑑], 0 ≀ 𝑑 ≀ 𝑇, then we

    define

    π›Ώπ‘˜

    𝑠𝑖𝑋𝑑:= E [𝐻 | G

    π‘˜

    𝑠𝑖 ,𝑑] ; 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖, (82)

    so that the related derivative operators are given by

    Dπ‘˜,𝑗

    𝑠𝑖𝑋 :=

    ∞

    βˆ‘π‘›=1

    D𝑗

    π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛

    π›Ώπ‘˜

    𝑠𝑖𝑋1

    βŸ¦π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛,π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛+1⟦, (83)

    where

    Dπ‘—π›Ώπ‘˜

    𝑠𝑖𝑋 :=

    ∞

    βˆ‘π‘›=1

    Ξ”π›Ώπ‘˜π‘ π‘–π‘‹π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛

    Ξ”π΄π‘˜,𝑗

    π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛

    1βŸ¦π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛,π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βŸ§; 𝑗 = 1, 2, π‘˜ β‰₯ 1. (84)

    A Gπ‘˜π‘ π‘–-predictable version of Dπ‘˜,𝑗

    𝑠𝑖𝑋 is given by

    Dπ‘˜,𝑗𝑠𝑖𝑋 := 01⟦0⟧ +

    ∞

    βˆ‘π‘›=1

    E [Dπ‘˜,𝑗

    π‘ π‘–π‘‹π‘‡π‘˜,𝑗

    𝑠𝑖 ,𝑛

    | Gπ‘˜

    𝑠𝑖 ,π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βˆ’1

    ] 1βŸ§π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βˆ’1,π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βŸ§;

    𝑗 = 1, 2.

    (85)

    In the sequel, we denote

    πœƒπ‘˜,𝐻

    𝑠𝑖:=

    ∞

    βˆ‘π‘›=1

    Dπ‘˜,1π‘ π‘–π‘‹π‘‡π‘˜,1𝑠𝑖 ,𝑛

    πœŽπ‘ π‘– ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    𝑆𝑠𝑖 ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    1βŸ¦π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1,π‘‡π‘˜,𝑗

    𝑠𝑖 ,π‘›βŸ¦; 𝑠

    π‘–βˆˆ Ξ , (86)

    where πœŽπ‘ π‘– ,β‹…is the volatility process driven by the shifted

    filtration {F𝑠𝑖 ,𝑑; 0 ≀ 𝑑 ≀ 𝑇 βˆ’ 𝑠

    𝑖} and 𝑆

    𝑠𝑖 ,β‹…is the risky asset

    price process driven by the shifted Brownian motionπ‘Š(1)𝑠𝑖.

    We are now able to present the main result of this section.

    Corollary 15. For a given Q ∈ M𝑒, let 𝐻 be a Q-square-integrable claim satisfying (𝑀). Let

    𝐻 = E [𝐻] + βˆ«π‘‡

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑+ 𝐿𝐻

    𝑇(87)

    be the correspondent GKW decomposition under Q. If𝑑P/𝑑Q ∈ 𝐿1(P) and

    EP

    βˆ«π‘‡

    0

    πœƒπ»

    𝑒𝑑𝑆𝑒

    < ∞. (88)

    Then, for any set of trading dates Ξ  = {(𝑠𝑖)π‘ž

    𝑖=0}, we have

    limπ‘˜β†’βˆž

    βˆ‘π‘ π‘–βˆˆΞ 

    ∞

    βˆ‘π‘›=1

    πœƒπ‘˜,𝐻

    𝑠𝑖 ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    (𝑆𝑠𝑖 ,π‘‡π‘˜,1𝑠𝑖 ,𝑛

    βˆ’ 𝑆𝑠𝑖 ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    ) 1{π‘‡π‘˜,1𝑠𝑖 ,𝑛≀𝑠𝑖+1βˆ’π‘ π‘–}

    = βˆ«π‘‡

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑

    (89)

    weakly in 𝐿1 under P.

    Proof. Let Ξ  = {(𝑠𝑖)π‘ž

    𝑖=0} be any set of trading dates where π‘ž is

    a fixed positive integer. To shorten notation, let us define

    𝑅 (πœƒπ‘˜,𝐻

    , Ξ , π‘˜)

    := βˆ‘π‘ π‘–βˆˆΞ 

    ∞

    βˆ‘π‘›=1

    πœƒπ‘˜,𝐻

    𝑠𝑖 ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    (𝑆𝑠𝑖 ,π‘‡π‘˜,1𝑠𝑖 ,𝑛

    βˆ’ 𝑆𝑠𝑖 ,π‘‡π‘˜,1

    𝑠𝑖 ,π‘›βˆ’1

    ) 1{π‘‡π‘˜,1𝑠𝑖 ,𝑛≀𝑠𝑖+1βˆ’π‘ π‘–}

    (90)

    for π‘˜ β‰₯ 1 and Ξ . At first, we recall that {π‘‡π‘˜,1𝑠𝑖 ,𝑛

    βˆ’ π‘‡π‘˜,1𝑠𝑖 ,π‘›βˆ’1

    ; 𝑛 β‰₯

    1, π‘ π‘–βˆˆ Ξ } is an i.i.d sequence with absolutely continuous

    distribution. In this one-dimensional case, the probability ofthe set {π‘‡π‘˜,1

    𝑠𝑖 ,𝑛≀ 𝑠𝑖+1

    βˆ’ 𝑠𝑖} is always strictly positive for every Ξ 

    and π‘˜, 𝑛 β‰₯ 1. Hence, 𝑅(πœƒπ‘˜,𝐻, Ξ , π‘˜) is a nondegenerate subsetof random variables. By making a change of variable on theItoΜ‚ integral, we will write

    βˆ«π‘‡

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑= ∫

    𝑇

    0

    πœ™π»,1

    π‘‘π‘‘π‘Š

    (1)

    𝑑= βˆ‘π‘ π‘–βˆˆΞ 

    βˆ«π‘ π‘–+1

    𝑠𝑖

    πœ™π»,1

    π‘‘π‘‘π‘Š

    (1)

    𝑑

    = βˆ‘π‘ π‘–βˆˆΞ 

    βˆ«π‘ π‘–+1βˆ’π‘ π‘–

    0

    πœ™π»,1

    𝑠𝑖+π‘‘π‘‘π‘Š

    (1)

    𝑠𝑖 ,𝑑.

