Foad Shokrollahi Option pricing in fractional models aaa ACTA WASAENSIA 425 ACADEMICDISSERTATION Tobepresented,withthepermissionoftheBoardoftheSchoolofTechnologyand InnovationoftheUniversityofVaasa,forpublicexaminationinAuditoriumNissi (,)onthe28UIofJune,2019,atnoon. Reviewers Docent Ehsan Azmoodeh Ruhr Universitat Bochum Fakulta¨t fur Mathematik Universitatsstr. 150 44780 Bochum GERMANY Docent Dario Gasbarra University of Helsinki Department of Mathematics and Statistics P.O. Box 68 FIN-00014 Helsingin yliopisto FINLAND III Julkaisija Julkaisupäivämäärä Vaasan yliopisto Kesäkuu 2019 Tekijä(t) Julkaisun tyyppi Foad Shokrollahi Artikkeliväitöskirja ORCID tunniste Julkaisusarjan nimi, osan numero https://orcid.org/0000-0003-1434-0949 Acta Wasaensia, 425 ISBN Yhteystiedot Vaasan Yliopisto Tekniikan ja innovaatiojohtamisen yksikkö MatemaattisHWWLHWHHW PL 700 65101 Vaasa 978-952-476-869-6 (painettu) 978-952-476-870-2 (verkkoaineisto) URN:ISBN:978-952-476-870-2 ISSN 0355-2667 (Acta Wasaensia 425, painettu) 2323-9123 (Acta Wasaensia 425, verkkoaineisto) Sivumäärä Kieli 88 Englanti Julkaisun nimike Optioiden hinnoittelu fraktionallisissa Malleissa Tiivistelmä Väitöskirja tarkastelee fraktionaalisen Black–Scholes -mallin ja sekoitetun fraktionallisen Black–Scholes -mallin käyttöä erityyppisten optioiden arvottamisessa. Tätä tutkitaan neljässä artikkelissa. Ensimmäisessä artikkelissa tarkastellaan geometrisia aasialaisia optioita ja potenssioptioita, kun osakehinta noudattaa aikamuunnettua sekoitettua fraktionaalista mallia. Tässä mallissa sekoitun fraktionaalisen Black–Scholes -mallin käänteinen subordinaattoriprosessi korvaa fysikaalisen ajan. Kolmannen artikkelin tarkoitus on hinnoitella eurooppalainen valuuttaoptio fraktionaalisen Brownin liikkeen mallissa aikamuunnetulla strategialla. Lisäksi aika-askeleen ja pitkän aikavälin riippuvuuden vaikutusta tutkitaan transaktiokulujen alaisuudessa. Ehdollinen keskiarvosuojaaminen fraktionaalisessa Black–Sholes -mallissa on toisen artikkelin aihe. Ehdollinen keskiarvosuojaus eurooppalaiselle vaniljaoptiolle, jolla on konveksi tai konkaavi positiivinen tuottofunktio transaktiokulujen vallitessa, on artikkelin päätulos. Neljännessä artikkelissa tutkitaan eurooppalaisia optioita diskreetissä ajassa mallissa, joka on hypyllinen sekoitettu fraktionaalinen Brownin liike. Käyttäen keskiarvoista deltasuojausstrategiaa artikkelissa johdetaan hinnoittelumalli eurooppalaisille optioille transaktiokulujen vallitessa. Asiasanat Optioiden hinnoittelu, Stokastinen mallinnus, Matemaattinen rahoitusteoria, Fraktionaaliset mallit, Fraktionaalinen Brownin liike V Publisher Date of publication Vaasan yliopisto June 2019 Author(s) Type of publication Foad Shokrollahi Doctoral thesis by publication ORCID identifier Name and number of series https://orcid.org/0000-0003-1434-0949 Acta wasaensia, 425 ISBN Contact information University of Vaasa School of Technology and Innovation Mathematics and Statistics P.O. Box 700 FI-65101 Vaasa Finland 978-952-476-869-6 (print) 978-952-476-870-2 (online) URN :ISBN :978-952-476-870-2 ISSN 0355-2667 (Acta wasaensia 425, print) 2323-9123 (Acta wasaensia 425, online) Number of pages Language 88 English Title of publication Option pricing in fractional models Abstract This thesis deals with application of the fractional Black-Scholes and mixed fractional Black-Scholes models to evaluate different type of options. These assessments are considered in four individual papers. In the first articles, the problem of geometric Asian and power options pricing is investigated when the stock price follows a time changed mixed fractional model. In this model, an inverse subordinator process in the mixed fractional Black-Scholes model replaces the physical time. The aim of the third paper is to evaluate the European currency option in a fractional Brownian motion environment by the time-changed strategy. Also, the impact of time step and long range dependence are obtained under transaction costs. Conditional mean hedging under fractional Black-Scholes model is the propose of the second article. The conditional mean hedge of the European vanilla type option with convex or concave positive payoff under transaction costs is obtained. In the fourth article, the mixed fractional Brownian motion with jump process are incorporated to analyze European options in discrete time case. By a mean delta hedging strategy, the pricing model is proposed for European option under transaction costs. Keywords Option pricing , Stochastic modeling, Mathematical finance, Fractional model, Fractional Brownian motion VII ACKNOWLEDGEMENTS I would like to first thank my supervisor Prof Tommi Sottinen, for his constant encouragement, generosity, and friendship during the period in which this work was completed. I owe him almost completely for the opportunity to devote myself to mathematics and work that was meaningful to me, and will be forever in his debt. I also would like to extend my gratitude to all the members of the Department of Mathematics and Statistics of the University of Vaasa for the friendly working environment. Furthermore, I am thankful to the University of Vaasa for their financial support which made it possible to attend a number of conferences and workshops. I wish to thank my pre-examiners Docent Dario Gasbarra and Docent Ehsan Azmoodeh for reviewing my thesis and for their feedbacks. Finally, I must thank my family for their constant support. In particular, I thank my wife, Arezoo, and my lovely daughter, Lara for their patience, friendship, and understanding while I completed this work. Vaasa, May 2019 Foad Shokrollahi IX CONTENTS 1 INTRODUCTION. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2 STOCHASTIC PROCESSES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 2.1 General facts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 2.2 Le´vy processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 2.3 Long-range dependence and self-similarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.4 Gaussian processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.4.1 Brownian motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.4.2 Fractional Brownian motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.4.3 Mixed fractional Brownian motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 3 ESSENTIALS OF STOCHASTIC ANALYSIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 3.1 Itoˆ’s lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 3.2 Girsanov’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 4 FUNDAMENTAL ELEMENTS OF STOCHASTIC FINANCE . . . . . . . . . . . . 15 4.1 Useful financial terminologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 4.2 Classical Black-Scholes market model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 4.3 PDE approach in option pricing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 4.4 (Mixed) Fractional Black-Scholes market model . . . . . . . . . . . . . . . . . . . . . . 19 4.4.1 Fractional Black-Scholes market model . . . . . . . . . . . . . . . . . . . . . . . 19 4.4.2 Mixed fractional Black-Scholes market model . . . . . . . . . . . . . . . . 19 5 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 5.1 Summaries of the articles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 XReferences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 XI LIST OF PUBLICATIONS The dissertation is based on the following three refereed articles: (I) Shokrollahi, F. (2018). The evaluation of geometric Asian power options under time changed mixed fractional Brownian motion. Journal of Computa- tional and Applied Mathematics 344, 716–724. (II) Shokrollahi, F., and Sottinen, T. (2017). Hedging in fractional Black-Scholes model with transaction costs. Statistics & Probability Letters. 130, 85–91. (III) Shokrollahi, F. (2018). Subdiffusive fractional Black-Scholes model for pri- cing currency options under transaction costs. Cogent Mathematics & Statis- tics 5, 1470145. (IV) Shokrollahi, F. (2018). Mixed fractional Merton model to evaluate European options with transaction costs. Journal of Mathematical Finance 8, 623–639. All the articles are reprinted with the permission of the copyright owners. (permis- sion is needed from the publisher!) XIII AUTHOR’S CONTRIBUTION Publication I: “The evaluation of geometric Asian power options under time changed mixed fractional Brownian motion” This is an independent work of the author. Publication II: “Hedging in fractional Black—Scholes model with transaction costs” This article is the outcome of a joint discussion and all the results are a joint work with Tommi Sottinen. Publication III: “Subdiffusive fractional Black-–Scholes model for pricing cur- rency options under transaction costs ” This is an independent work of the author. Publication IV: “Mixed fractional Merton model to evaluate European options with transaction costs” This is an independent work of the author. 1 INTRODUCTION In recent years, options have received increasing attention and their role has grown rapidly in most trading exchanges. Most common examples of variables underlying options are the price of stocks, bonds or commodities traded on the exchange. Ho- wever, options can depend on almost any variable, from the price of pork bellies to the amount of rainfall in a certain geographic area. Standardized options on stock prices have been traded at the exchange since 1973. There are two basic types of options. A call option gives the holder the right (and not the obligation) to buy the underlying asset at a certain time in the future for a certain price. A put option gives the holder the right to sell the underlying asset at a certain time in the future for a certain price. The price for which the asset is being exchanged is referred to as the strike price or the exercise price. The future time when the exchange takes place is referred to as the maturity of the option or the expiration date. Based on the exercise conditions, options are categorized into European options and American options. European options can only be exercised at maturity. American options can be exercised anytime during the life of the option. Throughout this section options are assumed European unless otherwise specified. Depending on the strategy, options trading can provide a variety of benefits, inclu- ding the security of limited risk and the advantage of leverage. Another benefit is that options can protect or enhance your portfolio in rising, falling and neutral markets. Since it appeared in the 1970s, the Black-Scholes (BS) model (Black & Scholes (1990)) has become the most popular method to option pricing and its generalized version has provided mathematically beautiful and powerful results on option pri- cing. However, they are still theoretical adoptions and not necessarily consistent with empirical features of financial returns, such as nonindependence, nonlinearity, ect. For example, Hull and White (Hull & White (1987)) introduced a bivariate diffusion model for pricing options on assets with stochastic volatilities. Heston (Heston (1993)) proposed affine stochastic volatility. Furthermore, since discon- tinuity or jumps is one of the significant component in financial asset pricing (see Andersen, Benzoni & Lund (2002), Chernov, Gallant, Ghysels & Tauchen (2003), Pan (2002), Eraker (2004)) and also some scholars have been represented pricing models based on the jump processes (see Merton (1976), Kou (2002), Cont & Tan- kov (2004), Ahn, Cho & Park (2007), Ma (2006)). Many realistic models have been described long memory behavior in financial time series (Lo, A. W. (1991), Willinger, W., Taqqu, M. S., & Teverovsky, V. (1999) (1999), Cont, R. (2005), Dai & Singleton (2000), Berg & Lyhagen (1998), Hsieh (1991), Huang & Yang (1995)). Since, fractional Brownian motion ( f Bm) is a self- similar and long-range dependence process, then it can be a appropriate candidate to capture these phenomena (Wang, Zhu, Tang & Yan (2010), Wang (2010), Sottinen 2 Acta Wasaensia (2003), Sottinen & Valkeila (2003), Wang, Wu, Zhou & Jing (2012), Xiao, Zhang, Zhang & Wang (2010), Zhang, Xiao & He (2009), Cartea & del Castillo-Negrete (2007)). Kolmogorov introduced the f Bm in 1940. A representation theorem for Kolmogorov’s process was introduced by Mandelbrot and Van Ness (Mandelbrot & Van Ness (1968)). Nowadays, the f Bm process play a significant role in stochastic finance and different extensions of the fractional Black-Scholes formulas for pricing options based on the geometric fractional Brownian motion are proposed to capture the behavior of underlying asset (Bayraktar, Poor & Sircar (2004), Meng & Wang (2010)). On the mathematical side, f Bm is neither a semi martingale nor a Markov process (except in the Brownian motion case). Hence, the classical stochastic integration theory developed for semimartingale is not handy to analyze financial markets based on fractional Brownian motion (Hu, Y., & Øksendal, B. (2003)). Further, some aut- hors discussed arbitrage under fractional Black-Scholes model and proposed some restrictions to exclude arbitrage in fractional markets (see Bender, Sottinen & Val- keila (2007), Bender, C., Sottinen, T., & Valkeila, E. (2008) (2008), Bender & Elliott (2004), Bjo¨rk & Hult (2005)). To better describe long memory property and fluctuations in the financial assets, the mixed fractional Brownian motion (mfBm) was presented (see El-Nouty (2003), Mishura (2008), Cheridito, P. (2001), Zili (2006)). A mfBm is a family of Gussian process which is a linear combination of Brownian motion and independent f Bm with Hurst parameter H ∈ (12 ,1). The pioneering work to apply the mfBm in finance was presented by Cheridito, P. (2001). He proved that for H ∈ (34 ,1), the mfBm is equivalent to one with Brownian motion, and then its free of arbitrage. For H ∈ (12 ,1), Mishura and Valkeila (Mishura & Valkeila (2002)) proved that the mixed model is arbitrage-free. Acta Wasaensia 3 2 STOCHASTIC PROCESSES In this section, we provide some definitions and auxiliary facts that are needed in this thesis (for further details, see Shiryaev, A. N., do Rosa´rio Grossinho, M., Oliveira, P. E., & Esquı´vel, M. L. (2006) (2006), Fo¨llmer, H., & Schied, A. (2011), Shiryaev, A. N. (1999), Kallenberg, O. (2006), Clark & Ghosh (2004), Melnikov, Pliska (1997), Mikosch (1998)). Throughout this section all random objects are defined in the probability space (Ω,F ,P). 2.1 General facts Definition 2.1. Let T ⊆ [0,+∞) be an interval. A stochastic process X indexed by interval T is a collection of random variables (Xt)t≥0. Also, for every ω ∈Ω, the real valued function t ∈ T → Xt(ω) is called a trajectory or a sample path of the process X . Definition 2.2. A filtration is a family (Ft)t≥0 of σ -algebrasFt ⊆F such that ∀0≤ s≤ t ⇒Fs ⊆Ft . Definition 2.3. Let (Ft)t≥0 be a filtration. A stochastic process X = (Xt)t≥0 is said to be adapted to the filtration (Ft)t≥0 if for every t ≥ 0, the random variable Xt is Ft-measurable. Definition 2.4. An stochastic process (Xt)t≥0 is called a martingale with respect to the filtration (Ft)t≥0 ( and probability measure P) if the following conditions are satisfied: (i) Xt isFt-measurable for all t ≥ 0, (ii) E[|Xt |]< ∞ for all t ≥ 0, and (iii) E[Xt |Fs] = Xs for all 0≤ s≤ t. Definition 2.5. AnFt - adapted stochastic process is called a local martingale with respect to the given filtration (Ft)t≥0 if there exists an increasing sequence ofFt - stopping times τk such that τk → ∞ as k → ∞, 4 Acta Wasaensia and (Xt∧τk)t≥0, (2.1) is anFt- martingale for all k, where t ∧ τk = min(t,τk). Definition 2.6. A process X = (Xt)t≥0 is called an (Ft)t≥0- semi-martingale, if it admits the representation Xt = X0+Mt +At , (2.2) where M is an (Ft)t≥0- local martingale with M0 = 0, A is a process of locally bounded variation and adapted to filteration (Ft)t≥0, X0 isF0-measurable. Definition 2.7. (Ho¨lder continuous) Let α ∈ (0,1]. A function f : R → R is said to be locally α-Ho¨lder continuous at x ∈ R, if there exists ε > 0 and c = cx such that | f (x)− f (y)| ≤ c|x− y|α , for all y ∈ R with |y− x|< ε. Definition 2.8. A stochastic process X = (Xt)t≥0 is said to have stationary incre- ments if for all s≥ 0, and every h > 0, (Xt −Xs)t≥0 f.d.= (Xt+h−Xs+h)t≥0. (2.3) Here f.d. = denotes equality in finite dimensional distribution. Definition 2.9. A stochastic process X = (Xt)t≥0 is said to have independent incre- ments if for every t ≥ 0 and any choice ti ∈ T with t0 < t1 < ... < tn and n≥ 1, Xt2 −Xt1 , ...,Xtn −Xtn−1 (2.4) are independent random variables. 2.2 Le´vy processes Le´vy processes are stochastic processes with independent and stationary incre- ments. Definition 2.10. Le´vy process (Xt)t>0 is a process with the following properties Acta Wasaensia 5 (1) Independent increments, (2) Stationary increments, and (3) Continuous paths in probability: That is limh→0P(|Xt+h −Xt | ≥ ε) = 0 for any ε > 0. Definition 2.11. (Subordinator process) A subordinator is a real-valued Le´vy process with nondecreasing sample paths. Definition 2.12. (Stable process) A stable process is a real-valued Le´vy process (Xt)t≥0 with initial value X0 = 0 that satisfies the self-similarity property (Xat)t≥0 f.d. = (a1/αXt)t≥0 ∀t > 0. The parameter α is called the exponent of the process. Definition 2.13. (Poisson process) A Poisson process (Xt)t≥0 satisfies the following conditions: (1) X0 = 0, (2) Xt −Xs are integer valued for 0≤ s < t < ∞ and P(Xt −Xs = k) = λ k(t− s)k k! e−λ (t−s) for k = 0,1,2, ... (2.5) (3) The increments Xt2 −Xt1 , Xt4 −Xt3 ... and Xtn −Xtn−1 are independent for every 0≤ t1 < t2 < t3 < t4 < ... < tn. Example 2.14. The fundamental Le´vy processes are the Brownian motion (defined later) and the Poisson process. The Poisson process is a subordinator, but is not stable; the Brownian motion is stable, with exponent α = 2. 2.3 Long-range dependence and self-similarity Definition 2.15. Let (Xt)t≥0 be a process with stationary trajectories and (rn)n∈N the autocovariance sequence defined by ∀n ∈ N, rn = E[Xn+1X1]. (2.6) 6 Acta Wasaensia Then, the process X = (Xt)t≥0 is called long-range dependence if ∑ n∈N rn = ∞. Remark 2.16. Since (Xt)t≥0 is a process with stationary trajectories ∀s≥ 0,∀n ∈ N, rn = E[Xn+sXs]. Definition 2.17. Let H ∈ (0,1]. A stochastic process X = (Xt)t≥0 is said to be self-similar with exponent H, if for any a > 0, (Xat)t≥0 f.d. = (aHXt)t≥0. Acta Wasaensia 7 2.4 Gaussian processes Definition 2.18. A stochastic process (Xt)t∈T is Gaussian if all finite dimensional projections (Xt)t∈T0 , T0 ⊂ T finite, are multivariate Gaussian. 2.4.1 Brownian motion Definition 2.19. Brownian motion is a process (Bt)t≥0 with the following proper- ties: (1) B0 = 0, (2) Bt has independent increments, (3) Bt −Bs ∼ N(0, t− s) for s < t, here N is the normal distribution function. Definition 2.20. (Markov Process) The process (Xt)t∈T is a Markov process if E[ f (Xt)|Fs] = E[ f (Xt)|Xs], ∀t > s, t,s ∈ T, whereFs = σ{Xu;u≤ s}, and f is a bounded Borel function. 