    (91)

    Let us fixQ ∈ M𝑒. By the very definition,

    𝑅 (πœƒπ‘˜,𝐻

    , Ξ , π‘˜) = βˆ‘π‘ π‘–βˆˆΞ 

    βˆ«π‘ π‘–+1βˆ’π‘ π‘–

    0

    Dπ‘˜,1π‘ π‘–π‘‹β„“π‘‘π΄π‘˜,1

    𝑠𝑖 ,β„“under Q. (92)

    Now we notice that Theorem 11 holds for the two-dimensional Brownian motion (π‘Š(1)

    𝑠𝑖,π‘Š(2)𝑠𝑖), for each

    π‘ π‘–βˆˆ Ξ  with the discretization of the Brownian motion given

    by π΄π‘˜,1𝑠𝑖. Moreover, using the fact that E|𝑑P/𝑑Q|2 < ∞

    and repeating the argument given by (77) restricted to theinterval [𝑠

    𝑖, 𝑠𝑖+1), we have

    limπ‘˜β†’βˆž

    𝑅 (πœƒπ‘˜,𝐻

    , Ξ , π‘˜) = βˆ‘π‘ π‘–βˆˆΞ 

    limπ‘˜β†’βˆž

    βˆ«π‘ π‘–+1βˆ’π‘ π‘–

    0

    Dπ‘˜,1π‘ π‘–π‘‹β„“π‘‘π΄π‘˜,1

    𝑠𝑖 ,β„“

    = βˆ«π‘‡

    0

    πœƒπ»

    𝑑𝑑𝑆𝑑,

    (93)

    weakly in 𝐿1(P) for each Π. This concludes the proof.

    Remark 16. In practice, one may approximate the gain pro-cess by a nonantecipative strategy as follows. Let Ξ  be agiven set of trading dates on the interval [0, 𝑇] so that |Ξ | =max

    0β‰€π‘–β‰€π‘ž|π‘ π‘–βˆ’ π‘ π‘–βˆ’1| is small. We take a large π‘˜ and we perform

    a nonantecipative buy-and-hold-type strategy among thetrading dates [𝑠

    𝑖, 𝑠𝑖+1); π‘ π‘–βˆˆ Ξ , in the full approximation (90)

    which results in

    βˆ‘π‘ π‘–βˆˆΞ 

    πœƒπ‘˜,𝐻

    𝑠𝑖 ,0(𝑆𝑠𝑖 ,𝑠𝑖+1βˆ’π‘ π‘–

    βˆ’ 𝑆𝑠𝑖 ,0) where

    πœƒπ‘˜,𝐻

    𝑠𝑖 ,0=

    E [Dπ‘˜,1π‘ π‘–π‘‹π‘‡π‘˜,1

    𝑠𝑖 ,1

    | F𝑠𝑖]

    πœŽπ‘ π‘– ,0𝑆𝑠𝑖 ,0

    ; π‘ π‘–βˆˆ Ξ .

    (94)

    Convergence (89) implies that approximation (94) results inunavoidable hedging errors with respect to the gain process

  • 12 International Journal of Stochastic Analysis

    due to the discretization of the dynamic hedging, but we donot expect large hedging errors provided that π‘˜ is large and|Ξ | is small. Hedging errors arising from discrete hedgingin complete markets are widely studied in the literature. Wedo not know optimal rebalancing dates in this incompletemarket setting, but simulation results presented in Section 6suggest that homogeneous hedging dates work very well fora variety of models with and without stochastic volatility. Amore detailed study is needed in order to get more preciserelations betweenΞ  and the stopping times, a topicwhichwillbe further explored in a future work.

    Let us now briefly explain how the results of this sectioncan be applied to well-known quadratic hedging methodolo-gies.

    Generalized Föllmer-Schweizer. If one takes the minimalmartingalemeasure PΜ‚, then 𝐿𝐻 in (70) is aP-localmartingaleand is orthogonal to the martingale component of 𝑆. Duethis orthogonality and the zero mean behavior of the cost𝐿𝐻, it is still reasonable to work with generalized Föllmer-Schweizer decompositions underPwithout knowing a priorithe existence of locally risk minimizing hedging strategies.

    Local Risk Minimization. One should notice that if ∫ πœƒπ»π‘‘π‘† ∈𝐡2(F), 𝐿𝐻 ∈ 𝐡2(F) under P and 𝑑PΜ‚/𝑑P ∈ 𝐿2(P), then πœƒπ»is the locally risk minimizing trading strategy and (70) is theFöllmer-Schweizer decomposition under P.

    Mean Variance Hedging. If one takes PΜƒ, then the meanvariance hedging strategy is not completely determined bytheGKWdecomposition under PΜƒ. Nevertheless, Corollary 15still can be used to approximate the optimal hedging strategyby computing the density process 𝑍 based on the so-calledfundamental equations derived by Hobson [16]. See (5)and (6) for details. For instance, in the classical Hestonmodel, Hobson derives analytical formulas for 𝜁. See (110) inSection 6.

    Hedging of Fully Path-Dependent Options. The most interest-ing application of our results is the hedging of fully path-dependent options under stochastic volatility. For instance,if 𝐻 = Ξ¦({𝑆

    𝑑; 0 ≀ 𝑑 ≀ 𝑇}), then Corollary 15 and

    Remark 16 jointly with the above hedging methodologiesallow us to dynamically hedge the payoff 𝐻 based on (94).The conditioning on the information flow {F

    𝑠𝑖; π‘ π‘–βˆˆ Ξ } in

    the hedging strategy πœƒπ‘˜,𝐻hedg := {πœƒπ‘˜,𝐻

    𝑠𝑖; π‘ π‘–βˆˆ Ξ } encodes the

    continuous monitoring of a path-dependent option. For eachhedging date 𝑠

    𝑖, one has to incorporate the whole history

    of the price and volatility until such date in order to getan accurate description of the hedging. If 𝐻 is not path-dependent, then the information encoded by {F

    𝑠𝑖; π‘ π‘–βˆˆ Ξ }

    in πœƒπ‘˜,𝐻hedg is only crucial at time 𝑠𝑖.Next, we provide the details of the Monte Carlo algo-

    rithm for the approximating pure hedging strategy πœƒπ‘˜,𝐻hedg ={πœƒπ‘˜,𝐻𝑠𝑖 ,0

    ; π‘ π‘–βˆˆ Ξ }.