2.4.2 Fractional Brownian motion f Bm has recently become a useful choice for modeling in mathematical finance and other sciences. On purely empirical data, some believe that f Bm is an ideal candidate since it enjoys two important statistical features of long memory and self-similarity. Even with its popularity, our understanding of the properties and behaviour of f Bm is limited. Kolmogorov (Kolmogorov (1941)) was the first to introduce the Gaussian process which is now known as f Bm in the theory of probability. This class of processes was studied by Kolmogorov in detail and it played an essential role in the series of problem s of the statistical theory of turbulence. Yaglom (Yaglom (1955)) dis- cussed the spectral density and correlation function of f Bm. A quadratic variation formula for f Bm follows from a general result of Baxter (Baxter (1956)). Gladys- hev (Gladyshev (1961)) extended Baxter’s result and provided a theoretical result to determine the value of the Hurst effect denoted by H. However, most of the en- comium to f Bm has been given to Mandelbrot and Van Ness (Mandelbrot & Van 8 Acta Wasaensia Ness (1968)) who used f Bm to model natural phenomena such as the speculative market fluctuations. For get more information about f Bm, you can see, Hu, Y., & Øksendal, B. (2003), Nualart (2006), Biagini, Hu, Øksendal & Zhang (2008), Mishura (2008). Definition 2.21. The fractional brownian motion with Hurst index H ∈ (0,1) deno- ted by (BHt )t∈R, is the centered Gaussian process with covariance function RH(s, t) = 1 2 (|t|2H + |s|2H −|t− s|2H) , s, t ∈ R. Figure (1) shows the sample path of the f Bm for different parameter. Figure 1. f Bm with different Hurst parameter H . The f Bm can be represented in terms of the one-sided or two-sided standard Brow- nin motion see (Nualart (2006)). First, we review some special function involved in the representation results. The Gamma function is defined by Γ(α) = ∫ ∞ 0 exp(−v)vα−1dv, α > 0. The Beta function is defined by β (α,β ) = ∫ 1 0 (1− v)α−1vβ−1dv, α,β > 0. Acta Wasaensia 9 The Gauss hypergeometric function of parameters a,b,c and variable z ∈ R is defi- ned by the formal power series F(a,b,c,z) = ∞ ∑ k=0 (a)k(b)kzk (c)kk! . where (a)k = a(a+1)...(a+ k−1). (1) The one-sided f Bm can be constructed from a one-sided Brownian motion: Theorem 2.22. Molchan-Golosov representation (Molchan & Golosov (1969)) For H ∈ (0,1), it holds that BHt =C(H) ∫ t 0 (t− s)H− 12F ( 1 2 ,H− 1 2 ,H+ 1 2 , s− t s ) dBs, where C(H) = 2H Γ(H+ 12 ) . (2) The two-sided f Bm can be written in terms of one-sided Brownian motion: Mandelbrot-Van Ness representation (Mandelbrot & Van Ness (1968)) Theorem 2.23. For H ∈ (0,1), it holds that BHt =C(H) ∫ ( (t− s)H− 121(−∞,t)(s)− (−s)H− 1 21(−∞,0)(s) ) dB˜s, here B˜ represents two-sided Brownian motion. Theorem 2.24. (Mishura (2008)) For H = 12 , f Bm is neither a Markov process nor a semimartingale. Remark 2.25. Since f Bm is not a semimartngale,the classical integration theory de- veloped for semimartingale is not available, Mishura (2008). Then, many scholars introduced two different approaches for stochastic integral with respect to f Bm (1) Pathwise approach (2) Malliavin calculus (Skorokhod integration) approach (Nual- art (2006), Sottinen, T., & Viitasaari, L. (2016)). Remark 2.26. Using stationarity increment of f Bm, it can be shown that the auto- covariance function γn = of the sequence (Xn)n≥1 := (BHn+1−Bn)n≥1 is given by γn = n ∑ k=0 E[BH1 (B H k+1−BHk )] = 1 2 [ (n+1)2H −2n2H +(n−1)2H ] , (2.7) 10 Acta Wasaensia therefore γn ≈ H(2H−1)n2H−2, as n→ ∞,H = 12 . (2.8) Notice that when (1) H = 12 ,γn = 0,∀n, therefore f Bm has independent increments. (2) If H > 12 ,γn > 0, the increments of the f Bm process are positively correlated and by p-series ∑∞n=1 |γn|= ∞, therefore has long-range dependence. (3) If H < 12 ,γn < 0, the increments of the f Bm process are negatively correlated and by p-series ∑∞n=1 |γn|= c < ∞, therefore has short-range dependence. 2.4.3 Mixed fractional Brownian motion Let a and b be two real constants such that (a,b) = (0,0). Definition 2.27. A mfBm with parameters a,b, and H is a process MH = (MHt (a,b))t≥0, defined by MHt = M H t (a,b) = aBt +bB H t , ∀t ≥ 0 (2.9) where B is a Brownian motion and BH is an independent f Bm with Hurst parame- ter H (Cheridito, P. (2001), van Zanten, H. (2007), Mishura (2008), Zili (2006), Marinucci & Robinson (1999)). Proposition 2.28. The m fBm has the following properties (i) MH is a centered Gaussian process, (ii) for all t ∈ R+,E((MHt (a,b))2) = a2t+b2t2H, (iii) one has that Cov ( MHt (a,b),M H s (a,b) ) = a2(t ∧ s)+ 1 2 b2 [ t2H + s2H −|t− s|2H ] ,∀s, t ∈ R+, (2.10) where t ∧ s = 1/2(t+ s+ |t− s|), (iv) the increments of the m fBm are stationary. Acta Wasaensia 11 Lemma 2.29. (Zili (2006)) For any h > 0,(MHht (a,b))t≥0 f.d. = (MHt (ah 1 2 ,bhH)))t≥0. This property will be called the mixed-self-similarity. Theorem 2.30. (Zili (2006)) For all H ∈ (0,1) − {12},a ∈ R and b ∈ R −{0},(MHt (a,b))t≥0 is not a Markov process. Theorem 2.31. (Cheridito, P. (2001)) For H ∈ (34 ,1], the m fBm is equivalent to Brownian motion. Theorem 2.32. (Zili (2006)) For all a ∈ R and b ∈ R− {0}, the increments of (MHt (a,b))t∈R+ are positively correlated if 1 2 < H < 1, uncorrelated if H = 1 2 , and negatively correlated if 0 < H < 12 . Lemma 2.33. (Zili (2006)) For all a ∈ R and b ∈ R− {0}, the increments of (MHt (a,b))t∈R+ are long-range dependence if and only if H > 1 2 . Theorem 2.34. (Ho¨lder continuity) (Zili (2006)) For all T > 0 and γ < 12 ∧H, the m fBm has a modification which sample paths have a Ho¨lder-continuity, with order γ , on the interval [0,T ]. 12 Acta Wasaensia 3 ESSENTIALS OF STOCHASTIC ANALYSIS 3.1 Itoˆ’s lemma Let (Xt)t≥0 be a stochastic process and suppose that there exists a real number X0 and two (Ft)t≥0-adapted processes μ = (μt)t≥0 and σ = (σt)t≥0 such that the fol- lowing relation holds for all t ≥ 0, Xt = X0+ ∫ t 0 μsds+ ∫ t 0 σsdBs. (3.1) where the stochastic integral in (3.1) is an Itoˆ integral, such processes are called Itoˆ diffusions. We can write the equation as follows dXt = μtdt+σtdBt , (3.2) Then, we can say X satisfies the SDE given by (3.2) with the initial condition X0 given. Note that the formal notation dXt = μtdt+σtdBt is only formal. It is simply a shorthand version of the expression (3.2) above. In option pricing, we often take as given a stochastic differential equation repre- sentation for some basic quantity such as stock price. Many other quantities of interest will be functions of that basic process. To determine the dynamics of these other processes, we shall apply Itoˆ’s Lemma, which is basically the chain rule for stochastic processes (Mikosch (1998), Tong (2012), Øksendal (2003)). Theorem 3.1. (Itoˆ’s Lemma) Assume the stochastic process Xt satisfies in the following equation dXt = μtdt+σtdBt , (3.3) where μ = (μt)t≥0 and σ = (σt)t≥0 are adapted processes to a filtration (Ft)t≥0. Let Y be a new process defined byYt = f (t,Xt) where f (t,x) is a function twice diffe- rentiable in its first argument and once in its second. Then Y satisfies the stochastic differential equation: dYt = (∂ f ∂ t (t,Xt)+μt ∂ f ∂x (t,Xt)+ 1 2 σ2t ∂ 2 f ∂x2 ) (t,Xt)dt+σt ∂ f ∂x (t,Xt)dBt . Theorem 3.2. For fBm, we have Skorokhod and Fo¨llmer types Itoˆ’s lemma ( Dun- can, T. E., Hu, Y., & Pasik-Duncan, B. (2000) (2000), Sottinen & Valkeila (2003)) Acta Wasaensia 13 If f : R→ R is a twice continuously differentiable function with bounded derivatives to order two, then the Skorokhod integral is f (BHT )− f (BH0 ) = ∫ T 0 f ′(BHs )δB H s +H ∫ T 0 s2H−1 f ′′(BHs )ds, (3.4) where δ is divergence operator connected to Brownian motion. For H > 12 , the Fo¨llmer or pathwise integral is f (BHT )− f (BH0 ) = ∫ T 0 f ′(BHs )dB H s , (3.5) and for H < 12 , ∫ T 0 f ′(BHs )dBHs does not exist as pathwise integral in general. 3.2 Girsanov’s Theorem Definition 3.3. Two measures P and Q on a measurable space (Ω,F ) are equivalent if P(A) = 0⇔ Q(A) = 0, ∀A ∈F . (3.6) The Radon-Nikodym derivative can be defined by using two equivalent measures as follows: Mt = dQ dP |F t , (3.7) which enables us to change a measure to another. It follows that for any random variable X that isFt-measurable EP[XM] = ∫ Ω X(w)Mt(ω)dP(ω) = ∫ Ω X(ω)dQ(ω) = EQ[X ]. (3.8) To change the measures for Brownian motion we can use the Girsanov’s theorem. Theorem 3.4. (Girsanov’s Theorem) 14 Acta Wasaensia Assume we have (Ft)t≥0 be a filtration on interval [0,T ] where T < ∞. Define a random process M: Mt = exp [ − ∫ t 0 λudBPu − 1 2 ∫ t 0 λ 2u du ] , t ∈ [0,T ]. (3.9) where BP is a Brownian motion under probability measure P and λ is an (Ft)t≥0- adapted process that satisfies the Novikov condition E { exp [1 2 ∫ t 0 λ 2u du ]} < ∞, t ∈ [0,T ]. (3.10) If we define BQ as BQt = B P t + ∫ t 0 λudu, t ∈ [0,T ], (3.11) then the following outcomes holds: (i) M is the Radon-Nikodym martingale Mt = dQ dP |Ft . (ii) BQ is a Brownian motion under the probability measure Q. Acta Wasaensia 15 4 FUNDAMENTAL ELEMENTS OF STOCHASTIC FI- NANCE 4.1 Useful financial terminologies Definition 4.1. (Financial derivatives) A financial derivative is a contract whose value depends on one or more securities or assets, called underlying assets. Typically the underlying asset is a stock, a bond, a currency exchange rate or the quotation of commodities such as gold, oil or wheat. Definition 4.2. The spot price (stock price) is the current price in the marketplace at which a given asset—such as a security, commodity, or currency—can be bought or sold for immediate delivery. The strike price will be denoted by S. Definition 4.3. The strike (exercise) price is the price at which a derivative can be exercised, and refers to the price of the derivative’s underlying asset. The strike price will be denoted by K. Definition 4.4. Expiration date (maturity time) is date on which the option can be exercised. This will be denoted by T . Definition 4.5. A European call (Put) option grants the right to purchase (sell) a stock at a specific time called maturity T for a specific amount K called the exercise price. The value of a European call option is denoted by (ST − K)+ where (x)+ = max(x,0). Similarly, the value of a European put option is (K−ST )+. This amount is called the option payoff. Here, ST is the spot price at time T . Definition 4.6. A risk free interest rate, denoted by r, is the rate of return on an asset that possesses no risk. Remark 4.7. A dividend payout during the life of an option will have the affect of decreasing the value of a call and increasing the value of a put, the stock price typically falls by the amount of the dividend when it is paid. This will be denoted by D. Definition 4.8. A currency option is a contract, which gives the owner the right but not the obligation to purchase or sell the indicated amount of foreign currency at a specified price within a specified period of time (American option) or on a fixed date (European option). 16 Acta Wasaensia Definition 4.9. Asian options are known as path dependent options whose payoff depends on the average stock price and a fixed or floating strike price during a specific period of time before maturity. There exist two types of Asian options such as fixed strike and floating strike options. The payoff for a fixed strike price option is (AT −K)+ and (K −AT )+ for a call and put option respectively where K denotes the strike price, T is the strike time and AT is the average price of the underlying asset over the predetermined interval. For a floating strike price option, the payoffs are (ST −AT )+ and (AT − ST )+, for a call and put option respectively where ST stands for the price of stock at time T . Further, Asian options can be also categorized based on different averaging, namely that (i) geometric average that is AT = exp{ 1T ∫ T 0 logStdt}. (ii) arithmetic average AT = 1 T ∫ T 0 Stdt. Consider a financial market consisting of n assets with prices S1t , ...,S n t , which under probability measure P are governed by the following stochastic differential equati- ons: dSit = μ i t dt+σ i t dB i t , i = 1,2, ...,n, (4.1) where B = (B1, ...,Bn) is an n dimentional Brownian motion and, μ i and σ i are adapted to the natural filtration of the Brownian motion B. Next, we denote an n-dimensional stochastic process θt = (δ 1t , ...,δ nt ) as a trading strategy, where δ it denotes the holding in asset i at time t. The value Vt(δ ) at time t of a trading strategy δ is given by Vt(δ ) = n ∑ i=1 δ it S i t . (4.2) Definition 4.10. A self-financing trading strategy is a strategy δ with the property: Vt(δ ) =V0(δ )+ n ∑ i=1 ∫ t 0 δ it dS i t , t ∈ [0,T ]. (4.3) Definition 4.11. An arbitrage opportunity is a self-financing trading strategy δ with Acta Wasaensia 17 (i) V0(δ )≤ 0 a.s. (ii) VT (δ )≥ 0 a.s. and P[VT (δ )> 0]> 0. In words, arbitrage is a situation where it is possible to make a profit without the possibility of incurring a loss. Definition 4.12. Time value of an option is basically the risk premium that the seller requires to provide the option buyer with the right to buy/sell the stock up to the expiration date. Definition 4.13. Portfolio hedging describes a variety of techniques used by inves- tment managers, individual investors and corporations to reduce risk exposure in an investment portfolio. Hedging uses one investment to minimize the negative impact of adverse price swings in another. Hedging of options is one of the central problems in mathematical finance and it has been studied extensively in various setups. The idea of hedging is to replicate a claim f (ST ) by trading only the underlying asset. In mathematical terms, we are interested in finding a predictableH such that f (ST ) =C+ ∫ T 0 HsdSs (4.4) where the deterministic constant C is called the hedging cost. In arbitrage free models the hedging cost can be interpreted as the fair price of the option. Definition 4.14. A risk-neutral measure is a probability measure such that each share price is exactly equal to the discounted expectation of the share price un- der this measure. In other words, a risk neutral measure is any probability mea- sure, equivalent to the market measure P , which makes all discounted asset prices martingales. This is heavily used in the pricing of financial derivatives due to the fundamental theorem of asset pricing, which implies that in a complete market a derivative’s price is the discounted expected value of the future payoff under the unique risk-neutral measure. Definition 4.15. Greeks are the quantities representing the sensitivity of the price of derivatives such as options to a change in underlying parameters on which the value of an instrument or portfolio of financial instruments is dependent. The name is used because the most common of these sensitivities are denoted by Greek letters. Traders use different Greek values, such as delta, theta, and others, to assess options risk and manage option portfolios. Definition 4.16. Transaction costs are the costs incurred during trading – the pro- cess of selling and purchasing – on top of the price of the product that is changing hands. Transaction costs may also refer to a fee that a bank, broker, underwriter or other financial intermediary charges. The difference between what a dealer and buyer paid for a security is one of the transaction costs. 18 Acta Wasaensia 4.2 Classical Black-Scholes market model In the classical Black-Scholes market model, two assets are traded continously over the time interval [0,T ]. Denote by A the riskless asset, or bond, and by S the risky asset, or stock. The dynamic of asset price is governed by a geometric Brownian motion: dAt = rtAtdt, dSt = μtStdt+σStdBt , where r is a deterministic interest rate, σ > 0 is constant, B is a Brownian motion. The function μ is the deterministic drift of the stock. 4.3 PDE approach in option pricing Given the drift rate μ and the volatility σ , the geometric Brownian motion for the stock price process St is given by: dSt = μtStdt+σStdBt , where B is a Brownian motion. Then, applying Itoˆ’s lemma and self financing strategy the value of an option V (t,St) with a bounded payoff f (ST ) satisfies the following SDE: dV = ( ∂V ∂ t +μSt ∂V ∂St + 1 2 σ2S2t ∂ 2V ∂S2t ) dt+σSt ∂V ∂St dBt Setting up a hedging portfolio with one unit in V and Δ units of the stock St with Δ= ∂V∂St gives us Black-Scholes PDE with the drift μ replaced by the risk-free rate r, i.e.: ∂V ∂ t + rSt ∂V ∂St + 1 2 σ2S2t ∂ 2V ∂S2t = rV. After delta hedging has removed the risk of the portfolio of one unit in V and Δ Acta Wasaensia 19 units in stock St , the right stock price process to consider is the one in the risk neutral measure, as: dSt = rStdt+σStdBt , Feynman-Kac representation of SDE tell us that PDE of the Black- Scholes kind have an equivalent probabilistic representation (see, Pascucci, A. (2011)). That is, Feynman-Kac assures that one can solve for the price of the derivative V (t,St) by either discretizing the Black-Scholes PDE using finite difference methods, or by exploiting the probabilistic interpretation and using Monte Carlo methods. The Black-Scholes PDE implies that the price of the derivative V (t,St) at time t is equi- valent to the discounted value of the expected payoff at expiration (time T ). This is the famous Feynman-Kac representation: V (t,St) = EQ[e−r(T−t)V (T,ST )|Ft ]. 4.4 (Mixed) Fractional Black-Scholes market model 4.4.1 Fractional Black-Scholes market model The first try to involve the fractional Brownian motion in modeling of the market was simply to replace the B with BH in the Black-Scholes model. To interpret the integrals in the pathwise way (which is possible if H > 12 ), which is natural from the point of view. In this case the dynamic of asset price in the fractional Black-Scholes market model is given by dSt = μtStdt+σStdBHt . 4.4.2 Mixed fractional Black-Scholes market model If one wants to introduce an economically meaningful market model with long range dependent returns, mixed models are a good option. In such models, one can have both long range dependence and no-arbitrage in the sense of (Bender et al. 20 Acta Wasaensia (2008)). In the mixed fractional market model the price of an asset is modeled as dSt = μtStdt+σStdBt +σStdBHt . where B is a Brownian motion and BH is a f Bm. Acta Wasaensia 21 5 CONCLUSIONS 5.1 Summaries of the articles I. The evaluation of geometric Asian power options under time changed mixed fractional Brownian motion This paper deals with pricing a geometric Asian option under time changed mixed fractional Brownian motion. In this model, to get better behaving financial market model, we replace the physical time t with the inverse α- subordinator process Tα(t) in the mixed fractional Black-Scholes model. Then, we apply this result to propose a new model for pricing asian power options. Finally, a lower bound for asian option price is introduced. II. Hedging in fractional Black—Scholes model with transaction costs In this paper, we consider the discounted fractional Black–Scholes pricing model where the riskless investment, or the bond, is taken as the numeraire and risky asset S = (St)t∈[0,T ] is given by the dynamics dSt = Stμdt+StσdBt , where B is the fractional Brownian motion with Hurst index H ∈ (12 ,1). Then, we obtain the conditional mean hedge of the European vanilla type option with transaction costs. III. Subdiffusive fractional Black—Scholes model for pricing currency options under transaction costs A generalization of the fractional Black-Scholed model is proposed for pricing an European currency option in a discrete time market model with transaction cost. We assume that the stock price follows the subdiffusive fractional Black-Scholes model St = S0 exp { (rd − r f )Tα(t)+σBHTα (t) } , S0 > 0, 22 Acta Wasaensia where Tα(t) is the inverse of an α-stable subordinator, with α ∈ (12 ,1),H ∈ [12 ,1),α +αH > 1. In this paper, an analytical pricing model for European cur- rency option under transaction costs is introduced in discrete time case. IV. Mixed fractional Merton model to evaluate European options with tran- saction costs The problem of European option pricing in discrete time are discussed under a mixed fractional version of the Merton model with transaction costs. In this case, it has been assumed that the price dynamics St satisfies in the following St = S0 exp { μt+σBt +σHBHt +Nt lnJ } , S0 > 0, where S0,μ,σ ,σH are constant, Bt is a Brownian motion; BHt is a fractional Brow- nian motion with Hurst parameter H ∈ (34 ,1), Nt is a Poisson process with intensity λ > 0 and J is a positive random variable. We assume that Bt ,BHt ,Nt and J are independent. Finally, we evaluate impact of parameters on the option price. Acta Wasaensia 23 REFERENCES Ahn, C.M., Cho, D.C. & Park, K. (2007). The pricing of foreign currency options under jump-diffusion processes. Journal of Futures Markets: Futures, Options, and Other Derivative Products 27, 7, 669–695. Andersen, T.G., Benzoni, L. & Lund, J. (2002). An empirical investigation of continuous-time equity return models. The Journal of Finance 57, 3, 1239–1284. Baxter, G. (1956). A strong limit theorem for Gaussian processes. Proceedings of the American Mathematical Society 7, 3, 522–527. Bayraktar, E., Poor, H.V. & Sircar, K.R. (2004). Estimating the fractal dimension of the S&P 500 index using wavelet analysis. International Journal of Theoretical and Applied Finance 7, 05, 615–643. Bender, C. & Elliott, R.J. (2004). Arbitrage in a discrete version of the Wick- fractional Black-Scholes market. Mathematics of Operations Research 29, 4, 935– 945. Bender, C., Sottinen, T. & Valkeila, E. (2007). Arbitrage with fractional Brownian motion? Theory Stoch. Process. 