    5. The Algorithm

    In this section we present the basic algorithm to evaluate thehedging strategy for a given European-type contingent claim𝐻 ∈ 𝐿2(Q) satisfying assumption (M) for a fixed Q ∈ M𝑒 ata terminal time 0 < 𝑇 < ∞. The core of the algorithm is thesimulation of the stochastic derivativeDπ‘˜,𝑗𝑋

    E[[

    [

    E [𝐻 | Fπ‘˜π‘‡π‘˜,𝑗

    1

    ] βˆ’ E [𝐻 | Fπ‘˜πœπ‘˜,𝑗

    1

    ]

    π΄π‘˜,𝑗

    π‘‡π‘˜,𝑗

    1

    ]]

    ]

    , 𝑗 = 1, . . . , 𝑑, (95)

    for 𝑑 ≀ 𝑝. Recall that Fπ‘˜ is a discrete jumping filtrationgenerated by the i.i.d families of Bernoulli and absolutelycontinuous random variables given, respectively, by {πœ‚π‘˜,𝑗

    𝑛; 𝑛 β‰₯

    1, 1 ≀ 𝑗 ≀ 𝑝} and {π‘‡π‘˜,𝑗𝑛

    βˆ’ π‘‡π‘˜,𝑗

    π‘›βˆ’1; 𝑛 β‰₯ 1, 1 ≀ 𝑗 ≀ 𝑝} which are

    amenable to an exact simulation by using Burq and Jones [8].By considering the payoff𝐻 as a functional of π΄π‘˜,1, . . . , π΄π‘˜,𝑝,this section explains how to perform a concrete and feasibleMonte Carlo method to obtain the hedging strategies πœƒπ».

    In the sequel, we fix the discretization level π‘˜ β‰₯ 1.

    Step 1 (simulation of the stopping times {π‘‡π‘˜,𝑗ℓ; 𝑗 = 1, . . . , 𝑝;

    β„“ β‰₯ 1} and the step processes {π΄π‘˜,𝑗; 𝑗 = 1, . . . , 𝑝})

    (1) One generates the increments {π‘‡π‘˜,𝑗ℓ

    βˆ’ π‘‡π‘˜,𝑗

    β„“βˆ’1; β„“ β‰₯ 1}

    according to the algorithm described by Burq andJones [8] and, consequently, the Fπ‘˜,𝑗-stopping times{π‘‡π‘˜,𝑗

    β„“; β„“ β‰₯ 1} for every 𝑗 = 1, . . . , 𝑝, such that all the

    Fπ‘˜,𝑗-stopping times π‘‡π‘˜,𝑗ℓ

    ≀ 𝑇.

    (2) One simulates the i.i.d family πœ‚π‘˜,𝑗 = {πœ‚π‘˜,𝑗ℓ; β„“ β‰₯ 1}

    independently of {π‘‡π‘˜,𝑗ℓ; β„“ β‰₯ 1}, according to the

    Bernoulli random variable πœ‚π‘˜,𝑗1with parameter 1/2 for

    𝑖 = βˆ’1, 1. This simulates the step process π΄π‘˜,𝑗 for𝑗 = 1, . . . , 𝑝.

    In the next step we need to simulate 𝐻 based onapproximations of the discounted price process {𝑆𝑖

    𝑑; 0 ≀ 𝑑 ≀

    𝑇; 𝑖 = 1, . . . , 𝑑} as follows.

    Step 2 (simulation of the discounted stock price process{𝑆𝑖; 𝑖 = 1, . . . , 𝑑}). Suppose that, using Step 1, we have the

    partitionsTπ‘˜,𝑗, the family πœ‚π‘˜,𝑗, and the step processesπ΄π‘˜,𝑗 for𝑗 = 1, . . . , 𝑑, 𝑑 + 1, . . . , 𝑝. The following steps show how tocompute approximations to the discounted stock price prices𝑆𝑖, 𝑖 = 1, . . . , 𝑑, and the payoff function𝐻.

    (1) We consider the order statistics Tπ‘˜ generated by allstopping times as defined by (29). This is the finestpartition generated by all partitionsTπ‘˜,𝑗.

    (2) We apply some appropriate method to evaluate anapproximation π‘†π‘˜,𝑖 of the discounted price 𝑆𝑖 for 𝑖 =1, . . . , 𝑑, where π‘†π‘˜,𝑖 is a functional of the noisy π΄π‘˜ =(π΄π‘˜,1, . . . , π΄π‘˜,𝑝). Generally speaking, we work withsome ItoΜ‚-Taylor expansion method driven by π΄π‘˜.

  • International Journal of Stochastic Analysis 13

    (3) Based on the approximation for π‘†π‘˜, we calculate theapproximation for the payoff οΏ½Μ‚οΏ½ as follows: οΏ½Μ‚οΏ½π‘˜ =Ξ¦(π‘†π‘˜,𝑗; 1 ≀ 𝑗 ≀ 𝑝).

    Next, we describe the crucial step in the algorithm: thesimulation of the stochastic derivative described by (41).

    Step 3 (simulation of the stochastic derivative Dπ‘˜,π‘—π‘‹π‘‡π‘˜,𝑗

    1

    ).Werecall that

    Dπ‘˜,π‘—π‘‹π‘‡π‘˜,𝑗

    1

    = E[[

    [

    E [𝐻 | Fπ‘˜π‘‡π‘˜,𝑗

    1

    ] βˆ’ E [𝐻 | Fπ‘˜πœπ‘˜,𝑗

    1

    ]

    π΄π‘˜,𝑗

    π‘‡π‘˜,𝑗

    1

    ]]

    ]

    , (96)

    where πœπ‘˜,𝑗1

    = max{π‘‡π‘˜π‘›; π‘‡π‘˜π‘›< 𝑇

    π‘˜,𝑗

    1}, andFπ‘˜

    π‘‡π‘˜,𝑗

    1

    is given by

    Fπ‘˜

    π‘‡π‘˜,𝑗

    1

    = Fπ‘˜

    πœπ‘˜,𝑗

    1

    ∨ 𝜎 (π‘‡π‘˜,𝑗

    1, πœ‚π‘˜,𝑗

    1) . (97)

    In the sequel, π‘‘π‘˜,𝑗ℓdenotes the realization of π‘‡π‘˜,𝑗

    β„“by means

    of Step 1 and π‘‘π‘˜β„“denotes the realization of π‘‡π‘˜

    β„“based on the

    finest random partition Tπ‘˜. Moreover, any sequence (π‘‘π‘˜1<

    π‘‘π‘˜2

    < β‹… β‹… β‹… < π‘‘π‘˜,𝑗

    1) encodes the information generated by

    the realization of Tπ‘˜ until the first hitting time of the 𝑗thpartition. In addition, we denote π‘‘π‘˜,𝑗

    1βˆ’as the last time in the

    finest partition before π‘‘π‘˜,𝑗1. For each (π‘˜, 𝑛) ∈ N Γ— N, let

    (]𝑛1,π‘˜, ]𝑛2,π‘˜) ∈ {1, . . . , 𝑝} Γ— N be the unique random pair which

    realizes

    π‘‡π‘˜

    𝑛= 𝑇

    π‘˜,]𝑛1,π‘˜

    ]𝑛2,π‘˜

    a.s, π‘˜, 𝑛 β‰₯ 1. (98)

    Based on these quantities, we define πœ‚π‘˜π‘‘π‘˜π‘›

    as the realization of

    the random variable πœ‚π‘˜,]𝑛

    1,π‘˜

    ]𝑛2,π‘˜

    , where {πœ‚π‘˜,π‘—π‘š; 1 ≀ 𝑗 ≀ 𝑝, π‘š β‰₯ 1} is

    given by (26).In the sequel, EΜ‚ denotes the conditional expectation

    computed in terms of the Monte Carlo method.