13, 1-2, 23–34. Bender, C., Sottinen, T., & Valkeila, E. (2008), Pricing by hedging and no-arbitrage beyond semimartingales. Finance and Stochastics. 12, 4, 441–468. Berg, L. & Lyhagen, J. (1998). Short and long-run dependence in Swedish stock returns. Applied Financial Economics 8, 4, 435–443. Biagini, F., Hu, Y., Øksendal, B. & Zhang, T. (2008). Stochastic calculus for fracti- onal Brownian motion and applications. Springer Science & Business Media. Bjo¨rk, T. & Hult, H. (2005). A note on Wick products and the fractional Black- Scholes model. Finance Stoch. 9, 2, 197–209. Black, F., & Scholes, M. (1973). The pricing of options and corporate liabilities Journal of Political Economy. 81, 3, 637–654. Cartea, A. & del Castillo-Negrete, D. (2007). Fractional diffusion models of option prices in markets with jumps. Physica A: Statistical Mechanics and its Applications 374, 2, 749–763. 24 Acta Wasaensia Cheridito, P. (2001). Mixed fractional Brownian motion. Bernoulli. 7, 6, 913–934. Chernov, M., Gallant, A.R., Ghysels, E. & Tauchen, G. (2003). Alternative models for stock price dynamics. J. Econometrics 116, 1-2, 225–257. Clark, E. & Ghosh, D.K. (2004). Arbitrage, hedging, and speculation: the foreign exchange market. Greenwood Publishing Group. Cont, R. (2005), Long range dependence in financial markets, In Fractals in engi- neering (pp. 159-179). Springer, London. Cont, R. & Tankov, P. (2004). Calibration of jump-diffusion option pricing models: a robust non-parametric approach. Dai, Q. & Singleton, K.J. (2000). Specification analysis of affine term structure models. The Journal of Finance 55, 5, 1943–1978. Duncan, T. E., Hu, Y., & Pasik-Duncan, B. (2000), Stochastic calculus for fractional Brownian motion I. Theory. Journal on Control and Optimization. 38, 2, 582–612. El-Nouty, C. (2003). The fractional mixed fractional Brownian motion. Statistics & Probability Letters 65, 2, 111–120. Eraker, B. (2004). Do stock prices and volatility jump? Reconciling evidence from spot and option prices. The Journal of Finance 59, 3, 1367–1403. Fo¨llmer, H., & Schied, A. (2011), Stochastic finance: an introduction in discrete time, Walter de Gruyter. Gladyshev, E. (1961). A new limit theorem for stochastic processes with Gaussian increments. Theory of Probability & Its Applications 6, 1, 52–61. Gylfadottir, G. (2010). Path-dependent option pricing: Efficient methods for Le´vy models. University of Florida. Heston, S.L. (1993). A closed-form solution for options with stochastic volatility with applications to bond and currency options. The Review of Financial Studies 6, 2, 327–343. Hsieh, D.A. (1991). Chaos and nonlinear dynamics: application to financial mar- kets. The Journal of Finance 46, 5, 1839–1877. Hu, Y., & Øksendal, B. (2003). Fractional white noise calculus and applications to finance. Infinite Dimensional Analysis, Quantum Probability and Related Topics. Acta Wasaensia 25 6, 01, 1–32. Huang, B.N. & Yang, C.W. (1995). The fractal structure in multinational stock returns. Applied Economics Letters 2, 3, 67–71. Hull, J. & White, A. (1987). The pricing of options on assets with stochastic vola- tilities. The Journal of Finance 42, 2, 281–300. Kallenberg, O. (2006), Foundations of modern probability, Springer Science & Business Media. Karatzas, I. & Shreve, S. (2012). Brownian motion and stochastic calculus, volume 113. Springer Science & Business Media. Kolmogorov, A.N. (1941). The local structure of turbulence in incompressible vis- cous fluid for very large Reynolds numbers. In Dokl. Akad. Nauk SSSR, volume 30. 299–303. Kou, S.G. (2002). A jump-diffusion model for option pricing. Management Science 48, 8, 1086–1101. Lamberton, D. & Lapeyre, B. (2011). Introduction to stochastic calculus applied to finance. Chapman and Hall/CRC. Lo, A. W. (1991), Long-term memory in stock market prices. Econometrica. 59, 1, 1279–1313. Ma, C. (2006). Intertemporal recursive utility and an equilibrium asset pricing model in the presence of levy jumps. Journal of Mathematical Economics 42, 2, 131–160. Mandelbrot, B. B., & Van Ness, J. W. (1968). Fractional Brownian motions, fracti- onal noises and applications. SIAM Review. 10, 4, 422–437. Marinucci, D. & Robinson, P.M. (1999). Alternative forms of fractional Brownian motion. Journal of Statistical Planning and Inference 80, 1, 111–122. Melnikov, A. (1999). Financial Markets: stochastic analysis and the pricing of derivative securities. Meng, L. & Wang, M. (2010). Comparison of Black–Scholes formula with fracti- onal Black–Scholes formula in the foreign exchange option market with changing volatility. Asia-Pacific Financial Markets 17, 2, 99–111. 26 Acta Wasaensia Merton, R.C. (1976). Option pricing when underlying stock returns are discontinu- ous. Journal of Financial Economics 3, 1-2, 125–144. Mikosch, T. (1998). Elementary stochastic calculus—with finance in view, volume 6 of Advanced Series on Statistical Science & Applied Probability. World Scientific Publishing Co., Inc., River Edge, NJ. Mishura, Y.S. (2008). Stochastic calculus for fractional Brownian motion and re- lated processes, volume 1929 of Lecture Notes in Mathematics. Springer-Verlag, Berlin. Mishura, Y.S. & Valkeila, E. (2002). The absence of arbitrage in a mixed Brownian– fractional Brownian model. Trudy Matematicheskogo Instituta imeni VA Steklova 237, 224–233. Molchan, G.M. & Golosov, Y.I. (1969). Gaussian stationary processes with asymp- totically a power spectrum. In Doklady Akademii Nauk, volume 184. Russian Aca- demy of Sciences, 546–549. Nualart, D. (2006). Fractional Brownian motion: stochastic calculus and applicati- ons. In International Congress of Mathematicians. Vol. III. Eur. Math. Soc., Zu¨rich, 1541–1562. Øksendal, B. (2003). Stochastic differential equations. Universitext, sixth edition. Springer-Verlag, Berlin. An introduction with applications. Pan, J. (2002). The jump-risk premia implicit in options: Evidence from an integra- ted time-series study. Journal of Financial Economics 63, 1, 3–50. Pascucci, A. (2011) PDE and martingale methods in option pricing, Springer Science & Business Media. Pliska, S. (1997). Introduction to mathematical finance. Blackwell publishers Ox- ford. Seydel, R.U. (2017). Tools for computational finance. Universitext, sixth edition. Springer-Verlag, London. Shiryaev, A. N. (1999), Essentials of stochastic finance: facts, models, theory, World Scientific. Shiryaev, A. N., do Rosa´rio Grossinho, M., Oliveira, P. E., & Esquı´vel, M. L. (2006), Stochastic finance, Springer Science & Business Media. Acta Wasaensia 27 Sottinen, T., & Viitasaari, L. (2016), Stochastic analysis of Gaussian processes via Fredholm representation. International Journal of Stochastic Analysis. 2016. Sottinen, T. & Valkeila, E. (2003). On arbitrage and replication in the fractional Black-Scholes pricing model. Statist. Decisions 21, 2, 93–107. Sottinen, T.P. (2003). Fractional Brownian motion in finance and queueing. Pro- Quest LLC, Ann Arbor, MI. Thesis (Ph.D.)–Helsingin Yliopisto (Finland). Tong, Z. (2012). Option pricing with long memory stochastic volatility models. Ph.D. thesis, Universite´ d’Ottawa/University of Ottawa. van Zanten, H. (2007), When is a linear combination of independent fBm’s equiva- lent to a single fBm?. Stochastic Processes and their Applications. 117, 1, 57-70. Wang, X.T. (2010). Scaling and long range dependence in option pricing, IV: pricing European options with transaction costs under the multifractional Black- Scholes model. Phys. A 389, 4, 789–796. Wang, X.T., Wu, M., Zhou, Z.M. & Jing, W.S. (2012). Pricing European option with transaction costs under the fractional long memory stochastic volatility model. Physica A: Statistical Mechanics and its Applications 391, 4, 1469–1480. Wang, X.T., Zhu, E.H., Tang, M.M. & Yan, H.G. (2010). Scaling and long-range dependence in option pricing II: pricing European option with transaction costs under the mixed Brownian-fractional Brownian model. Phys. A 389, 3, 445–451. Willinger, W., Taqqu, M. S., & Teverovsky, V. (1999), Stock market prices and long-range dependence. Finance and Stochastics. 3, 1, 1–13. Xiao, W.L., Zhang, W.G., Zhang, X.L. & Wang, Y.L. (2010). Pricing currency options in a fractional Brownian motion with jumps. Economic Modelling 27, 5, 935–942. Yaglom, A.M. (1955). Correlation theory of processes with random stationary n th increments. Matematicheskii Sbornik 79, 1, 141–196. Zhang, W.G., Xiao, W.L. & He, C.X. (2009). Equity warrants pricing model under fractional Brownian motion and an empirical study. Expert Systems with Applicati- ons 36, 2, 3056–3065. Zili, M. (2006). On the mixed fractional Brownian motion. J. Appl. Math. Stoch. Anal. , Art. ID 32435, Journal of Computational and Applied Mathematics 344 (2018) 716–724 Contents lists available at ScienceDirect Journal of Computational and Applied Mathematics journal homepage: www.elsevier.com/locate/cam The evaluation of geometric Asian power options under time changed mixed fractional Brownian motion Foad Shokrollahi Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700, FIN-65101 Vaasa, Finland a r t i c l e i n f o Article history: Received 14 December 2017 Received in revised form 17 April 2018 MSC: 91G20 91G80 60G22 Keywords: Mixed fractional Brownian motion Geometric Asian option Power option Time changed process a b s t r a c t In this paper, the geometric Asian option pricing problem is investigated under the assump- tion that the underlying stock price is assumed following a mixed fractional subdiffusive Black–Scholes model, and the geometric average Asian option pricing formula is derived under this assumption. We then apply the results to value Asian power options on the stocks that pay constant dividends when the payoff is a power function. Finally, lower bound of Asian options and some special cases are provided. © 2018 Elsevier B.V. All rights reserved. 1. Introduction A standard option is a financial contract which gives the owner of the contract the right, but not the obligation, to buy or sell a specified asset at a prespecified time (maturity) for a prespecified price (strike price). The specified asset (underlying asset) can be for example stocks, indexes, currencies, bonds or commodities. The option can be either a call option, which gives the owner the right to buy the underlying asset, or it can be a put option, which gives the owner the right to sell the underlying asset. There are several types of options that are traded in a market. American option allows the owner to exercise his option at any time up to and including the strike date. European options can be exercised only on the strike date. European options are also called vanilla options. Their payoffs at maturity depend on the spot value of the stock at the time of exercise. There are other options whose values depend on the stock prices over a predetermined time interval. For an Asian option, the payoff is determined by the average value over some predetermined time interval. The average price of the underlying asset can either determine the underlying settlement price (average price Asian options) or the option strike price (average strike Asian options). Furthermore, the average prices can be calculated using either the arithmetic mean or the geometric mean. The type of Asian option that will be examined throughout this research is geometric Asian option. Over the past three decades, academic researchers and market practitioners have developed and adopted different models and techniques for option valuation. Themost popular model on option pricing was introduced by Black and Scholes (BS) [1] in 1973. In the BS model it has been assumed that the asset price dynamics are governed by a geometric Brownian motion. However, a large number of empirical studies have shown that the distributions of the logarithmic returns of financial asset usually exhibit properties of self-similarity, heavy tails, long-range dependence in both auto-correlations and cross-correlations, and volatility clustering [2–4]. Actually, the most popular stochastic process that exhibits long-range E-mail address: foad.shokrollahi@uva.fi. https://doi.org/10.1016/j.cam.2018.05.042 0377-0427/© 2018 Elsevier B.V. All rights reserved. 28 Acta Wasaensia F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 717 dependence is of course the fractional Brownian motion. Moreover, the fractional Brownian motion produces a burstiness in the sample path behavior, which is the important behavior of financial time series. Since fractional Brownian motion is neither aMarkov process nor a semi-martingale, we cannot use the usual stochastic calculus to analyze it. Further, fractional Brownianmotion admits arbitrage in a complete and frictionlessmarket. To get around this problem and to take into account the long memory property, it is reasonable to use the mixed fractional Brownian motion (mfBm) to capture the fluctuations of the financial asset [5–7]. The mfBm is a linear combination of the Brownian motion and fractional Brownian motion with Hurst index H ∈ ( 12 , 1), defined on the filtered probability (Ω,F ,P) for any t ∈ R+ by: MHt (a, b) = aB(t) + bBH (t), (1.1) where B(t) is a Brownian motion, and BH (t) is an independent fractional Brownian motion with Hurst index H . Cheridito [7] proved that, for H ∈ ( 34 , 1), the mixed model is equivalent to the Brownian motion and hence it is also arbitrage free. For H ∈ ( 12 , 1), Mishura and Valkeila [8] demonstrated that the mixedmodel is arbitrage free. Rao [9] discussed geometric Asian power option undermfBm. To see more about the mixed model, one can refer to Refs. [6,7,10,11]. Analysis of various real-life data shows thatmanyprocesses observed in economics display characteristic periods inwhich they stay motionless [12]. This feature is most common for emerging markets in which the number of participants, and thus the number of transactions, is rather low. Notably, similar behavior is observed in physical systems exhibiting subdiffusion. The constant periods of financial processes correspond to the trapping events in which the subdiffusive test particle gets immobilized [13]. Subdiffusion is a well known and established phenomenon in statistical physics. Its usual mathematical description is in terms of the celebrated Fractional Fokker–Planck equation (FFPE). This equation was first derived from the continuous-time random walk scheme with heavy-tailed waiting times [14,15,10], and since then became fundamental in modeling and analysis of complex systems exhibiting slow dynamics. Following this line, and to model the observed long range dependence and fluctuations in the financial price time series, we introduce a time-changedmixed fractional BSmodel to value Asian power option when the underlying stock is St = X(Tα(t)) = S0e(r−q)Tα (t)+M H α (t)(σ ,σ )− 12 σ2 t α Γ (α+1)− 12 σ2 ( tα Γ (α+1) )2H , S0 = X(0) > 0, (1.2) whereMHα (t)(σ , σ ) = σB(Tα(t)) + σBH (Tα(t)), α ∈ ( 12 , 1), 2α − αH > 1 and Tα(t) is the inverse α-stable subordinator. We then apply the result to price geometric Asian power options that pay constant dividends when the payoff is a power function. We also provide some special cases and lower bound for the Asian option price. The rest of the paper is organized as follows. In Section 2, some useful concepts and theorems of time changed mixed fractional process are introduced. In Section 3, a brief introduction of Asian options is given. Analytical valuation formula for geometric Asian options is derived in Section 4 and then applied to geometric Asian power options in Section 5. The lower bound on the price of the Asian option is proposed in Section 6. 2. Auxiliary facts In this section, we recall some definitions and results about mixed fractional time changed process. More information about mixed fractional process can be found in [16,10]. The time-changed process Tα(t) is the inverse α-stable subordinator defined as below Tα(t) = inf{τ > 0,Uα(t) ≥ t} here Uα(τ )τ≥0 is a strictly increasing α-stable Lévy process [17] with Laplace transform: E(e−uUα (τ )) = e−τuα , α ∈ (0, 1). Uα(t) is 1α self-similar and Tα(t) is α self-similar, that is, for every h > 0, Uα(ht)  h 1 α Uα(t) Tα(ht)  hαTα(t), here  indicates that the random variables on both sides have the same distribution. Specially, when α ↑ 1, Tα(t) reduces to the physical time t . You can find more details about subordinator and its inverse processes in [18,19]. Consider the subdiffusion process MHα (t)(a, b) = aWα(t) + bWHα (t) = aB(Tα(t)) + bBH (Tα(t)), where B(τ ) is a Brownian motion, BH (τ ) is a fractional Brownian motion with Hurst index H and Tα(t) is inverse α- subordinator which are supposed to be independent. When a = 0, b = 1, then it is the process considered in [20] and if b = 0, a = 1, then it is the process considered in [21]. In this research, we assume that H ∈ ( 34 , 1) and (a, b) = (1, 1). Remark 2.1. When α ↑ 1, the processes Wα(t) and WHα (t) degenerate to B(t) and BH (t), respectively. Then, MHα (t)(a, b) reduces to themfBm in Eq. (1.1). Acta Wasaensia 29 718 F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 Remark 2.2. From [20,21], we know that E(Tα(t)) = tαΓ (α+1) . Then, by applying α-self-similar and non-decreasing sample path of Tα(t), we have E[(B(Tα(t)))2] = t α Γ (α + 1) (2.1) E[(BH (Tα(t)))2] = ( tα Γ (α + 1) )2H . (2.2) 3. Asian options The payoff of an Asian option is based on the difference between an asset’s average price over a given time period, and a fixed price called the strike price. Asian options are popular because they tend to have lower volatility than options whose payoffs are based purely on a single price point. It is also harder for big traders to manipulate an average price over an extended period than a single price, so Asian option offers further protection against risk. The Asian call and put options have a payoff that is calculated with an average value of the underlying asset over a specific period. The payoff for an Asian call and put option with strike price K and expiration time T is (S¯(T ) − K )+ and (K − S¯(T ))+ respectively, where S¯(T ) is the average price of the underlying asset over the prespecified interval. Since Asian options are less expensive than their European counterparts, they are attractive to many different investors. Apart from the regular Asian option there also exists Asian strike option. An Asian strike call option guarantees the holder that the average price of an underlying asset is not higher than the final price. The option will not be exercised if the average price of the underlying asset is greater than the final price. The holder of an Asian strike put optionmakes sure that the average price received for the underlying asset is not less thanwhat the final pricewill provide. The payoff for anAsian strike call and put option is (S¯(T )−S(T ))+ and (S(T )−S¯(T ))+ respectively, where S(T ) is the value of underlying stock at maturity date T . Asian options are divided into two different types, when calculating the average, the geometric Asian option G(T ) = exp { 1 T ∫ T 0 ln S(t)dt } , and the arithmetic Asian option. A(T ) = 1 T ∫ T 0 S(t)dt. We assume that the prespecified interval [0, T ] is fixed, thenwill price the geometric Asian option in the continuous average case under time changed mixed fractional Brownian motion environment. 4. Pricing model of geometric Asian option In order to derive an Asian option pricing formula in a time changed mixed fractional market, we make the following assumptions: (i) The price of underlying stock at time t is given by Eq. (1.2). (ii) There are no transaction costs in buying or selling the stocks or option. (iii) The risk free interest rate r and dividend rate q are known and constant through time. (iv) The option can be exercised only at the maturity time. From Eq. (1.2), we know that ln St  N(u, v), where u = ln S(0) + (r − q)Tα(t) − 12σ 2 t α Γ (α + 1) − 1 2 σ 2 ( tα Γ (α + 1) )2H (4.1) v = σ 2 t α Γ (α + 1) + σ 2 ( tα Γ (α + 1) )2H . (4.2) Let C(S(0), T ) be the price of a European call option at time 0 with strike price K and that matures at time T . Then, from [16], we can get C(S(0), T ) = S(0)e−qTφ(d1) − Ke−rTφ(d2), where d1 = ln S0 K + (r − q + σˆ 2 2 )T σˆ √ T , d2 = d1 − σˆ √ T , 30 Acta Wasaensia F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 719 σˆ 2 = σ 2 T α−1 Γ (α) + σ 2 ( Tα−1 Γ (α) )2H , and φ(.) denotes cumulative normal density function. Under the above assumptions (i)–(iv), we obtain the value of the geometric Asian call option by the following theorem Theorem4.1. Suppose the stock price St satisfied Eq. (1.2). Then, under the risk-neutral probabilitymeasure, the value of geometric Asian call option C(S(0), T ) with strike price K and maturity time T is given by C(S(0), T ) = S(0) exp { −rT + (r − q) T α Γ (α + 2) + σ 2(−T )α 2Γ (α + 3) − σ 2T 2αH 4(2αH + 1)(αH + 1)(Γ (α + 1))2H } φ(d1) − Ke−qTφ(d2), (4.3) where d2 = μG − ln K σG , d1 = d2 + σG, μG = ln S(0) + (r − q − σ 2 2 ) Tα Γ (α + 2) − σ 2T 2αH 2(2αH + 1)(Γ (α + 1))2H , σ 2G = σ 2Tα Γ (α + 2) + σ 2(−T )α Γ (α + 3) + σ 2T 2αH (2αH + 2)(Γ (α + 1))2H , the interest rate r and the dividend rate q are constant over time and φ(.) denotes cumulative normal density function. Proof. Suppose L(T ) = 1 T ∫ T 0 ln S(t)dt. Then G(T ) = eL(T ). (4.4) We know that ln St  N(u, v), then it is clear that the random variable L(T ) has Gaussian distribution under the risk-neutral probability measure. We will now compute its mean and variance under the risk-neutral probability measure. Let E denote the expectation and,μG and σ 2G denote themean and the variance of the randomvariableE under the risk-neutral probability measure. Note that μG = E[L(T )] = 1T ∫ T 0 E[ln S(t)]dt = ln S(0) + 1 T ∫ T 0 (r − q) t α Γ (α + 1)dt − σ 2 2T ∫ T 0 [ tα Γ (α + 1) + t2αH (Γ (α + 1))2H ] dt = ln S(0) + (r − q) T α Γ (α + 2) − σ 2Tα 2Γ (α + 2) − σ 2T 2αH (4αH + 2)(Γ (α + 1))2H , and σ 2G = Var[L(T )] = E[(L(T ) − μG)2] = σ 2 T 2 ∫ T 0 ∫ T 0 ( E[Wα(t)Wα(τ )] + E[WHα (t)WHα (τ )] ) dtdτ , by independence of the processes B(t), BH (t) and Tα(t), we obtain = σ 2 T 2 ∫ T 0 ∫ T 0 ( | t α Γ (α + 1) | + | τα Γ (α + 1) | − | (t − τ )α Γ (α + 1) | ) dtdτ + σ 2 T 2 ∫ T 0 ∫ T 0 ( | t α Γ (α + 1) | 2H + | τ α Γ (α + 1) | 2H − | (t − τ ) α Γ (α + 1) | 2H) dtdτ = σ 2Tα Γ (α + 2) + σ 2(−T )α Γ (α + 3) + σ 2T 2αH (2αH + 2)(Γ (α + 1))2H . From (4.4), we know that the random variable G(T ) is log-normally distributed, then lnG(T )  N(μG, σ 2G ). Let I = {x : ex > K } and φ(.) be the probability density function of a standard normal distribution, then the price of geometric Asian call Acta Wasaensia 31 720 F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 option is given by the following computations C(S(0), T ) = e−rTE[(G(T ) − K )+] = e−rT ∫ I (ex − K ) 1√ 2πσG exp { − (x − μG) 2 2σ 2G } dx = e−rT ∫ I (eμG+zσG − K ) 1√ 2πσG exp { − (x − μG) 2 2σ 2G } ϕ(z)dz = e−rT+μG+ 12 σ2G ∫ ∞ −d2 e− 1 2 (z−σG)2dz − Ke−rT ∫ −∞ −d2 ϕ(z)dz = e−rT+μG+ 12 σ2G ∫ ∞ −d2−σG ϕ(z)dz − Ke−rT ∫ d2 −∞ ϕ(z)dz = e−rT+μG+ 12 σ2G ∫ d2+σG −∞ ϕ(z)dz − Ke−rT ∫ d2 −∞ ϕ(z)dz = e−rT+μG+ 12 σ2Gφ(d1) − Ke−rTφ(d2), = S(0) exp { −rT + (r − q) T α Γ (α + 2) + σ 2 (−T )α 2Γ (α + 3) − σ 2 T 2αH 4(2αH + 1)(αH + 1)(Γ (α + 1))2H } φ(d1) − Ke−qTφ(d2), here I = {x : ex > K } = {z : eμG+zσG > K } = {z : μG + zσG > ln K } = {z : z > −d2}, thus we obtain the pricing formula.  Moreover, using the put–call parity, the valuation model for a geometric Asian put option under time changed mixed fractional BS model can be written P(S(0), T ) = Ke−qTφ(−d2) − S(0) exp { −rT + (r − q) T α Γ (α + 2) + σ 2(−T )α 2Γ (α + 3) − σ 2T 2αH 4(2αH + 1)(αH + 1)(Γ (α + 1))2H } φ(−d1), (4.5) where d1 and d2 are defined previously. Letting α ↑ 1, then the stock price follows themfBm shown below St = S0 exp { (r − q)T + σB(t) + σBH (t) − 1 2 σ 2t − 1 2 σ 2t2H } , 0 < t < T , (4.6) and the result is presented below. Corollary 4.1. The value of geometric Asian call option with maturity T and strike K , whose stock price follows Eq. (4.6), is given by C(S(0), T ) = S(0) exp { −1 2 (r + q)T − σ 2T 12 − σ 2T 2H 4(2H + 1)(H + 1) } φ(d1) − Ke−qTφ(d2), (4.7) where d2 = μG − ln K σG , d1 = d2 + σG, μG = ln S(0) + 12 (r − q − σ 2 2 )T − σ 2T 2H 2(2H + 1) , 32 Acta Wasaensia F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 721 σ 2G = σ 2T 3 + σ 2T 2H (2H + 2) , which is consistent with result in [9]. 5. Pricing model of Asian power option In this section, we consider the pricing model of Asian power call option with strike price K and maturity time T under time changed mixed fractional BS model where the payoff function is (Gn(T ) − K )+ for some constant integer n ≥ 1. Theorem5.1. Suppose the stock price St satisfied Eq. (1.2). Then, under the risk-neutral probabilitymeasure the value of geometric Asian power call option C(S(0), T ) with strike price K , maturity time T and payoff function (Gn(T ) − K )+ is given by C(S(0), T ) = S(0) exp { −rT + (r − q) nT α Γ (α + 2) − (n − n2)σ 2Tα 2Γ (α + 2) + n2σ 2(−T )α 2Γ (α + 3) − nσ 2T 2αH (4αH + 2)(Γ (α + 1))2H − n2σ 2T 2αH (4αH + 4)(Γ (α + 1))2H } φ(f1) − Ke−qTφ(f2), (5.1) where f2 = μG − 1 n ln K σG , f1 = f2 + nσG, μG = ln S(0) + (r − q − σ 2 2 ) Tα Γ (α + 2) − σ 2T 2αH 2(2αH + 1)(Γ (α + 1))2H , σ 2G = σ 2Tα Γ (α + 2) + σ 2(−T )α Γ (α + 3) + σ 2T 2αH (2αH + 2)(Γ (α + 1))2H , the interest rate r and the dividend rate q are constant over time and ϕ(.) denotes cumulative normal density function. Proof. The payoff function for Asian power option is (Gn(T ) − K )+ = (enL(T ) − K )+, then applying similar computation in Theorem 4.1, we obtain C(S(0), T ) = e−rTE[(Gn(T ) − K )+] = e−rT ∫ I (enx − K ) 1√ 2πσG exp { − (x − μG) 2 2σ 2G } dx = e−rT ∫ I (en(μG+zσG) − K ) 1√ 2πσG exp { − (x − μG) 2 2σ 2G } ϕ(z)dz = e−rT+nμG+ 12 n2σ2G ∫ ∞ −f2 e− 1 2 (z−nσG)2dz − Ke−rT ∫ −∞ −f2 ϕ(z)dz = e−rT+nμG+ 12 n2σ2G ∫ ∞ −f2−nσG ϕ(z)dz − Ke−rT ∫ f2 −∞ ϕ(z)dz = e−rT+nμG+ 12 n2σ2G ∫ f2+nσG −∞ ϕ(z)dz − Ke−rT ∫ f2 −∞ ϕ(z)dz = e−rT+nμG+ 12 n2σ2Gφ(f1) − Ke−rTφ(f2), = S(0) exp { −rT + (r − q) nT α Γ (α + 2) − (n − n2)σ 2Tα 2Γ (α + 2) + n2σ 2(−T )α 2Γ (α + 3) − nσ 2T 2αH (4αH + 2)(Γ (α + 1))2H − n2σ 2T 2αH (4αH + 4)(Γ (α + 1))2H } φ(f1) − Ke−qTφ(f2), Acta Wasaensia 33 722 F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 here I = {x : enx > K } = {z : en(μG+zσG) > K } = {z : μG + zσG > 1n ln K } = {z : z > −f2}, thus the proof is completed.  The time changed mixed fractional BS model includes the jump behavior of price process because the subordinator process is a pure jump Levy process so it can capture the random variations in volatility. Also, it can be used for data with long range dependence and visible constant time periods characteristic for processes delayed by inverse subordinators. 6. Lower bound of the Asian option price The aim of this section is to obtain the lower bound on the price of the Asian option. The next theorem shows that the normal distribution is stable when the random variables are jointly normal. Theorem 6.1 ([22]). The conditional distribution of ln Sti given lnG(T ) is a normal distribution (ln Sti | lnG(T ) = z)  N(μi + (z − μG) λi σ 2G , σ 2i − λ2i σ 2G ), i = 1, . . . , n, where μi = ln S(0) + (r − q)Tα(ti) − 12σ 2 t α i Γ (α + 1) − 1 2 σ 2 ( tαi Γ (α + 1) )2H σ 2i = σ 2 tαi Γ (α + 1) + σ 2 ( tαi Γ (α + 1) )2H , λi = Cov(ln Sti , lnG(T )), 0 ≤ t1 < t2 < · · · < tn ≤ T , Tα(t) is inverse α-stable subordinator and, μG and σ 2G are defined in Theorem 4.1. Moreover, (Sti | lnG(T )) has a lognormal distribution and E [ Sti | lnG(T ) = z ] = exp { μi + (z − μG) λi σ 2G + 1 2 (σ 2i − λ2i σ 2G ) } i = 1, . . . , n. (6.1) Now, we condition on the geometric average G(T ) in the pricing expression of the Asian option C(S(0), T ) = e−rTE[(A(T ) − K )+] = e−rTE[E[(A(T ) − K )+|G(T )]] = e−rT ∫ ∞ 0 E[(A(T ) − K )+|G(T ) = z]g(z)dz, where g is the lognormal density function of G. Let C1 = ∫ K 0 E[(A(T ) − K )+|G(T ) = z]g(z)dz, C2 = ∫ ∞ K E[(A(T ) − K )+|G(T ) = z]g(z)dz, then C(S(0), T ) = e−rT (C1 + C2). Since the geometric average is less than arithmetic average A(T ) ≥ G(T ), C2 = ∫ ∞ K E[A(T ) − K |G(T ) = z]g(z)dz, (6.2) from Theorem 6.1, we can calculate C2. Applying Jensen’s inequality we obtain a lower bound on C1 C1 = ∫ K 0 E[(A(T ) − K )+|G(T ) = z]g(z)dz ≥ ∫ K 0 (E[A(T ) − K |G(T ) = z])+g(z)dz = ∫ K K˜ E[A(T ) − K |G(T ) = z]g(z)dz = C˜1 (6.3) where K˜ = {z|E[A(T )|G(T ) = z] = K }. 34 Acta Wasaensia F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 723 Eq. (6.1) enables us to obtain K˜ , then we calculate the following expectation E[A(T )|G(T ) = z] = E [ 1 n n∑ i=1 Sti |G(T ) = z ] = 1 n n∑ i=1 E [ Sti |G(T ) = z ] = 1 n n∑ i=1 exp ( μi + (log z − μG) λi σ 2G + 1 2 (σ 2i − λ2i σ 2G ) ) . Theorem 6.2. A lower bound on the price of the Asian option with strike price K and maturity time T is given by C˜(S(0), T ) = e−rT (C˜1 + C2) = e−rT {1 n n∑ i=1 exp(μi + 12σ 2 i )φ ( μG − ln K˜ + γi σG ) − Kφ ( μG − ln K˜ σG )} , where all parameters are defined previously. Proof. Collecting Eqs. (6.2) and (6.3) gives C˜1 + C2 = ∫ ∞ K˜ E[A(T ) − K |G(T ) = z]g(z)dz = ∫ ∞ K˜ E[A(T )|G(T ) = z]g(z)dz − K ∫ ∞ K˜ g(z)dz = ∫ ∞ K˜ E [ 1 n n∑ i=1 Sti |G(T ) = z ] g(z)dz − K ∫ ∞ K˜ g(z)dz = ∫ ∞ K˜ 1 n n∑ i=1 E [ Sti |G(T ) = z ] g(z)dz − K ∫ ∞ K˜ g(z)dz = 1 n n∑ i=1 ∫ ∞ K˜ E [ Sti | lnG(T ) = ln z ] g(z)dz − K ∫ ∞ K˜ g(z)dz. From the proof of Theorem 4.1, we obtain K ∫ ∞ K˜ g(z)dz = Kφ ( μG − ln K˜ σG ) , and from Eq. (6.1)∫ ∞ K˜ E [ Sti | lnG(T ) = ln z ] g(z)dz = ∫ ∞ K˜ exp ( μi + (ln z − μG) λi σ 2G + 1 2 (σ 2i − λ2i σ 2G ) ) g(z)dz = exp ( μi + 12σ 2 i )∫ ∞ K˜ exp ( (ln z − μG) λi σ 2G − 1 2 λ2i σ 2G ) g(z)dz. Using the density of the lognormal distribution, we have∫ ∞ K˜ 1√ 2πσGz exp ( (ln z − μG) λi σ 2G − 1 2 λ2i σ 2G − 1 2 ( μG − ln z σG )2 ) dz. Making the change of variables y = μG−ln z+λi σG and dydz = − 1σGz , then we have∫ −∞ μG−ln z+λi σG − 1√ 2π exp ( ( λi σG − y) λi σG − 1 2 λ2i σ 2G − 1 2 (y − λi σG )2 ) dy = ∫ μG−ln z+λi σG −∞ 1√ 2π exp ( −y λi σG + 1 2 λ2i σ 2G − 1 2 y2 − 1 2 λ2i σ 2G + y λi σG ) dy = ∫ μG−ln z+λi σG −∞ 1√ 2π exp ( −1 2 y2 ) dy = φ ( μG − ln K˜ + γi σG ) , by collecting C˜1 and C2 the proof is completed.  Acta Wasaensia 35 724 F. Shokrollahi / Journal of Computational and Applied Mathematics 344 (2018) 716–724 Acknowledgments This paper is supported by the University of Vaasa, Finland. The author is deeply grateful to the anonymous referees for reading the present paper carefully and giving very helpful comments which contributed to improving the quality of the paper. References [1] F. Black, M. Scholes, The pricing of options and corporate liabilities, J. Polit. Econ. 81 (3) (1973) 637–654. [2] X.-T.Wang,M.Wu, Z.-M. Zhou,W.-S. Jing, Pricing European optionwith transaction costs under the fractional longmemory stochastic volatilitymodel, Physica A 391 (4) (2012) 1469–1480. [3] W.-G. Zhang, W.-L. Xiao, C.-X. He, Equity warrants pricing model under fractional Brownian motion and an empirical study, Expert Syst. Appl. 36 (2) (2009) 3056–3065. [4] C. Necula, Option pricing in a fractional Brownian motion environment, 2002. [5] C. El-Nouty, The fractional mixed fractional Brownian motion, Statist. Probab. Lett. 65 (2) (2003) 111–120. [6] I.S. Mishura, Y. Mishura, Stochastic Calculus for Fractional Brownian Motion and Related Processes, Vol. 1929, Springer Science & Business Media, 2008. [7] P. Cheridito, et al., Mixed fractional Brownian motion, Bernoulli 7 (6) (2001) 913–934. [8] Y.S. Mishura, E. Valkeila, The absence of arbitrage in a mixed Brownian–fractional Brownian model, Tr. Mat. Inst. Steklova 237 (2002) 224–233. [9] B.P. Rao, Pricing geometric Asian power options under mixed fractional Brownian motion environment, Physica A 446 (2016) 92–99. [10] F. Shokrollahi, A. Kılıçman, M. Magdziarz, Pricing European options and currency options by time changed mixed fractional Brownian motion with transaction costs, Int. J. Financ. Eng. 3 (01) (2016) 1650003. [11] F. Shokrollahi, A. Kılıçman, Pricing currency option in a mixed fractional brownian motion with jumps environment, Math. Probl. Eng. 2014 (2014). [12] J. Janczura, A. Wyłomańska, Subdynamics of financial data from fractional fokker-planck equation., Acta Phys. Pol. B 40 (5) (2009) 1341–1351. [13] I. Eliazar, J. Klafter, Spatial gliding, temporal trapping, and anomalous transport, Physica D 187 (1) (2004) 30–50. [14] E. Barkai, R. Metzler, J. Klafter, From continuous time random walks to the fractional Fokker-Planck equation, Phys. Rev. E 61 (1) (2000) 132. [15] R. Metzler, J. Klafter, The random walk’s guide to anomalous diffusion: a fractional dynamics approach, Phys. Rep. 339 (1) (2000) 1–77. [16] Z. Guo, H. Yuan, Pricing European option under the time-changed mixed Brownian–fractional Brownian model, Physica A 406 (2014) 73–79. [17] K.-i. Sato, Lévy Processes and Infinitely Divisible Distributions, Cambridge University Press, 1999. [18] A. Janicki, A. Weron, Simulation and Chaotic Behavior of Alpha-stable Stochastic Processes, Vol. 178, CRC Press, 1993. [19] A. Piryatinska, A. Saichev, W. Woyczynski, Models of anomalous diffusion: the subdiffusive case, Physica A 349 (3) (2005) 375–420. [20] H. Gu, J.-R. Liang, Y.-X. Zhang, Time-changed geometric fractional Brownian motion and option pricing with transaction costs, Physica A 391 (15) (2012) 3971–3977. [21] M. Magdziarz, Black–Scholes formula in subdiffusive regime, J. Stat. Phys. 136 (3) (2009) 553–564. [22] J. Hoffman-Jorgensen, Probability with a View Towards Statistics, Vol. 2, CRC Press, 1994. 36 Acta Wasaensia Statistics and Probability Letters 130 (2017) 85–91 Contents lists available at ScienceDirect Statistics and Probability Letters journal homepage: www.elsevier.com/locate/stapro Hedging in fractional Black–Scholes model with transaction costs Foad Shokrollahi *, Tommi Sottinen Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700, FIN-65101 Vaasa, Finland a r t i c l e i n f o Article history: Received 11 June 2017 Received in revised form 23 July 2017 Accepted 25 July 2017 Available online 4 August 2017 MSC: 91G20 91G80 91G22 Keywords: Delta-hedging Fractional Black–Scholes model Transaction costs Option pricing a b s t r a c t We consider conditional-mean hedging in a fractional Black–Scholes pricing model in the presence of proportional transaction costs. We develop an explicit formula for the conditional-mean hedging portfolio in terms of the recently discovered explicit conditional law of the fractional Brownian motion. © 2017 Elsevier B.V. All rights reserved. 1. Introduction We consider discrete hedging in fractional Black–Scholes models where the asset price is driven by a long-range dependent fractional Brownian motion. For a convex or concave European vanilla type option, we construct the so- called conditional-mean hedge. This means that at each trading time the value of the conditional mean of the discrete hedging strategy coincides with the frictionless price. By frictionless wemean the continuous trading hedging price without transaction costs. The key ingredient in constructing the conditional mean hedging strategy is the recent representation for the regular conditional law of the fractional Brownian motion given in Sottinen and Viitasaari (2017). Let us note that there are arbitrage strategies with continuous trading without transaction costs, but not with discrete trading strategies, even in the absence of trading costs. For details of the use of fractional Brownian motion in finance and discussion on arbitrage see Bender et al. (2011). For the classical Black–Scholes model driven by the Brownian motion, the study of hedging under transaction costs goes back to Leland (1985). See also Denis and Kabanov (2010) and Kabanov and Safarian (2009) for a mathematically rigorous treatment. For the fractional Black–Scholes model driven by the long-range dependent fractional Brownian motion, the study of hedging under transaction costs was studied in Azmoodeh (2013). In the series of articles (Shokrollahi et al., 2016; Wang, 2010a, b; Wang et al., 2010a, b) the discrete hedging in the fractional Black–Scholes model was studied by using the economically dubious Wick–Itô–Skorohod interpretation of the self-financing condition. Actually, with the economically solid forward-type pathwise interpretation of the self-financing condition, these hedging strategies are valid, not for the geometric fractional Brownianmotion, but for a geometric Gaussian processwhere the driving noise is a Gaussianmartingale * Corresponding author. E-mail addresses: foad.shokrollahi@uva.fi (F. Shokrollahi), tommi.sottinen@iki.fi (T. Sottinen). http://dx.doi.org/10.1016/j.spl.2017.07.014 0167-7152/© 2017 Elsevier B.V. All rights reserved. Acta Wasaensia 37 86 F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 with the same variance function as the corresponding fractional Brownian motion would have, see Gapeev et al. (2011), Shokrollahi and Kılıçman (2014), Shokrollahi and Kılıçman (2015) and Shokrollahi et al. (2015). 