    (1) For every 𝑗 = 1, . . . , 𝑑 we compute

    DΜ‚π‘˜,π‘—π‘‹π‘‡π‘˜,𝑗

    1

    :=1

    2βˆ’π‘˜πœ‚π‘˜,𝑗

    1

    {EΜ‚ [𝐻 | Fπ‘˜

    π‘‡π‘˜,𝑗

    1

    ] βˆ’ EΜ‚ [𝐻 | Fπ‘˜

    πœπ‘˜,𝑗

    1

    ]}

    =1

    2βˆ’π‘˜πœ‚π‘˜,𝑗

    1

    {EΜ‚ [𝐻 | (π‘‘π‘˜

    1, πœ‚π‘˜

    π‘‘π‘˜

    1

    ) , . . . , (π‘‘π‘˜,𝑗

    1, πœ‚π‘˜

    π‘‘π‘˜,𝑗

    1

    )]

    βˆ’ EΜ‚ [𝐻 | (π‘‘π‘˜

    1, πœ‚π‘˜

    π‘‘π‘˜

    1

    ) , . . . , (π‘‘π‘˜,𝑗

    1βˆ’, πœ‚π‘˜

    π‘‘π‘˜,𝑗

    1βˆ’

    )]} ,

    (99)

    where πœ‚π‘˜,𝑗1in (99) denotes the realization of the Bernoulli

    variable πœ‚π‘˜,𝑗1.

    (2) We define the stochastic derivative

    πœ™π»,𝑆

    0:= (DΜ‚

    π‘˜,1π‘‹π‘‡π‘˜,1

    1

    , . . . , DΜ‚π‘˜,π‘‘π‘‹π‘‡π‘˜,𝑑

    1

    ) . (100)

    (3) We compute πœƒπ»

    0as

    πœƒπ»

    0:= (πœ™

    𝐻,𝑆

    0)⊀

    [diag(𝑆0)𝜎0]βˆ’1

    . (101)

    (4) Repeat these steps several times and calculate the purehedging strategy as the mean of all πœƒ

    𝐻

    0. Consider

    πœƒπ»

    0:= mean of πœƒ

    𝐻

    0. (102)

    Quantity (102) is a Monte Carlo estimate of πœƒπ»0.

    Remark 17. TheMonte Carlo simulation of (99) is performedby considering the payoff 𝐻 = Ξ¦(𝑆𝑗; 1 ≀ 𝑗 ≀ 𝑝) asa functional Ξ¦(π‘†π‘˜,𝑗; 1 ≀ 𝑗 ≀ 𝑝) of the noisy π΄π‘˜ =(π΄π‘˜,1, . . . , 𝐴

    π‘˜,𝑝) in terms of any ItoΜ‚-Taylor/Euler-Maruyama

    scheme.

    Remark 18. In order to compute the hedging strategy πœƒπ» overa trading period {𝑠

    𝑖; 𝑖 = 0, . . . , π‘ž}, one performs Algorithms 1,

    2 and 3 (see Appendix) but based on the shifted filtration andthe Brownian motionsπ‘Š(𝑗)

    𝑠𝑖for 𝑗 = 1, . . . , 𝑝 as described in

    Section 4.1.

    Remark 19. In practice, one has to calibrate the parametersof a given stochastic volatility model based on liquid instru-ments such as vanilla options and volatility surfaces. Withthose parameters at hand, the trader must follow steps (99)and (102). The hedging strategy is then given by calibrationand the computation of quantity (102) over a trading period.

    6. Numerical Analysis andDiscussion of the Methods

    In this section, we provide a detailed analysis of the numericalscheme proposed in this work.

    6.1. Multidimensional Black-Scholes Model. At first, we con-sider the classical multidimensional Black-Scholes modelwith asmany risky stocks as underlying independent randomfactors to be hedged (𝑑 = 𝑝). In this case, there is only oneequivalent local martingale measure, the hedging strategy πœƒπ»is given by (43), and the cost is just the option price. Toillustrate our method, we study a very special type of exoticoption: a BLAC (Basket Lock Active Coupon) down-and-outbarrier option whose payoff is given by

    𝐻 =βˆπ‘– ΜΈ=𝑗

    1{minπ‘ βˆˆ[0,𝑇]Sπ‘–π‘ βˆ¨minπ‘ βˆˆ[0,𝑇]𝑆

    𝑗

    𝑠>𝐿}. (103)

    It is well known that, for this type of option, there exists aclosed formula for the hedging strategy. Moreover, it satisfiesthe assumptions of Theorem A.2. See, for example, Bernis etal. [34] for some formulas.

    For comparison purposes with Bernis et al. [34], weconsider 𝑑 = 5 underlying assets, π‘Ÿ = 0% for the interest rate,and 𝑇 = 1 year for the maturity time. For each asset, we setinitial values 𝑆𝑖

    0= 100, 1 ≀ 𝑖 ≀ 5, andwe compute the hedging

  • 14 International Journal of Stochastic Analysis

    5000 10000 15000 20000

    BLAC option

    Hed

    ge

    00.000

    0.001

    0.002

    0.003

    0.004

    0.005

    0.006

    True value

    k = 3

    k = 4

    k = 5

    k = 6

    Figure 1: Monte Carlo hedging strategy of a BLAC down-and-outoption for a 5-dimensional Black-Scholes model.

    strategy with respect to the first asset 𝑆1 with discretizationlevel π‘˜ = 3, 4, 5, 6 and 20000 simulations.

    Following thework [34], we consider the volatilities of theassets given by β€–πœŽ1β€– = 35%, β€–πœŽ2β€– = 35%, β€–πœŽ3β€– = 38%, β€–πœŽ4β€– =35%, and β€–πœŽ5β€– = 40% and the correlation matrix defined byπœŒπ‘–π‘—= 0, 4 for 𝑖 ΜΈ= 𝑗, where πœŽπ‘– = (𝜎

    𝑖1, . . . , 𝜎

    𝑖5)⊀, and we use

    the barrier level 𝐿 = 76. In Table 1, we present the numericalresults based on the Algorithms 1, 2 and 3 for the pointwisehedging strategy πœƒπ» at time 𝑑 = 0.