2. Preliminaries We are interested in pricing of European vanilla options f (ST ) of a single underlying asset S = (St )t∈[0,T ], where T > 0 is a fixed time of maturity of the option. We consider the discounted fractional Black–Scholes pricingmodel where the ‘‘riskless’’ investment, or the bond, is taken as the numéraire and the risky asset S = (St )t∈[0,T ] is given by the dynamics dSt St = μ dt + σ dBt , (2.1) where B is the fractional Brownian motion with Hurst index H ∈ ( 12 , 1). Recall that, qualitatively, the fractional Brownian motion is the (up to amultiplicative constant) unique Gaussian process with stationary increments and self-similarity index H . Quantitatively, the fractional Brownian motion is defined by its covariance function r(t, s) = 1 2 [ t2H + s2H − |t − s|2H] . Since the fractional Brownian motion with index H ∈ ( 12 , 1) has zero quadratic variation, the classical change-of-variables rule applies. Consequently, the pathwise solution to the stochastic differential equation (2.1) is St = S0eμt+σBt . (2.2) Also, it follows from the classical change-of-variables rule that f (ST ) = f (S0) + ∫ T 0 f ′(St ) dSt , (2.3) where f is a convex or concave function and f ′ is its left-derivative. We refer to Azmoodeh et al. (2009) for details. The economic interpretation of (2.3) is that under continuous trading and no transaction costs, the replication price of a European vanilla option f (ST ) is f (S0) and the replicating strategy is given by πt = f ′(St ), where πt denotes the number of the shares of the risky asset S held by the investor at time t . Furthermore, we note that the value Vπ of the hedging strategy π = f ′(S·) at time t is Vπt = Vπ0 + ∫ t 0 πu dSu = f (S0) + ∫ t 0 f ′(Su) dSu = f (St ). Indeed, the first equality is simply the self-financing condition and the rest follows immediately from (2.3). Note that this is very different from the value in the classical Black–Scholes model, where the value is determined by the Black–Scholes partial differential equation, which in turn comes to the Itô’s change-of-variables rule. We assume that the trading only takes place at fixed preset time points 0 = t0 < t1 < · · · < tN = T . We denote by πN the discrete trading strategy πNt = πN0 1{0}(t) + N∑ i=1 πNti−11(ti−1,ti](t). The value of the strategy πN is given by Vπ N ,k t = Vπ N ,k 0 + ∫ t 0 πNu dSu − ∫ t 0 kSu|dπNu |, (2.4) where k ∈ [0, 1) is the proportional transaction cost. Under transaction costs perfect hedging is not possible. In this case, it is natural to try to hedge on average in the sense of the following definition: Definition 2.1 (Conditional-Mean Hedge). Let f (ST ) be a European vanilla type optionwith convex or concave payoff function f . Let π be its Markovian replicating strategy: πt = f ′(St ). We call the discrete-time strategy πN a conditional-mean hedge, if for all trading times ti, E [ Vπ N ,k ti+1 |Fti ] = E [ Vπti+1 |Fti ] . (2.5) Here Fti is the information generated by the asset price process S up to time ti. 38 Acta Wasaensia F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 87 Remark 2.1 (Conditional-Mean Hedge as Tracking Condition). Criterion (2.5) is actually a tracking requirement. We do not only require that the conditional means agree on the last trading time before the maturity, but also on all trading times. In this sense the criterion has an ‘‘American’’ flavor in it. From a purely ‘‘European’’ hedging point of view, one can simply remove all but the first and the last trading times. Remark 2.2 (Arbitrage and Uniqueness of Conditional-Mean Hedge). The conditional-mean hedging strategy πN depends on the continuous-time hedging strategy π . Since there is strong arbitrage in the fractional Black–Scholes model (zero can be perfectly replicated with negative initial wealth), the replicating strategy π is not unique. However, the strong arbitrage strategies are very complicated. Indeed, it follows directly from the change-of-variables formula that in the class of Markovian strategies πt = g(t, St ), the choice πt = f ′(St ) is the unique replicating strategy for the claim f (ST ). We stress that the expectation in (2.5) is with respect to the true probability measure; not under any equivalent martingale measure. Indeed, equivalent martingale measures do not exist in the fractional Black–Scholes model. To find the solution to (2.5) one must be able to calculate the conditional expectations involved. This can be done by using Sottinen and Viitasaari (2017, Theorem 3.1), a version of which we state below as Lemma 2.1 for the readers’ convenience. Lemma 2.1 (Conditional Fractional Brownian Motion). The fractional Brownian motion B with index H ∈ ( 12 , 1) conditioned on its own past FBu is the Gaussian process B(u) = B|FBu with FBu -measurable mean Bˆt (u) = Bu − ∫ u 0 Ψ (t, s|u) dBs, where Ψ (t, s|u) = − sin(π (H − 1 2 )) π s 1 2−H (u − s) 12−H ∫ t u zH− 1 2 (z − u)H− 12 z − s dz, and deterministic covariance function rˆ(t, s|u) = r(t, s) − ∫ u 0 k(t, v)k(s, v)dv, where k(t, s) = ( H − 1 2 )√ 2HΓ ( 3 2 − H ) Γ ( H − 12 ) Γ (2 − 2H) s 1 2−H ∫ t s zH− 1 2 (z − s)H− 23 dz; Γ is the Euler’s gamma function. Remark 2.3 (Conditional Asset Process). By (2.2) we have the equality of filtrations: FBt = FSt = Ft , for t ∈ [0, T ]. Consequently, the conditional process S(u) = S|Fu is, informally, given by St (u) = S0eμt+σBt (u) = Sueμ(t−u)+σ(Bt (u)−Bu). More formally, this means, in particular, that for t > u, E [ f (St ) ∣∣FSu ] = E [f (S0eμt+σBt ) ∣∣FBu ] = ∫ ∞ −∞ f ( S0eμt+σ Bˆt (u)+σ √ rˆ(t|u) z ) φ(z)dz = ∫ ∞ −∞ f ( Sue μ(t−u)+σ ( Bˆt (u)−Bu ) +σ √ rˆ(t|u) z ) φ(z)dz, where we have denoted rˆ(t|u) = rˆ(t, t|u), and φ is the standard normal density function. 3. Conditional-Mean hedging strategies Denote ΔBˆti+1 (ti) = Bˆti+1 (ti) − Bti . Acta Wasaensia 39 88 F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 In Theorem 3.1 we will calculate the conditional-mean hedging strategy in terms of the following conditional gains: ΔSˆti+1 (ti) = Sˆti+1 (ti) − Sti = E [Sti+1 |Fti]− Sti , ΔVˆπti+1 (ti) = Vˆπti+1 (ti) − Vπti = E [ Vπti+1 |Fti ] − Vπti , ΔVˆπ N ,k ti+1 (ti) = Vˆπ N ,k ti+1 (ti) − Vπti = E [ Vπ N ,k ti+1 |Fti ] − VπN ,kti . Lemma 3.1 states that all these conditional gains can be calculated explicitly by using the prediction law of the fractional Brownian motion. Lemma 3.1 (Conditional Gains). ΔSˆti+1 (ti) = Sti (∫ ∞ −∞ eμΔti+1+σΔBˆti+1 (ti)+σ √ rˆ(ti+1|ti)zφ(z) dz − 1 ) , ΔVˆπti+1 (ti) = ∫ ∞ −∞ f ( Stie μΔti+1+σΔBˆti+1 (ti)+σ √ rˆ(ti+1|ti)z ) φ(z) dz − f (Sti ), ΔVˆπ N ,k ti+1 (ti) = πNti ΔSˆti+1 (ti) − kSti |ΔπNti |. Proof. Let g : R → R be such that E [∣∣g(Bti+1 )∣∣] < ∞. Then, by Lemma 2.1, E [ g(Bti+1 ) |Fti ] = ∫ ∞ −∞ g ( Bˆti+1 (ti) + √ rˆ(ti+1|ti)z ) φ(z) dz. Consider ΔSˆti+1 (ti). By choosing g(x) = S0eμt+σx, we obtain Sˆti+1 (ti) = E [ Sti+1 ∣∣Fti] = E [g (Bti+1) ∣∣Fti] = ∫ ∞ −∞ S0e μti+1+σ ( Bˆti+1 (ti)+ √ rˆ(ti+1|ti)z ) φ(z) dz. The formula for ΔSˆti+1 (ti) follows from this. Consider then ΔVπti+1 (ti). By choosing g(x) = f (S0eμt+σx) we obtain Vˆπti+1 (ti) = E [ Vπti+1 |Fti ] = E [f (Sti+1 ) |Fti] = E [g(Bti+1 ) |Fti] = ∫ ∞ −∞ g ( Bˆti+1 (ti) + √ rˆ(ti+1|ti)z ) φ(z) dz = ∫ ∞ −∞ f ( S0e μti+1+σ ( Bˆti+1 (ti)+ √ rˆ(ti+1|ti)z )) φ(z) dz. The formula for ΔVˆπti+1 (ti) follows from this. 40 Acta Wasaensia F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 89 Finally, we calculate Vˆπ N ,k ti+1 (ti) = E [ Vπ N ,k ti+1 ∣∣Fti] = VπN ,kti + E [∫ ti+1 ti πNu dSu − ∫ ti+1 ti kSu|dπNu | ∣∣∣Fti ] = VπN ,kti + πNti ( E [ Sti+1 ∣∣Fti]− Sti)− kSti |ΔπNti | = VπN ,kti + πNti ΔSˆti+1 (ti) − kSti |ΔπNti |. The formula for ΔVˆπ N ,k ti+1 (ti) follows from this.  Now we are ready to state and prove our main result. We note that, in principle, our result is general: it is true in any pricingmodelwhere the option f (ST ) can be replicated. In practice, our result is specific to the fractional Black–Scholesmodel via Lemma 3.1. Theorem 3.1 (Conditional-Mean Hedging Strategy). The conditional mean hedge of the European vanilla type option with convex or concave positive payoff function f with proportional transaction costs k is given by the recursive equation πNti = ΔVˆπti+1 (ti) + (Vπti − Vπ N ,k ti ) + kSti |ΔπNti | ΔSˆti+1 (ti) , (3.1) where Vπ N ,k ti is determined by (2.4). Proof. Let us first consider the left hand side of (2.5). We have E [ Vπ N ,k ti+1 ∣∣Fti] = E [ Vπ N ,k ti + ∫ ti+1 ti πNu dSu − k ∫ ti+1 ti Su|dπNu | ∣∣∣Fti ] = VπN ,kti + πNti E [ Sti+1 (ti) − Sti ∣∣Fti]− kSti |ΔπNti| = VπN ,kti + πNti ΔSˆti+1 (ti) − kSti |ΔπNti |. For the right-hand-side of (2.5), we simply write E [ Vπti+1 ∣∣Fti] = ΔVˆπti+1 (ti) + Vπti . Equating the sides we obtain (3.1) after a little bit of simple algebra.  Remark 3.1. Taking the expected gains ΔSˆti+1 (ti) to be the numéraire, one recognizes three parts in the hedging formula (3.1). First, one invests on the expected gains in the time-value of the option. This ‘‘conditional-mean Delta-hedging’’ is intuitively themost obvious part. Indeed, a naïve approach to conditional-mean hedgingwould only give this part, forgetting to correct for the tracking-errors already made, which is the second part in (3.1). The third part in (3.1) is obviously due to the transaction costs. Remark 3.2. Eq. (3.1) for the strategy of the conditional-mean hedging is recursive: in addition to the filtration Fti , the position πNti−1 is needed to determine the position π N ti . Consequently, to determine the conditional-mean hedging strategy by using (3.1), the initial position πN0 must be fixed. The initial position is, however, not uniquely defined. Indeed, let β N 0 be the position in the riskless asset. Then the conditional-mean criterion (2.5) only requires that βN0 + πN0 E[St1 ] − kS0|πN0 | = E[f (St1 )]. There are of course infinite number of pairs (βN0 , π N 0 ) solving this equation. A natural way to fix the initial position (β N 0 , π N 0 ) for the investor interested in conditional-mean hedging would be the one with minimal cost. If short-selling is allowed, the investor is then faced with the minimization problem min πN0 ∈R v(πN0 ), where the initial wealth v is the piecewise linear function v(πN0 ) = βN0 + πN0 S0 = ⎧⎨ ⎩ E[f (St1 )] − ( ΔSˆt1 (0) − kS0 ) πN0 , if π N 0 ≥ 0, E[f (St1 )] − ( ΔSˆt1 (0) + kS0 ) πN0 , if π N 0 < 0. Acta Wasaensia 41 90 F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 Clearly, the minimal solution πN0 is independent of E[f (St1 )], and, consequently, of the option to be replicated. Also, the minimization problem is bounded if and only if k ≥ ∣∣∣∣∣ΔSˆt1 (0)S0 ∣∣∣∣∣ , i.e. the proportional transaction costs are bigger than the expected return on [0, t1] of the stock. In this case, the minimal cost conditional mean-hedging strategy starts by putting all the wealth in the riskless asset. We end this note by applying Theorem 3.1 to European call options. Corollary 3.1 (European Call Option). Denote dˆ+ti+1 (ti) = ln Sti K − μΔti+1 − σΔBˆti+1 (ti) σ √ rˆ(ti+1|ti) − σ √ rˆ(ti+1|ti), dˆ−ti+1 (ti) = ln Sti K − μΔti+1 − σΔBˆti+1 (ti) σ √ rˆ(ti+1|ti) , Xˆti+1 (ti) = μΔti+1 + σΔBˆti+1 (ti) + 1 2 σ 2 rˆ(ti+1|ti), and let Φ be the cumulative distribution function of the standard normal law. Then the conditional-mean hedging strategy for the European call option with strike-price K is given by πNti = Stie Xˆti+1 (ti)Φ(dˆ+ti+1 (ti)) − KΦ(dˆ−ti+1 (ti)) − Vπ N ,k ti + kSti |ΔπNti | ΔSˆti+1 (ti) . (3.2) Proof. First we note that Vˆ callti+1 (ti) = ∫ ∞ −∞ ( Stie μΔti+1+σΔBˆti+1 (ti)+σ √ rˆ(ti+1|ti)z − K )+ φ(z)dz = StieXˆti+1 (ti)Φ ( dˆ+ti+1 (ti) ) − KΦ ( dˆ−ti+1 (ti) ) . Next we note that V callti = (Sti − K )+. So, ΔVˆ callti+1 (ti) = StieXˆti+1 (ti)Φ(dˆ+ti+1 (ti)) − KΦ(dˆ−ti+1 (ti)) − (Sti − K )+, and (3.2) follows from this.  Acknowledgments F. Shokrollahi was funded by the graduate school of the University of Vaasa.We thank the referee for valuable comments. References Azmoodeh, E., 2013. On the fractional Black–Scholes market with transaction costs. Commun. Math. Finance 2 (3), 21–40. Azmoodeh, E., Mishura, Y., Valkeila, E., 2009. On hedging European options in geometric fractional Brownianmotionmarketmodel. Statist. Decisions 27 (2), 129–143. Bender, C., Sottinen, T., Valkeila, E., 2011. Fractional processes as models in stochastic finance. In: Advanced Mathematical Methods for Finance. Springer, Heidelberg, pp. 75–103. Denis, E., Kabanov, Y., 2010. Mean square error for the Leland-Lott hedging strategy: convex pay-offs. Finance Stoch. 14 (4), 625–667. Gapeev, P.V., Sottinen, T., Valkeila, E., 2011. Robust replication inH-self-similar Gaussianmarketmodels under uncertainty. Statist. Decisions 28 (1), 37–50. Kabanov, Y., Safarian, M., 2009. Markets with Transaction Costs. In: Springer Finance, Springer-Verlag, Berlin, p. xiv+294. Mathematical theory. Leland, H.E., 1985. Option pricing and replication with transactions costs. J. Finance 40 (5), 1283–1301. Shokrollahi, F., Kılıçman, A., 2014. Pricing currency option in amixed fractional Brownianmotionwith jumps environment. Math. Probl. Eng. Art. ID 858210, 13. Shokrollahi, F., Kılıçman, A., 2015. Actuarial approach in a mixed fractional Brownian motion with jumps environment for pricing currency option. Adv. Differential Equations 2015:257, 8. Shokrollahi, F., Kılıçman, A., Ibrahim, N.A., Ismail, F., 2015. Greeks and partial differential equations for some pricing currency options models. Malays. J. Math. Sci. 9 (3), 417–442. Shokrollahi, F., Kılıçman, A., Magdziarz, M., 2016. Pricing European options and currency options by time changed mixed fractional Brownian motion with transaction costs. Int. J. Financ. Eng. 3 (1) 1650003, 22. Sottinen, T., Viitasaari, L., 2017. Prediction law of fractional Brownian motion. Statist. Probab. Lett. 129, 155–166. 42 Acta Wasaensia F. Shokrollahi, T. Sottinen / Statistics and Probability Letters 130 (2017) 85–91 91 Wang, X.-T., 2010a. Scaling and long-range dependence in option pricing I: pricing European optionwith transaction costs under the fractional Black-Scholes model. Physica A 389 (3), 438–444. Wang, X.-T., 2010b. Scaling and long range dependence in option pricing, IV: pricing European options with transaction costs under the multifractional Black-Scholes model. Physica A 389 (4), 789–796. Wang, X.-T., Yan, H.-G., Tang, M.-M., Zhu, E.-H., 2010a. Scaling and long-range dependence in option pricing III: a fractional version of the Merton model with transaction costs. Physica A 389 (3), 452–458. Wang, X.-T., Zhu, E.-H., Tang, M.-M., Yan, H.-G., 2010b. Scaling and long-range dependence in option pricing II: pricing European option with transaction costs under the mixed Brownian-fractional Brownian model. Physica A 389 (3), 445–451. Acta Wasaensia 43 APPLIED & INTERDISCIPLINARY MATHEMATICS | RESEARCH ARTICLE Subdiffusive fractional Black–Scholes model for pricing currency options under transaction costs Foad Shokrollahi1* Abstract: A new framework for pricing European currency option is developed in the case where the spot exchange rate follows a subdiffusive fractional Black– Scholes. An analytic formula for pricing European currency call option is proposed by a mean self-financing delta-hedging argument in a discrete time setting. The minimal price of a currency option under transaction costs is obtained as time-step Δt ¼ tα1ΓðαÞ  1 2 π   1 2H k σ  1 H, which can be used as the actual price of an option. In addition, we also show that time-step and long-range dependence have a significant impact on option pricing. Subjects: Science; Mathematics & Statistics; Applied Mathematics; Statistics & Probability Keywords: subdiffusion process; currency option; transaction costs; inverse subordinator process MR Subject classifications: 91G20; 91G80; 60G22 1. Introduction The standard European currency option valuation model has been presented by Garman and Kohlhagen ðG KÞ (Garman & Kohlhagen, 1983). However, some papers have provided evidence of the mispricing for currency options by the G K model. The most important reason why this ABOUT THE AUTHOR Foad Shokrollahi is a researcher at Department of Mathematics and Statistics, University of Vaasa, Finland. His research interests are Stochastic Processes, Stochastic Analysis, Fractional Brownian motion, and Mathematical Finance. PUBLIC INTEREST STATEMENT Subdiffusion refers to a well-known and estab- lished phenomenon in statistical physics. One description of subdiffusion is related to subordi- nation, where the standard diffusion process is time-changed by the so-called inverse subordi- nator. According to the features of the subdiffu- sion process and the fractional Brownian motion, we propose the new model for pricing European currency options by using the fractional Brownian motion, subdiffusive strategy, and scaling time in discrete time setting, to get behavior from finan- cial markets. Motivated by this objective, we illustrate how to price a currency options in dis- crete time setting for both cases: with and with- out transaction costs by applying subdiffusive fractional Brownian motion model. By considering the empirical data, we will demonstrate that the proposed model is further flexible in comparison with the previous models and it obtains suitable benchmark for pricing currency options. Additionally, impact of the parameters on our pricing formula is investigated. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 © 2018 The Author(s). This open access article is distributed under a Creative Commons Attribution (CC-BY) 4.0 license. Received: 16 November 2017 Accepted: 24 April 2018 First Published: 29 May 2018 *Corresponding author: Foad Shokrollahi, Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700, FIN-65101 Vaasa, Finland E-mail: foad.shokrollahi@uva.fi Reviewing editor: Carlo Cattani, University of Tuscia, Italy Additional information is available at the end of the article Page 1 of 14 44 Acta Wasaensia model may not be entirely satisfactory could be that currencies are different from stocks in important respects and the geometric Brownian motion cannot capture the behavior of currency return (Ekvall, Jennergren, & Näslund, 1997). Since then, many methodologies for currency option pricing have been proposed by using modifications of the G K model (Garman & Kohlhagen, 1983; Ho, Stapleton, & Subrahmanyam, 1995). All this research above assumes that the logarithmic returns of the exchange rate are indepen- dent identically distributed normal random variables. However, in general, the assumptions of the Gaussianity and mutual independence of underlying asset log returns would not hold. Moreover, the empirical research has also shown that the distributions of the logarithmic returns in the financial market usually exhibit excess kurtosis with more probability mass near the origin and in the tails and less in the flanks than would occur for normally distributed data (Dai & Singleton, 2000). That is to say the features of financial return series are non-normality, non-independence, and nonlinearity. To capture these non-normal behaviors, many researchers have considered other distributions with fat tails such as the Pareto-stable distribution and the Generalized Hyperbolic Distribution. Moreover, self-similarity and long-range dependence have become important con- cepts in analyzing the financial time series. There is strong evidence that the stock return has little or no autocorrelation. As fractional Brownian motion ðFBMÞ has two important properties called self-similarity and long-range depen- dence, it has the ability to capture the typical tail behavior of stock prices or indexes (Borovkov, Mishura, Novikov, & Zhitlukhin, 2018; Shokrollahi & Sottinen, 2017). The fractional Black–Scholes ðFBSÞ model is an extension of the Black–Scholes ðBSÞ model, which displays the long-range dependence observed in empirical data. This model is based on replacing the classic Brownian motion by the fractional Brownian motion ðFBMÞ in the Black–Scholes model. That is V^ðtÞ ¼ V^0 exp μtþ σ B^HðtÞ n o ; V^0 > 0; (1:1) where μ; and σ are fixed, and B^HðtÞ is a FBM with Hurst parameter H 2 ½12 ;1Þ. It has been shown that the FBS model admits arbitrage in a complete and frictionless market (Cheridito, 2003; Shokrollahi & Kılıçman, 2014; Sottinen & Valkeila, 2003; Wang, Zhu, Tang, & Yan, 2010; Xiao, Zhang, Zhang, & Wang, 2010). Wang (2010) resolved this contradiction by giving up the arbitrage argument and examining option replication in the presence of proportional transaction costs in discrete time setting (Mastinšek, 2006). Magdziarz (2009a) applied the subdiffusive mechanism of trapping events to describe properly financial data exhibiting periods of constant values and introduced the subdiffusive geometric Brownian motion VαðtÞ ¼ VðTαðtÞÞ; (1:2) as the model of asset prices exhibiting subdiffusive dynamics, where VαðtÞ is a subordinated process (for the notion of subordinated processes please refer to Refs. Janicki and Weron (1993, 1995), Kumar, Wyłomańska, Połoczański, and Sundar (2017), Piryatinska, Saichev, and Woyczynski (2005), in which the parent process VðτÞ is a geometric Brownian motion and TαðtÞ is the inverse α-stable subordinator defined as follows: TαðtÞ ¼ inffτ>0 : QαðτÞ> tg; 0< α<1: (1:3) Here, QαðtÞ is a strictly increasing α-stable subordinator with Laplace transform: E eηQαðτÞ   ¼ eτηα , 0< α<1, where E denotes the mathematical expectation. Magdziarz (2009a) demonstrated that the considered model is free-arbitrage but is incomplete and proposed the corresponding subdiffusive BS formula for the fair prices of European options. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 2 of 14 Acta Wasaensia 45 Subdiffusion is a well-known and established phenomenon in statistical physics. The usual model of subdiffusion in physics is developed in terms of FFPE (fractional Fokker-Planck equations). This equation was first derived from the continuous-time random walk scheme with heavy-tailed waiting times (Metzler & Klafter, 2000). It provides a useful way for the description of transport dynamics in complex systems (Magdziarz, Weron, & Weron, 2007). Another description of sub- diffusion is in terms of subordination, where the standard diffusion process is time-changed by the so-called inverse subordinator (Gu, Liang, & Zhang, 2012; Guo, 2017; Janczura, Orzeł, & Wyłomańska, 2011; Magdziarz, 2009b, Magdziarz et al., 2007; Scalas, Gorenflo, & Mainardi, 2000, Shokrollahi & Kılıçman, 2014; Yang, 2017). The objective of this paper is to study the European call currency option by a mean self financing delta hedging argument. The main contribution of this paper is to derive an analytical formula for European call currency option without using the arbitrage argument in discrete time setting when the exchange rate follows a subdiffusive FBS St ¼ V^ðTαðtÞÞ ¼ S0exp μTαðtÞ þ σB^HðTαðtÞÞ n o ; (1:4) S0 ¼ V^ð0Þ>0: We then apply the result to value European put currency option. We also provide representative numerical results. Making the change of variable, BHðtÞ ¼ μþrfrdσ tþ B^HðtÞ, under the risk-neutral measure, we have that St ¼ V^ðRβðtÞÞ ¼ S0exp ðrd  rf ÞðTαðtÞÞ þ σBHððTαðtÞÞ   ; (1:5) S0 ¼ V^ð0Þ>0: This formula is similar to the Black–Scholes option pricing formula, but with the volatility being different. We denote the subordinated process Wα;HðtÞ ¼ BHðTαðtÞÞ, here the parent process BHðτÞ is a FBM and TαðtÞ is assumed to be independent of BHðτÞ. The process Wα;HðtÞ is called a subdiffusion process. Particularly, when H ¼ 12 , it is a subdiffusion process presented in Karipova and Magdziarz (2017), Kumar et al. (2017), and Magdziarz (2010). Figure 1 shows typically the differences and relationships between the sample paths of the spot exchange rate in the FBS model and the subdiffusive FBS model. The rest of the paper proceeds as follows: In Section 2, we provide an analytic pricing formula for the European currency option in the subdiffusive FBS environment and some Greeks of our pricing model are also obtained. Section 3 is devoted to analyze the impact of scaling and long-range dependence on currency option pricing. Moreover, the comparison of our subdiffusive FBS model and traditional models is undertaken in this section. Finally, Section 4 draws the concluding remarks. The proof of Theorems are provided in Appendix. 2. Pricing model for the European call currency option In this section, we derive a pricing formula for the European call currency option of the subdiffusive FBS model under the following assumptions: (i) We consider two possible investments: (1) a stock whose price satisfies the equation: St ¼ S0exp ðrd  rf ÞTαðtÞ þ σWα;HðtÞ   ; S0 >0; (2:1) Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 3 of 14 46 Acta Wasaensia where α 2 12 ;1   , H 2 ½12 ;1Þ, αþ αH>1, and rd; and rf are the domestic and the foreign interest rates, respectively. (2) A money market account: dFt ¼ rdFtdt; (2:2) where rd shows the domestic interest rate. (ii) The stock pays no dividends or other distributions, and all securities are perfectly divisible. There are no penalties to short selling. It is possible to borrow any fraction of the price of a security to buy it or to hold it, at the short-term interest rate. These are the same valuation policy as in the BS model. (iii) There are transaction costs that are proportional to the value of the transaction in the underlying stock. Let k denote the round trip transaction cost per unit dollar of transaction. Suppose U shares of the underlying stock are bought ðU>0Þ or sold ðU<0Þ at the price St, then the transaction cost is given by k2 Uj jSt in either buying or selling. Moreover, trading takes place only at discrete intervals. (iv) The option value is replicated by a replicating portfolio  with UðtÞ units of stock and riskless bonds with value FðtÞ. The value of the option must equal the value of the replicating portfolio to reduce (but not to avoid) arbitrage opportunities and be consistent with economic equilibrium. (v) The expected return for a hedged portfolio is equal to that from an option. The portfolio is revised every Δt and hedging takes place at equidistant time points with rebalancing intervals of (equal) length Δt, where Δt is a finite and fixed, small time-step. Remark 2.1. From Guo and Yuan (2014), Magdziarz (2009c), we have EðTmα ðtÞÞ ¼ t mαm! Γðmαþ1Þ . Then, by using α-self-similar and non-decreasing sample paths of TαðtÞ, we can obtain that α-self-similar and non-decreasing sample paths of TαðtÞ, E ΔTαðtÞð Þ ¼ E Tαðtþ ΔtÞ  TαðtÞð Þ ¼ 1Γð1þαÞ ðtþ ΔtÞα  tα½  ¼ t α1 ΓðαÞΔt: (2:3) and E ðΔBHðTαðtÞÞ2   ¼ t α1 ΓðαÞ  2H Δt2H: (2:4) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.99 0.995 1 1.005 1.01 1.015 1.02 1.025 1.03 1.035 t V t 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.985 0.99 0.995 1 1.005 1.01 1.015 1.02 1.025 1.03 1.035 t S t Figure 1. Comparison of the spot exchange rate’ sample paths in the FBS model (left) and the subdiffusive FBS model (right) for rd ¼ 0:03; rf ¼ 0:02; α¼ 0:9; H¼ 0:8; σ¼ 0:1; S0 ¼ 1. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 4 of 14 Acta Wasaensia 47 Let C ¼ Cðt; StÞ be the price of a European currency option at time t with a strike price K that matures at time T. Then, the pricing formula for currency call option is given by the following theorem. Theorem 2.1. C ¼ Cðt; StÞ is the value of the European currency call option on the stock St satisfied (1.5) and the trading takes place discretely with rebalancing intervals of length Δt. Then, C satisfies the partial differential equation @C @t þ ðrd  rf ÞSt @C @St þ 1 2 σ^2S2t @2C @S2t  rdC ¼ 0; (2:5) with boundary condition CðT; STÞ ¼ max ST  K;0f g. The value of the currency call option is Cðt; StÞ ¼ Sterf ðTtÞΦðd1Þ  KerdðTtÞΦðd2Þ; (2:6) and the value of the put currency option is Pðt; StÞ ¼ KerdðTtÞΦðd2Þ  Sterf ðTtÞΦðd1Þ; (2:7) where d1 ¼ ln StK  þ rd  rf ðT  tÞ þ σ^22 ðT tÞ σ^ ffiffiffiffiffiffiffiffiffiffi T tp ; d2 ¼ d1  σ^ðtÞ ffiffiffiffiffiffiffiffiffiffi T t p ; (2:8) σ^2 ¼ σ2 t α1 ΓðαÞ 2H Δt2H1 þ ffiffiffi 2 π r k σ tα1 ΓðαÞ H ΔtH1 " # ; (2:9) where Φð:Þ is the cumulative normal distribution function. In what follows, the properties of the subdiffusive FBS model are discussed, such as Greeks, which summarize how option prices change with respect to underlying variables and are critically important to asset pricing and risk management. The model can be used to rebalance a portfolio to achieve the desired exposure to certain risk. More importantly, by knowing the Greeks, particular exposure can be hedged from adverse changes in the market by using appropriate amounts of other related financial instruments. In contrast to option prices that can be observed in the market, Greeks cannot be observed and must be calculated given a model assumption. The Greeks are typically computed using a partial differentiation of the price formula. Theorem 2.2. The Greeks can be written as follows: Δ ¼ @C @St ¼ erf ðTtÞΦðd1Þ; (2:10)  ¼ @C @K ¼ erdðTtÞΦðd2Þ; (2:11) ρrd ¼ @C @rd ¼ KðT  tÞerdðTtÞΦðd2Þ; (2:12) ρrf ¼ @C @rf ¼ StðT  tÞerf ðTtÞΦðd1Þ; (2:13) Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 5 of 14 48 Acta Wasaensia Θ ¼ @C @t ¼ Strf erf ðTtÞΦðd1Þ  KrderdðTtÞΦðd2Þ þ Sterf ðTtÞσ2ðα 1Þ t α2 ΓðαÞ H t α1 ΓðαÞ  2H1 Δt2H1 bσ ffiffiffiffiffiffiffiffiffiffiT tp ðT tÞΦ0ðd1Þ þ Sterf ðTtÞ ffiffiffi 2 π r kσðβ 1Þ t α2 ΓðαÞ H t α1 ΓðαÞ  H1 ΔtH1 2bσ ffiffiffiffiffiffiffiffiffiffiT tp ðT  tÞΦ0ðd1Þ  Sterf ðTtÞ bσ 2 ffiffiffiffiffiffiffiffiffiffi T tp Φ 0ðd1Þ; (2:14) Γ ¼ @ 2C @S2t ¼ erf ðTtÞ Φ 0ðd1Þ Stσ^ ffiffiffiffiffiffiffiffiffiffi T tp ; (2:15) #σ^ ¼ @C @σ^ ¼ Sterf ðTtÞ ffiffiffiffiffiffiffiffiffiffi T  t p Φ0ðd1Þ: (2:16) Remark 2.2. The modified volatility without transaction costs ðk ¼ 0Þ is given by σ^2 ¼ σ2 t α1 ΓðαÞ 2H Δt2H1 " # ; (2:17) specially if α " 1, σ^2 ¼ σ2Δt2H1; (2:18) which is consistent with the result in Necula (2002). Furthermore, from Equation (2.18), if H # 12 , then σ^2 ¼ σ2, which is according to the results with the G K model (Garman & Kohlhagen, 1983). Letting α " 1, from Equation (2.9), we obtain Remark 2.3. The modified volatility under transaction costs is given by σ^2 ¼ σ2 Δt2H1 þ ffiffiffi 2 π r k σ ΔtH1 " # ; (2:19) that is in line with the findings in Wang (2010). 3. Empirical studies The objective of this section is to obtain the minimal price of an option with transaction costs and to show the impact of time scaling Δt, transaction costs k, and subordinator parameter α on the subdiffusive FBSmodel. Moreover, in the last part, we compute the currency option prices using our model and make comparisons with the results of the G K and FBS models. As kσ< ffiffi π 2 p often holds (for example: σ ¼ 0:1; k ¼ 0:01), from Equation (2.9), we have σ^2 σ2 ¼ tα1ΓðαÞ  2H Δt2H1 þ ffiffi 2 π q k σ tα1 ΓðαÞ  H ΔtH1  2 tα1ΓðαÞ  3 2H Δt 3 2H1 2 π  1 4 k σ  1 2; (3:1) Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 6 of 14 Acta Wasaensia 49 where H> 12 . Then, the minimal volatility σ^min is ffiffiffi 2 p σ t α1 ΓðαÞ  1 2 2 π  1 2 14H k σ  1 12H as Δt ¼ tα1ΓðαÞ 1 2π  12H kσ 1H. Thus, the minimal price of an option under transaction costs is represented as Cminðt; StÞ with σ^min in Equation (2.8). Moreover, the option rehedging time interval for traders to take is Δt ¼ tα1ΓðαÞ  1 2 π   1 2H k σ  1 H. The minimal price Cminðt; StÞ can be used as the actual price of an option. In particular, as Δt<1;α 2 12 ;1   and @C@σ^ ¼ Sterf ðTtÞ ffiffiffiffiffiffi Tt p ffiffiffiffi 2π p e d2 2 >0; @bσ @H ¼ σ 2 t α1 ΓðαÞ 2H Δt2H1 þ ffiffiffi 2 π r k σ tα1 ΓðαÞ H ΔtH1 " # ln tα1 ΓðαÞ þ lnΔt   2 t α1 ΓðαÞ 2H Δt2H1 þ ffiffiffi 2 π r k σ tα1 ΓðαÞ H ΔtH1 " #12 ¼ 2 t α1 ΓðαÞ 2H Δt2H1 þ ffiffiffi 2 π r k σ tα1 ΓðαÞ H ΔtH1 " #  σ2 ln t α1 ΓðαÞ   þ lnΔt h i 2bσ <0; (3:2) and @C@H ¼ @C@σ^ @σ^@H , then we have @C @H <0 asH 2 ½1 2 ;1Þ; (3:3) which displays that an increasing Hurst exponent comes along with a decrease of the option value (see Figure 2). On the other hand, if H # 12 , then σ^min ¼ ffiffiffi 2 p σ tα1 ΓðαÞ 1 2 2 π 1 2 14H k σ 1 12H ! σ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 tα1 ΓðαÞ s (3:4) and if α " 1, then σ^min ! ffiffiffi 2 p σ as H # 12 : In addition, if H # 12 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.05 0.1 0.15 0.2 0.25 k O pt io n V al ue 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1 3 4 5 6 7 8 9 10 x 10 −3 α O pt io n V al ue 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.005 0.01 0.015 0.02 0.025 0.03 Δ t O pt io n V al ue 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1 0 0.005 0.01 0.015 0.02 0.025 H O pt io n V al ue Figure 2. Call currency option values. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 7 of 14 50 Acta Wasaensia Δt ¼ t α1 ΓðαÞ 1 2 π 1 2H k σ 1 H ! t α1 ΓðαÞ 1 2 π k σ 2 ; (3:5) and if α " 1, then Δt! 2π   k σ  2 as H # 12 : Lux and Marchesi (1999) have shown that Hurst exponent H ¼ 0:51 0:004 in some cases, so Equations (3.4) and (3.5) have a practical application in option pricing. For example: if H # 12 ; α " 1; k ¼ 2% and σ ¼ 20%, then σ^min ! ffiffi 2 p 20 , and Δt! 0:02π ; and if H " 12 ; α " 1; k ¼ 0:2% and σ ¼ 20%, then σ^min ! ffiffi 2 p 20 , and Δt ! 2π  104: In the following, we investigate the impact of scaling and long-range dependence on option pricing. It is well known that Mantegna and Stanley (1995) introduced the method of scaling invariance from the complex science into the economic systems for the first time. Since then, a lot of research for scaling laws in finance has begun. If H ¼ 12 and k ¼ 0, from Equation (2.9), we know that σ^2 ¼ σ2 tα1ΓðαÞ   shows that fractal scaling Δt has not any impact on option pricing if a mean self- financing delta-hedging strategy is applied in a discrete time setting, while subordinator para- meter β has remarkable impact on option pricing in this case. In particular, from Equations (3.4) and (3.5), we know that σ^min ! σ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 tα1ΓðαÞ  r as H  12 and Δt ! t α1 ΓðαÞ  1 2 π   k σ  2 ; as H  12 . Therefore, Cminðt; StÞ is approximately scaling free with respect to the parameter k, if H  12 , but is scaling dependent with respect to subordinator parameter α. However, Δt ! tα1ΓðαÞ  1 2 π   k σ  2 ; is scaling dependent with respect to parameters k and α, if H  12 . On the other hand, if H> 12 and k ¼ 0, from Equation (2.17), we know that σ^2 ¼ σ2 tα1ΓðαÞ  2H Δt2H1  , which displays that the fractal scaling Δt and sabordinator parameter α have a significant impact on option pricing. Furthermore, for k0, from Equation (2.8), we know that option pricing is scaling dependent in general. Now, we present the values of currency call option using subdiffusive FBS model for different parameters. For the sake of simplicity, we will just consider the out-of-the-money case. Indeed, using the same method, one can also discuss the remaining cases: in-the-money and at-the-money. First, the prices of our subdiffusive FBS model are investigated for some Δt and prices for different exponent parameters. The prices of the call currency option versus its parameters H;Δt;α and k are revealed in Figure 2. The selected parameters are St ¼ 1:4; K ¼ 1:5; σ ¼ 0:1; rd ¼ 0:03; rf ¼ 0:02; T ¼ 1; t ¼ 0:1; Δt ¼ 0:01; k ¼ 0:01; H ¼ 0:8; α ¼ 0:9: Figure 2 indicates that the option price is an increasing function of k and Δt, while it is a decreasing function of H and α. For a detailed analysis of our model, the prices calculated by the G K, FBS and subdiffusive FBS models are compared for both out-of-the-money and in-the-money cases. The following para- meters are chosen: St ¼ 1:2; σ ¼ 0:5; rd ¼ 0:05; rf ¼ 0:01; t ¼ 0:1;Δt ¼ 0:01; k ¼ 0:001, and H ¼ 0:8, along with time maturity T 2 ½0:1;2, strike price K 2 ½0:8;1:19 for the in-the-money case and K 2 ½1:21;1:4 for the out-of-the-money case. Figures 3 and 4 show the theoretical values difference by the G K, FBS, and our subdiffusive FBS models for the in-the-money and out-of-the-money, respectively. As indicated in these figures, the values computed by our subdiffusive FBS model are better fitted to the G K values than the FBS model for both in-the-money and out-of-the money cases. Hence, when compared to these figures, our subdiffusive FBS model seems reasonable. 4. Conclusion Without using the arbitrage argument, in this paper, we derive a European currency option pricing model with transaction costs to capture the behavior of the spot exchange rate price, where the Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 8 of 14 Acta Wasaensia 51 spot exchange rate follows a subdiffusive FBS with transaction costs. In discrete time case, we show that the time scaling Δt and the Hurst exponent H play an important role in option pricing with or without transaction costs and option pricing is scaling dependent. In particular, the minimal price of an option under transaction costs is obtained. Acknowledgments I would like to thank the referees and the editor for their careful reading and their valuable comments. Funding The author received no direct funding for this research. This paper is supported by the university of Vaasa, Finland. Author details Foad Shokrollahi1 E-mail: foad.shokrollahi@uva.fi 1 Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700, FIN-65101 Vaasa, Finland. Citation information Cite this article as: Subdiffusive fractional Black–Scholes model for pricing currency options under transaction costs, Foad Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145. References Borovkov, K., Mishura, Y., Novikov, A., & Zhitlukhin, M. (2018). New and refined bounds for expected max- ima of fractional Brownian motion. Statistics & Probability Letters, 137, 142–147. doi:10.1016/j. spl.2018.01.025 Cheridito, P. (2003). Arbitrage in fractional Brownian motion models. Finance and Stochastics, 7(4), 533–553. doi:10.1007/s007800300101 Dai, Q., & Singleton, K. J. (2000). Specification analysis of affine term structure models. The Journal of Finance, 55(5), 1943–1978. doi:10.1111/0022- 1082.00278 Ekvall, N., Jennergren, L. P., & Näslund, B. (1997). Currency option pricing with mean reversion and uncovered interest parity: A revision of the Garman-Kohlhagen model. European Journal of Operational Research, 100(1), 41–59. doi:10.1016/S0377-2217(95)00366-5 Garman, M. B., & Kohlhagen, S. W. (1983). Foreign currency option values. Journal of International Money and -0.1 1.4 0 1.38 0.1 1.36 0.2 21.34 1.8 D iff er en ce 0.3 1.32 1.6 0.4 Strike Price 1.41.3 1.2 0.5 Time to Maturity (Year) 1.28 1 0.6 1.26 0.8 0.61.24 0.41.22 0.21.2 0 Subdiffusive FBS versus G-K FBS versus G-KFigure 4. Relative difference between the G  K, FBS, and subdiffusive FBS models for the out-of-the-money case. -0.2 1.2 0 1.15 0.2 21.1 0.4 1.8 D iff er en ce 1.05 1.6 0.6 Strike Price 1.41 0.8 1.2 Time to Maturity (Year) 0.95 1 1 0.80.9 0.6 0.40.85 0.20.8 0 Subdiffusive FBS versus G-K FBS versus G-KFigure 3. Relative difference between the G  K, FBS, and subdiffusive FBS models for the in-the-money case. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 9 of 14 52 Acta Wasaensia Finance, 2(3), 231–237. doi:10.1016/S0261-5606(83) 80001-1 Gu, H., Liang, J.-R., & Zhang, Y.-X. (2012). Time-changed geometric fractional Brownian motion and option pricing with transaction costs. Physica A: Statistical Mechanics and Its Applications, 391(15), 3971–3977. doi:10.1016/j.physa.2012.03.020 Guo, Z. (2017). Option pricing under the merton model of the short rate in subdiffusive Brownian motion regime. Journal of Statistical Computation and Simulation, 87(3), 519–529. doi:10.1080/ 00949655.2016.1218880 Guo, Z., & Yuan, H. (2014). Pricing European option under the time-changed mixed Brownian-fractional Brownian model. Physica A: Statistical Mechanics and Its Applications, 406, 73–79. doi:10.1016/j. physa.2014.03.032 Ho, T. S., Stapleton, R. C., & Subrahmanyam, M. G. (1995). Correlation risk, cross-market derivative products and portfolio performance. European Financial Management, 1(2), 105–124. doi:10.1111/j.1468- 036X.1995.tb00011.x Janczura, J., Orzeł, S., & Wyłomańska, A. (2011). Subordinated α-stable Ornstein–Uhlenbeck process as a tool for financial data description. Physica A: Statistical Mechanics and Its Applications, 390(23– 24), 4379–4387. doi:10.1016/j.physa.2011.07.007 Janicki, A., & Weron, A. (1993). Simulation and chaotic behavior of alpha-stable stochastic processes (Vol. 178). CRC Press. Janicki, A., & Weron, A. (1995). Simulation and chaotic behaviour of a-stable stochastic processes. Journal of the Royal Statistical Society-Series A Statistics in Society, 158(2), 339. Karipova, G., & Magdziarz, M. (2017). Pricing of basket options in subdiffusive fractional Black–Scholes model. Chaos, Solitons & Fractals, 102, 245–253. doi:10.1016/j.chaos.2017.05.013 Kumar, A., Wyłomańska, A., Połoczański, R., & Sundar, S. (2017). Fractional Brownian motion time-chan- ged by gamma and inverse gamma process. Physica A: Statistical Mechanics and Its Applications, 468, 648–667. doi:10.1016/j. physa.2016.10.060 Lux, T., & Marchesi, M. (1999). Scaling and criticality in a stochastic multi-agent model of a financial market. Nature, 397(6719), 498–500. doi:10.1038/17290 Magdziarz, M. (2009a). Black-Scholes formula in subdif- fusive regime. Journal of Statistical Physics, 136(3), 553–564. doi:10.1007/s10955-009-9791-4 Magdziarz, M. (2009b). Langevin picture of subdiffusion with infinitely divisible waiting times. Journal of Statistical Physics, 135(4), 763–772. doi:10.1007/ s10955-009-9751-z Magdziarz, M. (2009c). Stochastic representation of sub- diffusion processes with time-dependent drift. Stochastic Processes and Their Applications, 119(10), 3238–3252. doi:10.1016/j.spa.2009.05.006 Magdziarz, M. (2010). Path properties of subdiffusiona martingale approach. Stochastic Models, 26(2), 256– 271. doi:10.1080/15326341003756379 Magdziarz, M., Weron, A., & Weron, K. (2007). Fractional Fokker-Planck dynamics: Stochastic representation and computer simulation. Physical Review E, 75(1), 016708. doi:10.1103/PhysRevE.75.016708 Mantegna, R. N., Stanley, H. E. (1995). Scaling behaviour in the dynamics of an economic index. Nature, 376 (6535), 46–49. doi:10.1038/376046a0 Mastinšek, M. (2006). Discrete–time delta hedging and the Black–Scholes model with transaction costs. Mathematical Methods of Operations Research, 64(2), 227–236. doi:10.1007/s00186-006-0086-0 Metzler, R., & Klafter, J. (2000). The random walk’s guide to anomalous diffusion: A fractional dynamics approach. Physics Reports, 339(1), 1–77. doi:10.1016/ S0370-1573(00)00070-3 Necula, C. (2002).Option pricing in a fractional Brownian motion environment. Available at SSRN 1286833. Piryatinska, A., Saichev, A., & Woyczynski, W. (2005). Models of anomalous diffusion: The subdiffusive case. Physica A: Statistical Mechanics and Its Applications, 349(3–4), 375–420. doi:10.1016/j. physa.2004.11.003 Scalas, E., Gorenflo, R., & Mainardi, F. (2000). Fractional calculus and continuous-time finance. Physica A: Statistical Mechanics and Its Applications, 284(1–4), 376–384. doi:10.1016/S0378-4371(00)00255-7 Shokrollahi, F., & Kılıçman, A. (2014). Delta-hedging strategy and mixed fractional Brownian motion for pricing currency option. Mathematical Problems in Engineering, 501, 718768. Shokrollahi, F., & Kılıçman, A. (2014). Pricing currency option in a mixed fractional Brownian motion with jumps environment. Mathematical Problems in Engineering, 2014, 1–13. doi:10.1155/2014/858210 Shokrollahi, F., & Sottinen, T. (2017). Hedging in fractional Black–Scholes model with transaction costs. Statistics & Probability Letters, 130, 85–91. doi:10.1016/j.spl.2017.07.014 Sottinen, T., & Valkeila, E. (2003). On arbitrage and repli- cation in the fractional Black–Scholes pricing model. Statistics & Decisions/International Mathematical Journal for Stochastic Methods and Models, 21(2/ 2003), 93–108. Wang, X.-T. (2010). Scaling and long-range dependence in option pricing I: Pricing European option with transaction costs under the fractional Black–Scholes model. Physica A: Statistical Mechanics and Its Applications, 389(3), 438–444. doi:10.1016/j. physa.2009.09.041 Wang, X.-T., Zhu, E.-H., Tang, -M.-M., & Yan, H.-G. (2010). Scaling and long-range dependence in option pricing II: Pricing European option with transaction costs under the mixed Brownian–fractional Brownian model. Physica A: Statistical Mechanics and Its Applications, 389(3), 445–451. doi:10.1016/j. physa.2009.09.043 Xiao, W.-L., Zhang, W.-G., Zhang, X.-L., & Wang, Y.-L. (2010). Pricing currency options in a fractional Brownian motion with jumps. Economic Modelling, 27(5), 935–942. doi:10.1016/j. econmod.2010.05.010 Yang, Z. (2017). A new factorization of American frac- tional lookback option in a mixed jump-diffusion fractional Brownian motion environment. Journal of Modeling and Optimization, 9(1), 53–68. Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 10 of 14 Acta Wasaensia 53 Appendix Proof of Theorem 2.1. The movement of St on time interval ½t; tþ ΔtÞ of length Δt is ΔSt ¼ StþΔt  St ¼ Stðeðrdrf ÞΔTαðtÞþ σΔWα;HðtÞ  1Þ ¼ Stððrd  rf ÞΔTαðtÞ þ σΔWα;HðtÞ þ 1 2 ððrd  rf ÞΔTαðtÞ þ σΔWα;HðtÞÞ2Þ þ 1 6 Ste½θððrdrf ÞΔTαðtÞþ σΔWα;HðtÞÞ  ððrd  rf ÞΔTαðtÞ þ σΔWα;HðtÞÞ3; (4:1) here θ ¼ θðt;ΔtÞ 2 ð0;1Þ is a random variable corresponding to process St. Based on Lemmas 2.1 and 2.2 of Gu et al. (2012), we can get ððrd  rf ÞΔTαðtÞ þ σΔWα;HðtÞÞ2 ¼ ðoðΔtαεÞ þ oðΔtαHεÞÞ2 ¼ oðΔtαþ αH2εÞ þ oðΔt2αH2εÞ; (4:2) e½θððrdrf ÞΔTαðtÞþ σΔWα;HðtÞÞððrd  rf ÞΔTαðtÞ þ σΔWα;HðtÞÞ3 ¼ oðΔt3α3εÞ þ oðΔt2αþ αH3εÞ þ oðΔt2αHþ α3εÞ þ oðΔt3α3εÞ ¼ oðΔt3αH3εÞ; (4:3) oðΔtαþ αH2εÞ þ oðΔt3αH3εÞ ¼ oðΔtαþ αH2εÞ: (4:4) From the above equations, Equation (4.1) can be rewritten as follows ΔSt ¼ ðrd  rf ÞStΔTαðtÞ þ σStΔWα;HðtÞ þ 1 2 σ2StðΔWα;HðtÞÞ2 þ oðΔtαþ αH2εÞ: (4:5) By the assumption αHþ α>1, we obtain ΔSt ¼ ðrd  rf ÞStΔTαðtÞ þ σStΔWα;HðtÞ (4:6) Applying the Taylor expansion to Cðt; StÞ, we have ΔCðt; StÞ ¼ @C @t Δt þ @C @St ΔSt þ 12 @2C @S2t ΔS2t þ 1 2 @2C @t2 Δt2 þ @ 2C @St@t ΔtΔSt þ oðΔt3αHεÞ ¼ @C @t Δt þ @C @St ΔSt þ 12 @2C @S2t ΔS2t þ oðΔtÞ ¼ @C @t Δt þ ðrd  rf ÞSt @C @St ΔTαðtÞ þ σSt @C @St ΔWα;HðtÞ þ 1 2 σ2St @C @St ðΔWα;HðtÞÞ2 þ 1 2 σ2S2t @2C @S2t ðΔWα;HðtÞÞ2 þ oðΔtÞ: (4:7) From Equations (4.1)–(4.5), we obtain that @ 2C @S2t , @ 3C @S3t , @ 2C @C@t is oðΔt 1 2ð1HαÞε Þ and Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 11 of 14 54 Acta Wasaensia Δ @C @St ¼ @ 2C @St@t Δtþ @ 2C @S2t ΔSt þ 12 @3C @S3t ΔS2t þ oðΔtÞ; (4:8) and Δ @C @St :StþΔt ¼ σS2t @2C@S2t ΔWα;HðtÞ þ oðΔtÞ: (4:9) Moreover, from assumptions (iii) and (iv), it is found that the change in the value of portfolio t is Δt ¼ UtðΔSt þ rf StΔtÞ þ ΔFt  k2 ΔUtj jStþΔt¼ UtðΔSt þ rf StΔtÞ þ rdFtΔt  k2 ΔUtj jStþΔt þ oðΔtÞ; (4:10) where the number of bonds Ut is constant during time-step Δt. From assumption (v), Cðt; StÞ is replicated by portfolio ðtÞ. Thus, at time points Δt, 2Δt, 3Δt; :::; we have Cðt; StÞ ¼ UtSt þ Ft and Ut ¼ @C@St . Therefore, according to Equations (4.5)–(4.10), we have Δ ¼ @C @St ðrd  rf ÞStΔTαðtÞ þ σStΔWα;HðtÞ þ 1 2 σ2StðΔWα;HðtÞÞ2 þ rf StΔt  þ rdFtΔt k 2 Δ @C @St :StþΔt þ oðΔtÞ ¼ @C @St ðrd  rf ÞStΔTαðtÞ þ σStΔWα;HðtÞ þ 1 2 σ2StðΔWα;HðtÞÞ2 þ rf StΔt  þ C t; Stð Þ  St @C @St rdΔt k 2 σS2t @2C @S2t ΔWα;HðtÞ þ oðΔtÞ: (4:11) Consequently, Δ ΔC ¼ rdC rd  rf   St @C @St  @C @t Δt 1 2 σ2S2t @2C @S2t ΔWα;HðtÞ  2  k 2 σS2t @2C @S2t ΔWα;HðtÞ þ oðΔtÞ: (4:12) The time subscript, t, has been suppressed. As expected, using Equation (4.12), (iv), Remark 2.1, and (4.13), we infer EðΔ ΔCÞ ¼ rdC rd  rf   St @C @St  @C @t Δt  1 2 tα1 ΓðαÞ  2H Δt2Hσ2S2t @2C @S2t  1 2 ffiffiffi 2 π r kσS2t tα1 ΓðαÞ  H ΔtH @2C @S2t ¼ rdC ðrd  rf ÞSt @C @St  @C @t  1 2 tα1 ΓðαÞ  2H Δt2H1σ2S2t @2C @S2t 1 2 ffiffiffi 2 π r kσS2t tα1 ΓðαÞ  H ΔtH1 @2C @S2t ! Δt ¼ 0: (4:13) Thus, from Equation (4.13), we can derive Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 12 of 14 Acta Wasaensia 55 rdC ¼ ðrd  rf ÞSt @C @St þ @C @t þ 1 2 tα1 ΓðαÞ  2H Δt2H1σ2S2t @2C @S2t þ 1 2 ffiffiffi 2 π r kσS2t tα1 ΓðαÞ  H ΔtH1 @2C @S2t : (4:14) We define σ^2ðtÞ as follows: σ^2 ¼ σ2 t α1 ΓðαÞ  2H Δt2H1 þ ffiffiffi 2 π r kσ1 tα1 ΓðαÞ  H ΔtH1 ! : (4:15) where @ 2C @S2t is ever positive for the ordinary European currency call option without transaction costs, if the same conduct of @ 2C @S2t is postulated here and σ^ðtÞ remains fixed during the time-step ½t;ΔtÞ. Then, from Equations (4.14) and (4.15), we obtain @C @t þ ðrd  rf ÞSt @C @St þ 1 2 σ^2S2t @2C @S2t  rdC ¼ 0: (4:16) Followed by C ¼ Cðt; StÞ ¼ Sterf ðTtÞΦðd1Þ  KerdðTtÞΦðd2Þ; (4:17) and d1 ¼ ln St K   þðrdrf ÞðTtÞþσ^22 ðTtÞ σ^ ffiffiffiffiffiffi Tt p ; d2 ¼ d1  σ^ ffiffiffiffiffiffiffiffiffiffi T  tp : (4:18) Proof of Theorem 2.2. First, we derive a general formula. Let y be one of the influence factors. Thus @C @y ¼ @Ste  rfð ÞðTtÞ @y Φðd1Þ þ Sterf ðTtÞ @Φðd1Þ@y  @KerdðTtÞ@y Φðd2Þ  KerdðTtÞ @Φðd2Þ@y (4:19) But @Φðd2Þ @y ¼ Φ0ðd2Þ @d2 @y ¼ 1ffiffiffiffiffi 2π p e d2 2 2 @d2 @y ¼ 1ffiffiffiffiffi 2π p exp ðd1  bσ ffiffiffiffiffiffiffiffiffiffiT  tp Þ2 2 ! @d2 @y ¼ 1ffiffiffiffiffi 2π p e d2 1 2 expðd1bσ ffiffiffiffiffiffiffiffiffiffiffiffiT  tÞp Þ exp bσ2ðT  tÞ2 ! @d2 @y ¼ 1ffiffiffiffiffi 2π p e d2 1 2 exp ln St K þ ðrd  rf ÞðT  tÞ @d2 @y ¼ 1ffiffiffiffiffi 2π p e d2 1 2 S K expððrd  rf ÞðT  tÞÞ @d2 @y : (4:20) Then Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 13 of 14 56 Acta Wasaensia @C @y ¼ @Ste ðrf ÞðTtÞ @y Φðd1Þ  @Ke rdðTtÞ @y Φðd2Þ þ Sterf ðTtÞΦ0ðd1Þ @σ^ ffiffiffiffiffiffiffi TtÞ p @y : (4:21) Substituting in (4.21), we get the desired Greeks. ©2018 The Author(s). This open access article is distributed under a Creative Commons Attribution (CC-BY) 4.0 license. You are free to: Share — copy and redistribute the material in any medium or format. Adapt — remix, transform, and build upon the material for any purpose, even commercially. The licensor cannot revoke these freedoms as long as you follow the license terms. Under the following terms: Attribution — You must give appropriate credit, provide a link to the license, and indicate if changes were made. You may do so in any reasonable manner, but not in any way that suggests the licensor endorses you or your use. No additional restrictions Youmay not apply legal terms or technological measures that legally restrict others from doing anything the license permits. Cogent Mathematics & Statistics (ISSN: 2574-2558) is published by Cogent OA, part of Taylor & Francis Group. Publishing with Cogent OA ensures: • Immediate, universal access to your article on publication • High visibility and discoverability via the Cogent OA website as well as Taylor & Francis Online • Download and citation statistics for your article • Rapid online publication • Input from, and dialog with, expert editors and editorial boards • Retention of full copyright of your article • Guaranteed legacy preservation of your article • Discounts and waivers for authors in developing regions Submit your manuscript to a Cogent OA journal at www.CogentOA.com Shokrollahi, Cogent Mathematics & Statistics (2018), 5: 1470145 https://doi.org/10.1080/25742558.2018.1470145 Page 14 of 14 Acta Wasaensia 57 Journal of Mathematical Finance, 2018, 8, 623-639 http://www.scirp.org/journal/jmf ISSN Online: 2162-2442 ISSN Print: 2162-2434 DOI: 10.4236/jmf.2018.84040 Nov. 7, 2018 623 Journal of Mathematical Finance Mixed Fractional Merton Model to Evaluate European Options with Transaction Costs Foad Shokrollahi Department of Mathematics and Statistics, University of Vaasa, Vaasa, Finland Abstract This paper deals with the problem of discrete-time option pricing by the mixed fractional version of Merton model with transaction costs. By a mean-self-financing delta hedging argument in a discrete-time setting, a Eu- ropean call option pricing formula is obtained. We also investigate the effect of the time-step tG and the Hurst parameter H on our pricing option mod- el, which reveals that these parameters have high impact on option pricing. The properties of this model are also explained. Keywords Transaction Costs, Mixed Fractional Brownian Motion, European Option, Merton Model 1. Introduction Over the last few years, the financial markets have been regarded as complex and nonlinear dynamic systems. A series of studies has found that many financial market time series display scaling laws and long-range dependence. Therefore, it has been proposed that the Brownian motion in the classical Black-Scholes (BS) model [1] should be replaced by a process with long-range dependence. Nowadays, the BS model is the one most commonly used for analyzing financial data, and some scholars have presented modified forms of the BS model which have influential and significant outcomes on option pricing. However, they are still theoretical adaptations and not necessarily consistent with the empirical features of financial return series, such as nonnormality, long-range dependence, etc. For example, some scholars [2] [3] [4] [5] [6] have showed that returns are of long-range (or short-range) dependence, which suggests strong time-correlations between different events at different time How to cite this paper: Shokrollahi, F. (2018) Mixed Fractional Merton Model to Evaluate European Options with Transac- tion Costs. Journal of Mathematical Finance, 8, 623-639. https://doi.org/10.4236/jmf.2018.84040 Received: May 7, 2018 Accepted: November 4, 2018 Published: November 7, 2018 Copyright © 2018 by author and Scientific Research Publishing Inc. This work is licensed under the Creative Commons Attribution-NonCommercial International License (CC BY-NC 4.0). http://creativecommons.org/licenses/by-nc/4.0/ Open Access 58 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 624 Journal of Mathematical Finance scales [7] [8] [9]. In the search for better models for describing long-range dependence in financial return series, a mixed fractional Brownian (MFBM) model has been proposed as an improvement of the classical BS model [10]-[18]. The advantage of using the MFBM is that the markets are free of arbitrage. Moreover, Cheridito [10] has proved that, for 3 ,1 4 H § ·¨ ¸© ¹ , the MFBM is equivalent to one with Brownian motion, and hence time-step and long-range dependence in return series have no impact on option pricing in a complete financial market without transaction costs. In addition, a number of empirical studies show that the paths of asset prices are discontinuous and that there are jumps in asset prices, both in the stock market and foreign exchange [9] [19] [20] [21] [22]. The above researches have an important implication for option pricing. Merton [23] created a revolution in option pricing when the underlying asset was governed by a diffusion process. Based on this theory, Kou [24], Cont and Tankov [25] also considered the problems of pricing options under a jump diffusion environment in a larger setting. In this paper, to capture jumps or discontinuities, fluctuations and to take into account the long memory property of financial markets, a mixed fractional version of the Merton model is introduced, which is based on a combination of Poisson jumps and MFBM. The mixed fractional Merton (MFM) model is based on the assumption that the underlying asset price is generated by a two-part stochastic process: 1) small, continuous price movements are generated by a MFBM, and 2) large, infrequent price jumps are generated by a Poisson process. This two-part process is intuitively appealing, as it is consistent with an impressive market in which major information arrives infrequently and randomly. This process may provide a description for empirically observed distributions of exchange rate changes that are skewed, leptokurtic, have long memory and fatter tails than comparable normal distributions and apparent nonstationary variance. Further, we will show the impact of the time-step and long-range dependence in return series exactly on option pricing, regardless of whether proportional transaction costs are considered or not in a discrete time setting. Leland [26] is a pioneer scholar, who investigated option replication where transaction costs exist in a discrete time setting. In this view, the arbitrage-free arguments presented by Black and Scholes [1] are not applicable in a model where transaction costs occur at all moments of trading of the stock or bond. The problem is that perfect replication incurs an infinite number of transaction costs because of the infinite variation which exists in the geometric Brownian motion. In this regard, a delta hedge strategy is constructed in accordance with revision conducted a discrete number of times. Transaction costs lead to the failure of the no arbitrage principle and the continuous time trade in general: instead of no arbitrage, the principle of hedge pricing, according to which the price of an option is defined as the minimum level of initial wealth needed to hedge the option, comes into force. Acta Wasaensia 59 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 625 Journal of Mathematical Finance According to the empirical findings obtained before and the views of behavioral finance and econophysics, we are motivated to examine the problem that exists in option pricing, while the dynamics of price tS follows a mixed fractional jump-diffusion process under the transaction costs. We assume that tS satisfies ln 0e .H H t t B t B t N J tS S P V V   (1.1) where 0 , ,S P V and HV are fixed; B t is a Brownian motion; HB t is a fractional Brownian motion with Hurst parameter 3 ,1 4 H § ·¨ ¸© ¹ ; tN is a Poisson process with intensity 0O ! ; and J is a positive random variable. We assume that , ,H tB t B t N and J are independent. This paper is organized into several sections. In Section 2, we will study the problem of option pricing with transaction costs by applying delta hedging strategy. In addition, a new framework for pricing European option is obtained when the stock price tS is satisfied in Equation (1.1). Section 3 is devoted to empirical studies and simulations to show the performance of the MFM model. A conclusion is presented in Section 4. 2. Pricing Option by Mixed Fractional Version of Merton Model with Transaction Costs Suppose ^ ` 0tB t t be a standard Brownian motion and ^ ` 0H tB t t be a fractional Brownian motion with the Hurst parameter 3 ,1 4 H § ·¨ ¸© ¹ , both defined on complete probability space , , ,t P: F F , the absolute price jump size J is a nonnegative random variable drawn from lognormal distribution, i.e. ln ,J JJ N P V , which implies 2 2 222~ e ,e e 1JJ J J JJ Lognormal VP P V V § ·¨ ¸¨ ¸© ¹ and a Poisson process 0t tN N t with rate O . Additionally, the processes , ,HB B N and J are independent, P is the real world probability measure and > @0,t t TF denotes the P-augmentation of filtration generated by , ,HB B tW W W d . The objective of this section is to derive a stock pricing formula under transaction costs in a discrete time setting. Consider ,D S -market with a bond tD and a stock tS , where 0e . rt tD D (2.1) and ln 0 0 0e , , , , , , .H H t t B t B t N J t HS S R D S t R P V V P V V      (2.2) The groundwork of modeling the effects of transaction costs was done by Leland [26]. He adopted the hedging strategy of rehedging at every time-step 60 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 626 Journal of Mathematical Finance tG . That is, with every tG the portfolio is rebalanced, whether or not this is optimal in any sense. In the following proportional transaction cost option pricing model, we follow the other usual assumptions in the Black-Scholes model, but with the following exceptions: 1) The price tS of the underlying stock at time t satisfies Equation (2.2). 2) The portfolio is revised every tG where tG is a finite and fixed, small time-step. 3) Transaction costs are proportional to the value of the transaction in the underlying. Let k denote the round trip transaction cost per unit dollar of transaction. Suppose 0U ! shares are bought 0U ! or sold 0U  at the price tS , then the transaction cost is given by 2 t k U S in either buying or selling, where k is a constant. The value of k will depend on the individual investor. In the MFM model, where transaction costs are incurred at every time the stock or the bond is traded, the no arbitrage argument used by Black and Scholes no longer applies. The problem is that due to the infinite variation of the MFBM, perfect replication incurs an infinite amount of transaction costs. 4) The hedge portfolio has an expected return equal to that from an option. This is exactly the same valuation policy as earlier on discrete hedging with no transaction costs. 5) Traditional economics assumes that traders are rational and maximize their utility. However, if their behaviour is assumed to be boundedly rational, the traders' decisions can be explained both by their reaction to the past stock price, according to a standard speculative behaviour, and by imitation of other traders’ past decisions, according to common evidence in social psychology. It is well known that the delta-hedging strategy plays a central role in the theory of option pricing and that it is popularly used on the trading floor. Therefore, based on the availability heuristic, suggested by Tversky and Kahneman [27], traders are assumed to follow, anchor, and imitate the Black-Scholes delta-hedging strategy to price an option. In this case, delta-hedging argument is a partial and imperfect hedging strategy, which does not eliminate all of the risk. However, as mentioned in the paper [28], in most models of stock fluctuations, except for very special cases, risk in option trading cannot be eliminated and strict arbitrage opportunities do not exist, whatever be the price of the option. The risk cannot be eliminated is furthermore the fundamental reason for the very existence of option markets. Delta hedging is an options strategy that aims to reduce, or hedge, the risk associated with price movements in the underlying asset, by offsetting long and short positions. For example, a long call position may be delta hedged by shorting the underlying stock. This strategy is based on the change in premium, or price of option, caused by a change in the price of the underlying security. In this section we use the delta hedging strategy to obtain a pricing formula for European call option. Let the price of European call option be denoted with expiration T and strike Acta Wasaensia 61 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 627 Journal of Mathematical Finance price K by , tC t S with boundary conditions: , , ,0 0, , as .T T t t tC T S S K C t C t S S S  o of (2.3) Then, , tC t S is derived by the following theorem. Theorem 2.1. The price at every > @0,t T of a European call option with strike price K that matures at time T is given by 1 2 0 e , e . ! nT t r T t t t n T t C t S S d K d n O O I I c f   c  ª º ¬ ¼¦ (2.4) Moreover, , tC t S satisfies the following equation > @ 2 2 2 2 ˆ , , 2 1 0, t t t t t t t t SC C CrS rC E C t JS C t S t S S CE J S S V O O w w w ª º    ¬ ¼w w w w  w (2.5) where 2 1 2 1 ln 2 , , t n n n n S r T t T t Kd d d T t T t V VV § ·    ¨ ¸© ¹   (2.6) 2 2 2 22 ˆe , , J J J n nE J T t VP VO O O V Vc   (2.7) 2 2 2 2ln 1 e 1 , J J J J n n n E J r r E J r T t T t VP VP O O  § ·¨ ¸§ · © ¹¨ ¸      ¨ ¸ © ¹ (2.8) 22 1 2 22 2 2 22ˆ ,ʌ H H H Ht k t signt VV V V G V GG  § ·    *¨ ¸© ¹ (2.9) sign * is the signum function of 2 2 t C S w w ; n is the number of prices jumps; tG is a small and fixed time-step; k is the transaction costs and .I is the cumulative normal distribution. Moreover, using the put call parity, we can easily obtain the valuation model for a put currency option, which is provided by the following corollary. Corollary 2.1. The value of European put option with transaction costs is given by 2 1 0 e , e . ! nT t r T t t t n T t P t S K d S d n O O I I c f   c  ª º   ¬ ¼¦ 3. Properties of Pricing Formula In this section, we present the properties of MFM’s log-return density. The effects of Hurst parameter and time-step on our modified volatility 2nV are also discussed in the discrete time and continuous time cases. Then we show that these parameters play a significant role in a discrete time setting, both with and 62 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 628 Journal of Mathematical Finance without transaction costs. 