    Table 1 reports the difference between the true andthe estimated hedging value, the standard error =standard deviation/(number of simulations)1/2, the %error = difference/true valor, and the lower (LL) andupper limits (UL) of the 95% confidence interval for theempirical mean of the estimated pointwise hedging strategiesat time 𝑑 = 0. Due to Theorem A.2, we expect that whenthe discretization level π‘˜ increases, we obtain results closerto the true value and this is what we find in Monte Carloexperiments, confirmed by the small % error when usingπ‘˜ = 6. We also emphasize that when π‘˜ = 6, the confidenceinterval contains the true value 0.00338, and we can reallyassume the convergence of the algorithm.

    In Figure 1, we plot the average hedging estimates withrespect to the number of simulations. One should notice thatwhen π‘˜ increases, the standard error also increases, whichsuggests more simulations for higher values of π‘˜.

    6.2. Average Hedging Errors. Next, we present some aver-age hedging error results for two well-known nonconstantvolatility models: the constant elasticity of variance (CEV)model and the classical Heston stochastic volatility model[18].The typical exampleswe have inmind are the generalizedFöllmer-Schweizer, local risk minimization, and mean vari-ance hedging strategies, where the optimal hedging strategiesare computed by means of the minimal martingale measureand the variance optimal martingale measure, respectively.

    We analyze the one-touch one-dimensional European-typecontingent claims as follows:

    One-touch option: 𝐻 = 1{maxπ‘‘βˆˆ[0,𝑇]𝑆𝑑>105}. (104)

    By using the Algorithms 1, 2 and 3, we compute theerror committed by approximating the payoff𝐻 by EΜ‚Q[𝐻] +βˆ‘π‘›βˆ’1

    𝑖=0πœƒπ‘˜,𝐻𝑑𝑖 ,0

    (𝑆𝑑𝑖 ,𝑑𝑖+1βˆ’π‘‘π‘–

    βˆ’ 𝑆𝑑𝑖 ,0). This error will be called hedging

    error. The computation of this error is summarized in thefollowing steps.

    Computation of the Average Hedging Error(1) We first simulate paths under the physical measureand compute the payoff𝐻.

    (2) Then, we consider some deterministic partition of theinterval [0, 𝑇] into 𝑛 (number of hedging strategies inthe period) points 𝑑

    0, 𝑑1, . . . , 𝑑

    π‘›βˆ’1such that 𝑑

    𝑖+1βˆ’ 𝑑𝑖=

    𝑇/𝑛, for 𝑖 = 0, . . . , 𝑛 βˆ’ 1.(3) One simulates, at time 𝑑

    0= 0, the option price EΜ‚Q[𝐻]

    and the initial hedging estimate πœƒπ‘˜,𝐻0,0

    through (100),(101), and (102) under a fixed Q ∈ M𝑒. We follow theAlgorithms 1, 2 and 3.

    (4) We simulate πœƒπ‘˜,𝐻𝑑𝑖 ,0

    by means of the shifting argumentbased on the strongMarkov property of the Brownianmotion as described in Section 4.1.

    (5) We compute οΏ½Μ‚οΏ½ by

    οΏ½Μ‚οΏ½ := EΜ‚Q [𝐻] +π‘›βˆ’1

    βˆ‘π‘–=0

    πœƒπ‘˜,𝐻

    𝑑𝑖 ,0(𝑆𝑑𝑖 ,𝑑𝑖+1βˆ’π‘‘π‘–

    βˆ’ 𝑆𝑑𝑖 ,0) . (105)

    (6) We compute the hedging error estimate 𝛾 given by𝛾 := |𝐻 βˆ’ οΏ½Μ‚οΏ½|.

    (7) We compute the average hedging error given by AV :=(1/𝑀)βˆ‘

    𝑀

    β„“=1𝛾ℓwhere 𝛾

    β„“is the hedging error at the β„“th

    scenario and𝑀 is the total number of scenarios usedin the experiment.

    (8) We compute 𝐸(AV) := 100 Γ— AV/EΜ‚Q[𝐻].

    Remark 20. When no locally risk minimizing strategy isavailable, we also expect to obtain low average hedging errorswhen dealing with generalized Föllmer-Schweizer decom-positions due to the orthogonal martingale decomposition.In the mean variance hedging case, two terms appear inthe optimal hedging strategy: the pure hedging componentπœƒπ»,PΜƒ of the GKW decomposition under the optimal variancemartingale measures PΜƒ and 𝜁 as described by (5) and (6).For the Heston model, 𝜁 was explicitly calculated by Hobson[16]. We have used his formula in our numerical simulationsjointly with πœƒπ‘˜,𝐻 under PΜƒ in the calculation of the meanvariance hedging errors. See expression (110) for details.

    6.2.1. Constant Elasticity of Variance (CEV) Model. The dis-counted risky asset price process described by theCEVmodelunder the physical measure is given by

    𝑑𝑆𝑑= 𝑆𝑑[(π‘π‘‘βˆ’ π‘Ÿπ‘‘) 𝑑𝑑 + πœŽπ‘†

    (π›½βˆ’2)/2

    𝑑𝑑𝐡𝑑] , 𝑆

    0= 𝑠, (106)

  • International Journal of Stochastic Analysis 15

    Table 1: Monte Carlo hedging strategy of a BLAC down-and-out option for a 5-dimensional Black-Scholes model.

    π‘˜ Hedging St. error LL UL True value Difference % error3 0.00376 2.37 Γ— 10βˆ’5 0.00371 0.00380 0.00338 0.00038 11.15%4 0.00365 4.80 Γ— 10βˆ’5 0.00356 0.00374 0.00338 0.00027 8.03%5 0.00366 9.31 Γ— 10βˆ’5 0.00348 0.00384 0.00338 0.00028 8.35%6 0.00342 1.82 Γ— 10βˆ’4 0.00306 0.00378 0.00338 0.00004 1.29%

    where 𝐡 is a P-Brownian motion. The instantaneous Sharperatio is πœ“

    𝑑= (𝑏

    π‘‘βˆ’ π‘Ÿπ‘‘)/(πœŽπ‘†

    (π›½βˆ’2)/2

    𝑑) such that the model can be

    rewritten as

    𝑑𝑆𝑑= πœŽπ‘‘π‘†π›½/2

    π‘‘π‘‘π‘Š

    𝑑, (107)

    where π‘Š is a Q-Brownian motion and Q is the equivalentlocal martingale measure. In this Monte Carlo experiment,we consider a total number of scenarios 𝑀 equal to 1000with the following parameters: the barrier for the one-touchoption in (104) is 105, π‘Ÿ = 0 for the interest rate, 𝑏 = 0.01,𝑇 = 1 (month) for the maturity time, 𝜎 = 0.2, 𝑆

    0=

    100, and 𝛽 = 1.6 such that the constant of elasticity isβˆ’0.4. We simulate the average hedging errors by consideringdiscretization levels π‘˜ = 3, 4, 5. We perform 11, 16, 22, and44 hedging strategies along the interval [0, 𝑇]. We observethat, supposing 22 business days per month, we can assumethat 11, 22, and 44 hedging strategies on the interval [0, 1]correspond to one hedging strategy for every two days, onehedging strategy per day, and two hedging strategies per day,respectively. From Corollary 15, we know that this procedureis consistent.

    Table 2 reports the average hedging errors for the one-touch option. It provides the standard error = standard devi-ation of {𝛾

    𝑖; 1 ≀ 𝑖 ≀ 𝑀}/(total number of scenarios 𝑀)1/2,

    the % error = 𝐸(AV), the lower (LL) and upper limits (UL)of the 95% confidence interval for AV, and the price of theoption. It is important to notice that when π‘˜ increases, thepercentage error 𝐸(AV) decreases, which is expected due tothe weak convergence results of this paper. We also pointout that for π‘˜ = 5 all the 95% confidence intervals containthe zero. Moreover, we notice that as the number of hedgingstrategies increases, the standard error becomes smaller.

    6.2.2. Heston’s Stochastic Volatility Model. Here we considertwo types of hedging methodologies: local risk minimizationand mean variance hedging strategies as described in theIntroduction and Remark 20. The Heston dynamics of thediscounted price under the physical measure is given by

    𝑑𝑆𝑑= 𝑆𝑑(π‘π‘‘βˆ’ π‘Ÿπ‘‘) Σ𝑑𝑑𝑑 + 𝑆

    π‘‘βˆšΞ£π‘‘π‘‘π΅(1)

    𝑑,

    𝑑Σ𝑑= 2πœ… (πœƒ βˆ’ Ξ£

    𝑑) 𝑑𝑑 + 2𝜎√Σ

    𝑑𝑑𝑍𝑑, 0 ≀ 𝑑 ≀ 𝑇,

    (108)

    where 𝑍 = 𝜌𝐡(1) + 𝜌𝐡(2)𝑑, 𝜌 = √1 βˆ’ 𝜌2, (𝐡(1), 𝐡(2)) is a pair

    of two independent P-Brownian motions, and πœ…,π‘š, 𝛽0, πœ‡ are

    suitable constants in order to have a well-defined Markov

    process (see, e.g., [16, 18]). Alternatively, we can rewrite thedynamics as

    𝑑𝑆𝑑= π‘†π‘‘π‘Œ2

    𝑑(π‘π‘‘βˆ’ π‘Ÿπ‘‘) 𝑑𝑑 + 𝑆

    π‘‘π‘Œπ‘‘π‘‘π΅(1)

    𝑑,

    π‘‘π‘Œπ‘‘= πœ…(

    π‘š

    π‘Œπ‘‘

    βˆ’ π‘Œπ‘‘)𝑑𝑑 + πœŽπ‘‘π‘

    𝑑, 0 ≀ 𝑑 ≀ 𝑇,

    (109)

    where π‘Œ = βˆšΞ£π‘‘andπ‘š = πœƒ βˆ’ (𝜎2/2πœ…).

    Local Risk Minimization. For comparison purposes withHeath et al. [4], we consider the hedging of a Europeanput option 𝐻 written on a Heston model with correlationparameter 𝜌 = 0. We set 𝑆

    0= 100, strike price 𝐾 = 100, and

    𝑇 = 1 (month) andwe use discretization levels π‘˜ = 3, 4, and 5.We set the parameters πœ… = 2.5, πœƒ = 0.04, 𝜌 = 0, 𝜎 = 0.3, π‘Ÿ = 0,and π‘Œ

    0= 0.02. The hedging strategy πœƒπ»,PΜ‚ based on the local

    risk minimization methodology is bounded with continuouspaths so that Theorem A.2 applies to this case. Moreover, asdescribed by Heath et al. [4], πœƒπ»,PΜ‚ can be obtained by a PDEnumerical analysis.

    Table 3 presents the results of the hedging strategy πœƒπ‘˜,𝐻0,0

    by using Algorithms 1, 2 and 3. Figure 2 provides the MonteCarlo hedging strategy with respect to the number of simula-tions of order 10000. We notice that our results agree withthe results obtained by Heath et al. [4] by PDE methods.In this case, the true value of the hedging at time 𝑑 = 0is approximately βˆ’0.44. Table 3 provides the standard errorsrelated to the computed hedging strategy and the MonteCarlo prices.

    Hedging with Generalized Föllmer-Schweizer Decompositionfor One-Touch Option. Based on Corollary 15, we also presentthe averaging hedging error associated with one-touchoptions written on a Heston model with nonzero correlation.We consider a total number of scenarios 𝑀 = 1000 andwe set πœ… = 3.63, πœƒ = 0.04, 𝜌 = βˆ’0.53, 𝜎 = 0.3, π‘Ÿ = 0,𝑏 = 0.01, π‘Œ

    0= 0.3, and 𝑆

    0= 100 where the barrier is

    105. We simulate the average hedging error along the interval[0, 1] with discretization levels π‘˜ = 3, 4. We compute 22 and44 hedging strategies in the period (which corresponds toone and two hedging strategies per day, resp.). The averagehedging error results are summarized in Table 4. It providesthe standard error (St. error) = standard deviation of {𝛾

    β„“; 1 ≀

    β„“ ≀ 𝑀}/(total number of scenarios 𝑀 = 1000)1/2, the priceof the option, the lower (LL) and upper (UL) limits of the 95%confidence interval of AV, and the percentage error 𝐸(AV)related to AV.

    To the best of our knowledge, there is no result con-cerning the existence of locally risk minimizing hedging

  • 16 International Journal of Stochastic Analysis

    Table 2: Average hedging error of the one-touch option written on the CEV model.

    π‘˜ Hedging strategies AV St. error LL UL Price 𝐸(AV)3 11 0.0449 0.0073 0.0305 0.0592 0.4803 9.34%3 16 0.0446 0.0063 0.0323 0.0569 0.4804 9.28%3 22 0.0441 0.0056 0.0332 0.0550 0.4804 9.18%3 44 0.0431 0.0044 0.0345 0.0516 0.4803 8.96%4 11 0.0213 0.0071 0.0017 0.0295 0.5062 4.22%4 16 0.0203 0.0064 0.0078 0.0327 0.5060 4.00%4 22 0.0167 0.0053 0.0062 0.0271 0.5061 3.29%4 44 0.0158 0.0038 0.0084 0.0232 0.5057 3.12%5 11 0.0067 0.0072 βˆ’0.0074 0.0209 0.5205 1.30%5 16 0.0056 0.0065 βˆ’0.0186 0.0073 0.5196 1.08%5 22 0.0050 0.0055 βˆ’0.0057 0.0157 0.5187 0.97%5 44 0.0044 0.0040 βˆ’0.0034 0.0122 0.5204 0.85%

    Table 3: Monte Carlo local risk minimization hedging strategy of a European put option with Heston model.

    π‘˜ Hedging Standard error Monte Carlo price Standard error3 βˆ’0.4480 6.57 Γ— 10βˆ’4 10.417 5.00 Γ— 10βˆ’3

    4 βˆ’0.4506 1.28 Γ— 10βˆ’3 10.422 3.35 Γ— 10βˆ’3

    5 βˆ’0.4453 2.54 Γ— 10βˆ’3 10.409 2.75 Γ— 10βˆ’3

    10000

    Hed

    ge

    0 2000 4000 6000 8000

    βˆ’0.50

    βˆ’0.48

    βˆ’0.46

    βˆ’0.44

    βˆ’0.42

    βˆ’0.40

    k = 3

    k = 4

    k = 5

    Put option-average hedging

    Figure 2: Monte Carlo local risk minimization hedging strategy ofa European put option with Heston model.

    strategies for one-touch options written on a Heston modelwith nonzero correlation. Nevertheless, as pointed out inRemark 20, it is expected that pure hedging strategies basedon the generalized Föllmer-Schweizer decomposition miti-gate very well the average hedging error. This is what we getin the simulation results. In Table 4, we see that as π‘˜ increases,the percentage error𝐸(AV) decreases. For π‘˜ = 3, we also havea decrease in the standard error, but when π‘˜ = 4, the standarderror is almost the same (with a small increase).

    Mean Variance Hedging Strategy.Here we present the averagehedging errors associated with one-touch options written

    on a Heston model with nonzero correlation under themean variance methodology. Again, we simulate the averagehedging error along the interval [0, 1] by using π‘˜ = 3, 4 asdiscretization levels of the Brownianmotions.We perform 22and 44 hedging strategies in the period (which correspondsto one and two hedging strategies per day, resp.) withparameters π‘Ÿ = 0, 𝑏 = 0.01, πœ… = 3.63, πœƒ = 0.04, 𝜌 = βˆ’0.53,𝜎 = 0.3, π‘Œ

    0= 0.3, and 𝑆

    0= 100. The barrier of the one-touch

    option (104) is 105. There are some quantities which are notrelated to the GKW decomposition that must be computed(see Remark 20). The quantity 𝜁 is not related to the GKWdecomposition but it is described by Theorem 1.1 in Hobson[16] as follows.The process 𝜁 appearing in (5) and (6) is givenby

    πœπ‘‘= 𝑍

    0𝜌𝜎𝐹 (𝑇 βˆ’ 𝑑) βˆ’ 𝑍

    0𝑏; 0 ≀ 𝑑 ≀ 𝑇, (110)

    where 𝐹 is given by (see case 2 of Prop. 5.1 in [16])

    𝐹 (𝑑) =𝐢

    𝐴tanh(𝐴𝐢𝑑 + tanhβˆ’1 (𝐴𝐡

    𝐢)) βˆ’ 𝐡; 0 ≀ 𝑑 ≀ 𝑇,

    (111)

    with 𝐴 = √|1 βˆ’ 2𝜌2|𝜎2, 𝐡 = (πœ… + 2πœŒπœŽπ‘)/𝜎2|1 βˆ’ 2𝜌2|, and𝐢 = √|𝐷| where 𝐷 = 2𝑏2 + (πœ… + 2πœŒπœŽπ‘)2/𝜎2(1 βˆ’ 2𝜌2). Theinitial condition 𝑍

    0is given by

    𝑍0=π‘Œ20

    2𝐹 (𝑇) + πœ…πœƒβˆ«

    𝑇

    0

    𝐹 (𝑠) 𝑑𝑠. (112)

    The average hedging error results are summarized inTable 5. It reports the standard error (St. error) = stan-dard deviation of {𝛾

    β„“; 1 ≀ β„“ ≀ 𝑀}/(total number of

    scenarios 𝑀 = 1000)1/2, the price of the option, the lower

  • International Journal of Stochastic Analysis 17

    Table 4: Average hedging error with generalized Follmer-Schweizer decomposition: one-touch option with Heston model.

    π‘˜ Hedging strategies AV St. error LL UL Price 𝐸(AV)3 22 0.0422 0.0084 0.0258 0.0586 0.7399 5.70%3 44 0.0382 0.0067 0.0250 0.0515 0.7397 5.17%4 22 0.0210 0.0080 0.0053 0.0366 0.7733 2.71%4 44 0.0198 0.0082 0.0036 0.0360 0.7737 2.56%

    (LL) and upper (UL) limits of the 95% confidence interval ofAV, and the percentage error𝐸(AV) related to AV. Comparedto the local risk minimization methodology, the results showsmaller percentage errors for π‘˜ = 4. Also, in all the cases,the results show smaller values of the standard errors whichsuggests the mean variance methodology provides moreaccurate values of the hedging strategy. Again, for a fixedvalue π‘˜, when the number of hedging strategies increases, thestandard error decreases.

    Appendix

    This appendix provides a deeper understanding of the MonteCarlo algorithm proposed in this work when the represen-tation (πœ™π»,𝑆, πœ™π»,𝐼) in (43) admits additional integrability andpath smoothness assumptions. We present stronger approx-imations which complement the asymptotic result givenin Theorem 11. Uniform-type weak and strong pointwiseapproximations for πœƒπ» are presented and they validate thenumerical experiments in Tables 1 and 4 in Section 6. At first,we need some technical lemmas.

    Lemma A.1. Suppose that πœ™π» = (πœ™π»,1, . . . , πœ™π»,𝑝) is a 𝑝-dimensional progressive process such that Esup

    0β‰€π‘‘β‰€π‘‡β€–πœ™π»π‘‘β€–2R𝑝 <

    ∞. Then, the following identity holds:

    Ξ”π›Ώπ‘˜π‘‹π‘‡π‘˜,𝑗

    1

    = E [βˆ«π‘‡π‘˜,𝑗

    1

    0πœ™π»,π‘—π‘ π‘‘π‘Š(𝑗)

    𝑠| Fπ‘˜

    π‘‡π‘˜,𝑗

    1

    ]

    a.s; 𝑗 = 1, . . . , 𝑝; π‘˜ β‰₯ 1.(A.1)

    Proof. It is sufficient to prove for 𝑝 = 2 since the argumentfor 𝑝 > 2 easily follows from this case. LetH be the linearspace constituted by the bounded R2-valued F-progressiveprocesses πœ™ = (πœ™1, πœ™2) such that (A.1) holds with 𝑋 = 𝑋

    0+

    βˆ«β‹…

    0πœ™1π‘ π‘‘π‘Š(1)

    𝑠+ ∫

    β‹…

    0πœ™2π‘ π‘‘π‘Š(2)

    𝑠where 𝑋

    0∈ F

    0. Let U be the class

    of stochastic intervals of the form βŸ¦π‘†, +∞⟦ where 𝑆 is a F-stopping time. We claim that πœ™ = (1βŸ¦π‘†,+∞⟦, 1⟦𝐽,+∞⟦) ∈ H forevery F-stopping times 𝑆 and 𝐽. In order to check (A.1) forsuch πœ™, we only need to show for 𝑗 = 1 since the argumentfor 𝑗 = 2 is the same. With a slight abuse of notation, anysubsigma-algebra ofF

    𝑇of the form Ξ©βˆ—

    1βŠ—G will be denoted

    by G where Ξ©βˆ—1is the trivial sigma-algebra on the first copy

    Ξ©1.

    At first, we split Ξ© = β‹ƒβˆžπ‘›=1

    {π‘‡π‘˜π‘›= π‘‡π‘˜,1

    1} and we make the

    argument on the sets {π‘‡π‘˜π‘›= π‘‡π‘˜,1

    1}, 𝑛 β‰₯ 1. In this case, we know

    thatFπ‘˜π‘‡π‘˜,1

    1

    = Fπ‘˜,1π‘‡π‘˜,1

    1

    βŠ—Fπ‘˜,2π‘‡π‘˜,2

    π‘›βˆ’1

    a.s and

    Ξ”π›Ώπ‘˜π‘‹π‘‡π‘˜,𝑗

    1

    = Ξ”π›Ώπ‘˜(π‘Š

    (1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆) 1{𝑆

  • 18 International Journal of Stochastic Analysis

    Table 5: Average hedging error in the mean variance hedging methodology for one-touch option with Heston model.

    π‘˜ Hedging strategies AV St. error LL UL Price 𝐸(AV)3 22 0.0674 0.0052 0.0572 0.0777 0.7339 9.19%3 44 0.0577 0.0044 0.0490 0.0663 0.7340 7.86%4 22 0.0143 0.0056 0.0034 0.0252 0.7767 1.84%4 44 0.0134 0.0038 0.0060 0.0209 0.7765 1.73%

    on the set {𝐽 < π‘‡π‘˜π‘›βˆ’1

    }. By constructionFπ‘˜π‘‡π‘˜,1

    1

    = Fπ‘˜,1π‘‡π‘˜,1

    1

    βŠ—Fπ‘˜,2π‘‡π‘˜,2

    π‘›βˆ’1

    a.s and again the independence betweenπ‘Š(1) andπ‘Š(2) yields

    Ξ”π›Ώπ‘˜(π‘Š

    (1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆)

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    ) βˆ’ E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜

    π‘›βˆ’1

    )

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    )

    (A.5)

    on {π‘‡π‘˜π‘›βˆ’1

    ≀ 𝑆 < π‘‡π‘˜π‘›= π‘‡π‘˜,1

    1}. Similarly,

    Ξ”π›Ώπ‘˜(π‘Š

    (1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆)

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    ) βˆ’ E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜

    π‘›βˆ’1

    )

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    ) βˆ’ E (π‘Š(1)

    π‘‡π‘˜

    π‘›βˆ’1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜,2

    π‘‡π‘˜

    π‘›βˆ’1

    )

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    ) βˆ’ E (π‘Š(1)

    π‘‡π‘˜

    π‘›βˆ’1

    | Fπ‘˜,2

    π‘‡π‘˜

    π‘›βˆ’1

    )

    + E (π‘Š(1)

    𝑆| F

    π‘˜,2

    π‘‡π‘˜

    π‘›βˆ’1

    )

    = E(π‘Š(1)

    π‘‡π‘˜,1

    1

    βˆ’π‘Š(1)

    𝑆| F

    π‘˜

    π‘‡π‘˜,1

    1

    ) + E (π‘Š(1)

    𝑆| F

    π‘˜,2

    π‘‡π‘˜

    π‘›βˆ’1

    )

    (A.6)

    on {𝑆 < π‘‡π‘˜π‘›βˆ’1

    }. By assumption 𝑆 is an F-stopping time, whereF is a product filtration. Hence, E(π‘Š(1)

    𝑆| Fπ‘˜,2

    π‘‡π‘˜

    π‘›βˆ’1

    ) = 0 a.s on

    {𝑆 < π‘‡π‘˜π‘›βˆ’1

    }.Summing up the above identities, we will conclude

    (1βŸ¦π‘†,+∞⟦, 1⟦𝐽,+∞⟦) ∈ H. In particular, the constant process(1, 1) ∈ H and if πœ™π‘› is a sequence inH such that πœ™π‘› β†’ πœ™a.s Leb Γ— Q with πœ™ bounded, then a routine application ofBurkhölder inequality shows that πœ™ ∈ H. Since U generatesthe optional sigma-algebra, we will apply the monotone classtheorem and, by localization, wemay conclude the proof.

    Strong Convergence under Mild Regularity. In this section,we provide a pointwise strong convergence result for GKWprojectors under rather weak path regularity conditions. Letus consider the stopping times

    πœπ‘—:= inf {𝑑 > 0;

    π‘Š(𝑗)

    𝑑

    = 1} ; 𝑗 = 1, . . . , 𝑝, (A.7)

    and we set

    πœ“π»,𝑗

    (𝑒) := Eπœ™π»,𝑗

    πœπ‘—π‘’βˆ’ πœ™π»,𝑗

    0

    2

    , for 𝑒 β‰₯ 0, 𝑗 = 1, . . . , 𝑝.(A.8)

    Here, if 𝑒 satisfies πœπ‘—π‘’ β‰₯ 𝑇, we set πœ™π»,π‘—πœπ‘—π‘’

    := πœ™π»,𝑗

    𝑇and for

    simplicity we assume that πœ“π»,𝑗(0βˆ’) = 0.

    Theorem A.2. If𝐻 is a Q-square-integrable contingent claimsatisfying (M) and there exists a representation πœ™π» =(πœ™π»,1, . . . , πœ™π»,𝑝) of 𝐻 such that


Recommended