3.1. Log-Return Density In the case of MFM the log return jump size is assumed to be 2ln ~ ,i i J JY J N P V and the probability density of log return lnt tx S S is achieved as a quickly converging series of the following form: 0 |t t n t n P x A P N n P x A N n f   ¦ 2 2 2 2 0 e ; , , ! nt H t t J H J n t P x N x t n t t n n O O P P V V V f   ¦ (3.1) where 2 2 2 2 2 2 2 2 22 2 2 2 ; , 1 exp 22ʌ H t J H J t J HH H JH J N x t n t t n x t n t t nt t n P P V V V P P V V VV V V    ª º « » « »   ¬ ¼ (3.2) The term e ! nt t t P N n n O O is the probability that the asset price jumps n times during the time interval of length t. And 2 2 2 2| ; , Hn t t J H JP x A N n N x t n t t nP P V V V    is the mixed fractional normal density of log-return. It supposes that the asset price jumps i times in the time interval of t. As a result, in the MFM model, the log-return density can be described as the weighted average of the mixed fractional normal density by the probability that the asset price jumps n times. The outstanding properties of log-return density tP x are observed in the MFM. Firstly, the JP sign refers to the expected log-return jump size, ln JE Y E J P , which indicates the skewness sign. If 0JP  , the log-return density tP x shows negatively skewed, and if 0JP , it is symmetric as displayed in Figure 1 (Table 1). Secondly, larger value of intensity O (which means that jumps are expected to occur more frequently) makes the density fatter-tailed as illustrated in Figure 2. Note that the excess kurtosis in the case 20O is much smaller than in the case 1O or 10O . This is because excess kurtosis is a standardized measure (by standard deviation) (Table 2). 3.2. The Impact of Parameters Mantegna and Stanley [29] as pioneer scholars proposed the scaling invariance method from the complex science of economic systems which led to numerous investigations into scaling laws in finance. The major question in economics is whether the price impact of scaling law and long-range dependence is significant in option pricing. The answer to this question is assured. For instance, one of the significant issues in finance concerning the modeling of high-frequency data is related to analyzing the volatility in different time scales. Acta Wasaensia 63 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 629 Journal of Mathematical Finance Figure 1. MFM’s log-return density. Fixed parameters are 0.25V , 0.25HV , 0.76H , 0.1JV , 0JP , 0.009P , and 0.5t . Figure 2. MFM’s log-return density. Fixed parameters are 0.25V , 0.25HV , 0.76H , 0.1JV , 3O , 0.009P , and 0.5t . Table 1. Annualized moments of Merton’s log-return density in Figure 1. Model Mean Standard Deviation Skewness Excess Kurtosis 0.4JP  −1.1910 0.6161 −0.5082 0.2806 0JP 0.0090 0.1361 0 0.706 0.4JP 1.2090 0.6161 0.5082 0.2806 64 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 630 Journal of Mathematical Finance Table 2. Annualized moments of Merton’s log-return density in Figure 2. Model Mean Standard Deviation Skewness Excess Kurtosis 1O 0.0040 0.1161 0 0.0223 10O −0.0411 0.2061 0 0.706 20O −0.0913 0.3061 0 0.0640 Remark 3.1. In a continuous time setting 0, 0tG O z without transaction costs the implied volatility is 2 2 2ˆ Jn n T t VV V   , thus the option value is similar to the Merton jump diffusion model [19]. Moreover, if 0tG in the absence of transaction costs and jump case, the MFM model reduces to the BS model 2 2 2 2 0,2 t t t t SC C CrS rC t S S Vw w w   w w w (3.3) which shows that the Hurst parameter H and time-step tG have no effect on option pricing model in a continuous time setting 0tG . Remark 3.2. In a discrete time setting without transaction costs 0, 0k tG z , if jump occurs, the modified volatility is 22 12 2 2ˆ H Jn H nt T t VV V V G     and when jump does not occur 0O , from Equation (2.5), we have 2 22 12 2 2 0,2H tt Ht tSC C CrS t rCt S SV V G w w w    w w w (3.4) which demonstrates that the delta hedging strategy in a discrete time case is fundamentally different in comparison with a continuous time case. It also indicates that the scaling exponent 2 1H  and time-step tG play a significant role in option pricing theory. Figure 3 illustrates the impacts of the time-step, Hurst parameter, mean jump, and jump intensity on our option pricing model. Remark 3.3. From [30] we infer there exists 10,t M G § ·¨ ¸© ¹ such that 2 10, ˆmin , t M G V§ ·¨ ¸© ¹ (3.5) Holds, where 1M ! , k is small enough 22 1 2 22 2 2 22ˆ .ʌ 2 H H H Ht k tt VV V V G V GG  § ·   ¨ ¸© ¹ (3.6) Indeed, 2 2 1 2 22 2 1 1 2 41 2 2222 2 ʌ 22 .ʌ H H H H H H H H t k t t t k t t VV G V GG VV G V GG     § · ¨ ¸© ¹ § ·§ ·t ¨ ¸¨ ¸¨ ¸© ¹© ¹ (3.7) Acta Wasaensia 65 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 631 Journal of Mathematical Finance Figure 3. Modified volatility. Fixed parameters are 0.1V , 0.1HV , 0.76H , 0.03JV , 0.2T , 0.2T , and 0.1t . Set 22 2 2 12 2 2 2 .ʌ H H H H kt t t VVV G G G   § ·¨ ¸ ¨ ¸© ¹ (3.8) Thus 22 2 2 2 22 2 2 8 ʌ ʌ ʌ . 2 H H k k k t t V G V G § · ¨ ¸© ¹ (3.9) Suppose 22 2 2 2 2 2 2 2 8 ʌ ʌ ʌ . 2 H H k k k x f x x V V § · ¨ ¸© ¹  (3.10) Since 0 0f  and 22 2 2 2 2 2 2 2 8 1 ʌ ʌ ʌ1 1 0, 2 H H k k k M f M M V V § · ¨ ¸§ · § · © ¹  !¨ ¸ ¨ ¸© ¹ © ¹ (3.11) as k is small enough. Hence, there exists a 10,t M G § ·¨ ¸© ¹ such that 2 10, ˆmin t M G V§ ·¨ ¸© ¹ holds. Suppose 2 2 10, ˆ ˆmin min , t M G V V§ ·¨ ¸© ¹ (3.12) so 22 2 2 1 10, 0, ˆmin min min .n n t t M M n T tG G GV V V§ · § · ¨ ¸ ¨ ¸© ¹ © ¹   (3.13) Then the minimal price of an option with respect to transaction costs is displayed as min , tC t S with 2 minnV in Equation (2.4). min , tC t S can be 66 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 632 Journal of Mathematical Finance applied to the real price of an option. 4. Conclusion To capture the long memory and discontinuous property, this article focuses on the problem of pricing European option in a mixed fractional Merton environment without using the arbitrage argument. We obtain a mixed fractional version of Merton model for pricing European option with transaction costs. Some properties of mixed fractional Merton’s log-return density are discussed. Moreover, we derive that the Hurst parameter H and time-step tG play a significant role in pricing option in a discrete time setting for cases both with and without transaction costs. But these parameters have no impact on option pricing in a continuous time setting. Conflicts of Interest The authors declare no conflicts of interest regarding the publication of this pa- per. References [1] Black, F. and Scholes, M. (1973) The Pricing of Options and Corporate Liabilities. The Journal of Political Economy, 81, 637-654. https://doi.org/10.1086/260062 [2] Willinger, W., Taqqu, M.S. and Teverovsky, V. (1999) Stock Market Prices and Long-Range Dependence. Finance and Stochastics, 3, 1-13. https://doi.org/10.1007/s007800050049 [3] Ozdemir, Z.A. (2009) Linkages between International Stock Markets: A Multivariate Long-Memory Approach. Physica A: Statistical Mechanics and Its Applications, 388, 2461-2468. https://doi.org/10.1016/j.physa.2009.02.023 [4] Mariani, M., Florescu, I., Varela, M.B. and Ncheuguim, E. (2009) Long Correlations and Levy Models Applied to the Study of Memory Effects in High Frequency (Tick) Data. Physica A: Statistical Mechanics and Its Applications, 388, 1659-1664. https://doi.org/10.1016/j.physa.2008.12.038 [5] Sottinen, T. and Valkeila, E. (2003) On Arbitrage and Replication in the Fractional Black-Scholes Pricing Model. Statistics & Decisions/International Mathematical Journal for Stochastic Methods and Models, 21, 93-108. [6] Sottinen, T. and Viitasaari, L. (2016) Pathwise Integrals and Itô-Tanaka Formula for Gaussian Processes. Journal of Theoretical Probability, 29, 590-616. https://doi.org/10.1007/s10959-014-0588-2 [7] Cajueiro, D.O. and Tabak, B.M. (2007) Long-Range Dependence and Multifractality in the Term Structure of Libor Interest Rates. Physica A: Statistical Mechanics and Its Applications, 373, 603-614. https://doi.org/10.1016/j.physa.2006.04.110 [8] Cajueiro, D.O. and Tabak, B.M. (2007) Long-Range Dependence and Market Structure. Chaos, Solitons & Fractals, 31, 995-1000. https://doi.org/10.1016/j.chaos.2005.10.077 [9] Mandelbrot, B.B. and Stewart, I. (1998) Fractals and Scaling in Finance. Nature, 391, 758-758. [10] Cheridito, P. (2001) Mixed Fractional Brownian Motion. Bernoulli, 7, 913-934. https://doi.org/10.2307/3318626 Acta Wasaensia 67 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 633 Journal of Mathematical Finance [11] Mishura, Y.S. and Valkeila, E. (2002) The Absence of Arbitrage in a Mixed Brow- nian-Fractional Brownian Model. Trudy Matematicheskogo Instituta imeni VA Steklova, 237, 224-233. [12] Mishura, I.S. and Mishura, Y. (2008) Stochastic Calculus for Fractional Brownian Motion and Related Processes. Vol. 1929, Springer Science & Business Media, Kyiv. https://doi.org/10.1007/978-3-540-75873-0 [13] Shokrollahi, F. and Kılıऊman, A. (2014) Pricing Currency Option in a Mixed Frac- tional Brownian Motion with Jumps Environment. Mathematical Problems in En- gineering, 2014, 13. [14] Shokrollahi, F. and Kılıऊman, A. (2015) Actuarial Approach in a Mixed Fractional Brownian Motion with Jumps Environment for Pricing Currency Option. Advances in Difference Equations, 2015, 1-8. https://doi.org/10.1186/s13662-015-0590-8 [15] Shokrollahi, F. and Kılıऊman, A. (2014) Delta-Hedging Strategy and Mixed Frac- tional Brownian Motion for Pricing Currency Option. Mathematical Problems in Engineering, 501, Article ID: 718768. [16] Shokrollahi, F., Kılıऊman, A. and Magdziarz, M. (2016) Pricing European Options and Currency Options by Time Changed Mixed Fractional Brownian Motion with Transaction Costs. International Journal of Financial Engineering, 3, 1650003. [17] Xiao, W.-L., Zhang, W.-G., Zhang, X. and Zhang, X. (2012) Pricing Model for Eq- uity Warrants in a Mixed Fractional Brownian Environment and Its Algorithm. Physica A: Statistical Mechanics and Its Applications, 391, 6418-6431. https://doi.org/10.1016/j.physa.2012.07.041 [18] Cont, R. (2005) Long Range Dependence in Financial Markets. In: Fractals in Engi- neering, Springer, Berlin, 159-179. [19] Merton, R.C. (1976) Option Pricing When Underlying Stock Returns Are Discon- tinuous. Journal of Financial Economics, 3, 125-144. https://doi.org/10.1016/0304-405X(76)90022-2 [20] Jarrow, R. and Rosenfeld, E. (1988) Jump Risks and Intemporal Capital Asset Pric- ing Model. The Journal of Business, 57, 337-351. https://doi.org/10.1086/296267 [21] Ball, C.A. and Torous, W.N. (1985) On Jumps in Common Stock Prices and Their Impact on Call Option Pricing. The Journal of Finance, 40, 155-173. https://doi.org/10.1111/j.1540-6261.1985.tb04942.x [22] Shokrollahi, F. and Kılıऊman, A. (2016) The Valuation of Currency Options by Fractional Brownian Motion. Springer Plus, 5, 1145. https://doi.org/10.1186/s40064-016-2784-2 [23] Merton, R.C. (1973) Theory of Rational Option Pricing. The Bell Journal of Eco- nomics and Management Science, 4, 141-183. [24] Kou, S.G. (2002) A Jump-Diffusion Model for Option Pricing. Management Science, 48, 1086-1101. https://doi.org/10.1287/mnsc.48.8.1086.166 [25] Cont, R. and Tankov, P. (2003) Calibration of Jump-Diffusion Option-Pricing Models: A Robust Nonparametric Approach. Journal of Computational Finance, 7, 1-49. https://doi.org/10.21314/JCF.2004.123 [26] Leland, H.E. (1985) Option Pricing and Replication with Transactions Costs. The Journal of Finance, 40, 1283-1301. https://doi.org/10.1111/j.1540-6261.1985.tb02383.x [27] Tversky, A. and Kahneman, D. (1973) Availability: A Heuristic for Judging Fre- quency and Probability. Cognitive Psychology, 5, 207-232. [28] Wang, X.-T., Yan, H.-G., Tang, M.-M. and Zhu, E.-H. (2010) Scaling and 68 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 634 Journal of Mathematical Finance Long-Range Dependence in Option Pricing III: A Fractional Version of the Merton Model with Transaction Costs. Physica A: Statistical Mechanics and Its Applica- tions, 389, 452-458. https://doi.org/10.1016/j.physa.2009.09.044 [29] Mantegna, R.N. and Stanley, H.E. (1995) Scaling Behaviour in the Dynamics of an Economic Index. Nature, 376, 46-49. https://doi.org/10.1038/376046a0 [30] Wang, X.-T., Zhu, E.-H., Tang, M.-M. and Yan, H.-G. (2010) Scaling and Long-Range Dependence in Option Pricing II: Pricing European Option with Transaction Costs under the Mixed Brownian-Fractional Brownian Model. Physica A: Statistical Mechanics and Its Applications, 389, 445-451. https://doi.org/10.1016/j.physa.2009.09.043 Acta Wasaensia 69 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 635 Journal of Mathematical Finance Appendix Proof of Theorem 2.1. We consider a replicating portfolio with t\ units of financial underlying asset and one unit of the option. Then, the value of the portfolio at time t is , .t t tP t S C t S\  (4.1) Now, the movement in tS and tP is considered under discrete time interval tG . In view of this, we suppose that trading takes place at the specific time points of t and t tG . It can be said that the number of shares through the use of delta-hedging strategy and the present stock price tS are constantly held during the rebalancing interval > ,t t tG . Then, the movement in the value of the portfolio after time interval tG is defined as follows: , . 2t t t t kP t S C t S t SG \ G G G\   (4.2) where tSG is the movement of the underlying stock price, tG\ is the movement of the number of units of stock held in the portfolio, and tPG is the change in the value of the portfolio. Since the time-step tG and the asset change are both small, according to Taylor’s formulae we have if 0tNG with probability 1 tOG , so 2 3 2 e , 6 H H t t t t t H H H H t B t B tt H H SS S t S B t S B t t B t B t S t B t B tT PG VG V G G PG GV GV PG VG V G PG VG V G ª º¬ ¼         (4.3) where , ,t w wT T : , and 0 1T  . Since B t and HB t are continuous, then from [28] we have 1 1log ,HHt B t O t tG G G G § · ¨ ¸¨ ¸© ¹ (4.4) 32 1log ,t B t O t t G G G G § · ¨ ¸¨ ¸© ¹ (4.5) 0 as 0, HB t t B t G GG o o (4.6) and 3 53 1 2 1 2 3 33 31 12 22 2 e log log log log . H Ht B t B t H Ht B t B t O t O t t O t t O t t O t t T PG VG V G PG VG V G G G G G G G G G G  ª º¬ ¼      ª º¬ ¼ § ·  ¨ ¸© ¹ § · § · ¨ ¸ ¨ ¸¨ ¸ ¨ ¸© ¹ © ¹ Thus, we can get 32 12 log , 2 t t t H H t H H S S t S B t B t S B t B t O t t G P G VG V G VG V G G G   ª º¬ ¼ § ·  ª º ¨ ¸¬ ¼ © ¹ (4.7) 70 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 636 Journal of Mathematical Finance 322 12 2 log ,t t H HS S B t B t O t tG VG V G G G § ·  ª º ¨ ¸¬ ¼ © ¹ (4.8) 2 32 1221, log ,2t t tt t C C CC t S t S S O t t t S S G G G G G G w w w § ·    ¨ ¸w w w © ¹ (4.9) and 2 32 1221 log .2t tt tt t S S O t tt S S \ \ \G\ G G G G G w w w § ·    ¨ ¸w w w © ¹ (4.10) If 1tNG with probability tOG and the jump of tN in > @,t t t ' is assumed to occur at current time t, then ln 0e ,H H t B t B t J t S S P V V    (4.11) ln 0e ,H H t t B t t B t t J t tS S P G V G V G G        (4.12) e 1 ,H Ht B t B tt tt t tS S S S P V V GG      ª º  ¬ ¼ (4.13) e 1H H t t t t t t tt t t B t B t tt t S S S S S S S S S S G G P V V G             ª º   ¬ ¼ (4.14) 2 32 122 , , , , , , , 1 log , 2 t t t tt t tt t t t t C t S C t t S C t S C t S C t S C CC t S C t S t S t S C S O t t S GG G G G G G G              w w   w w w § ·  ¨ ¸w © ¹ (4.15) 2 32 122 , , , , , ,1 log , 2 t t t tt t t t t t t S t S t S t S t S t S t S t S S O t t S \ \G\ \ \ G G \ G G G          w w   w w w § ·  ¨ ¸w © ¹ (4.16) where t tt tS S SGG    . Based on the above assumptions iv and v, we have t t t t P DE P D G G§ · ¨ ¸© ¹ , i.e. 2t tE P rP O tG G  . Then 2 32 1 2 2 2 2 1 , 2 e 1 , , 1 log 2 , , , , 2 ,1 2 H H t t t t B t B t t tt t t t t t t t tt tt t t t t kSt E S C t S t tE S t S S t C t S C t S C C Ct S S O t t t S S t S t SkS t S t S t S t S t S S S P V V OG \G G G\ OG \ \ G G G G G \ \\ \ G G \ G                  ª º  « »¬ ¼ ª     «¬ ·w w w § ·    ¸¨ ¸w w w © ¹¹ w w   w w w w 32 1 2 log ,tt O t t rP tG G G  º§ · »¨ ¸© ¹¼ Acta Wasaensia 71 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 637 Journal of Mathematical Finance i.e. 2 32 1 2 2 1 , e 1 , , , , ,1 log 1 2 2 , , , 2 H Ht B t B t t t t t t t tt t t t t t t t t tt t E S C t S tE S t C t S C t S S S t C t S C t S t S t S C t S kSS O t t t E t S tkStE t S t S P V VOG \G G OG \ \ G G G G G OG G\ \OG \ \               ªª º   «¬ ¼ ¬ § w w¨     ¨ w w© º·w § · ª º»¸   ¨ ¸ « »»¸w ¬ ¼© ¹¹¼ wª  «¬ 2 32 1 222 , ,1 log , 2 t t t t t tt t S t S t S t S t S S O t t rP t O t S \G G \ G G G G G         ww w w º§ ·  »¨ ¸w © ¹ ¼ where the current stock price tS is given. Since 2 32 1 2 2 2 2 2 3 1 2 1 , 2 2 2 , , , ,1 log 2 , , ,1 1 2 2 log , 2 t tt t t t t t t t t t t t tt t t t t t t t kS t kSkSE t E t t S t S t S t S t S t S t S S O t t S t S t S t St kS E t S S t S S t kS O t t t S OG OGG\ G\ \ \ \\ G G \ G G G \ \ \OG G G G OGG G \           ª º « »¬ ¼ w w  w w w § ·  ¨ ¸w © ¹ w w w  w w w § · ¨ ¸© ¹ , , tt t S t S t t \\ G w  w 2 32 1 2 2 2 2 22 2 , ,1 log 2 , , 2 2 2 , , , 2 ʌ 2 t t t t t t t t H H tt t Ht t H tt t t S t S S S O t t S S kS k S tE B t B t t S t S O t S kS k S tt t t S t S O t S \ \G G G G O G\ VG V G \ \ G O G\ V G V G \ \ G          w w § ·   ¨ ¸w w © ¹ w|    w w    w (4.17) and t C S \ w w , from Equations (4.1) - (4.17), we can get > @ 2 2 2 12 2 2 2 2 2 2 22 2 2 , , 1 2 0. 2 ʌ Ht t H t t t t t t t Ht H t SC C CrS t t S S CrC E C t J S C t S E J S S kS Ct t O t t S V V G O O V V G G GG   ªw w w  « w w w¬ wª º    ¬ ¼ w º§ · w »   ¨ ¸ w »© ¹ ¼ (4.18) 72 Acta Wasaensia F. Shokrollahi DOI: 10.4236/jmf.2018.84040 638 Journal of Mathematical Finance Hence, we assume that > @ 2 2 2 12 2 2 2 2 2 2 22 2 2 , , 1 2 0. 2 ʌ Ht t H t t t t t t t Ht H t SC C CrS t t S S CrC E C t J S C t S E J S S kS Ct t S V V G O O V V GG   w w w  w w w wª º    ¬ ¼ w § · w  ¨ ¸ w© ¹ (4.19) Note that the term 2 2 2 222 2 ʌ Ht H kS t t V V GG § ·¨ ¸© ¹ is nonlinear, except when 2 2 t C S w* w does not change sign for all tS . Since it represents the degree of mishedging of the portfolio, it is not surprising to observe that * is involved in the transaction cost term. We may rewrite Equation (4.19) in the form which resembles the Merton equation: > @ 2 2 2 2 ˆ , , 2 1 0. t t t t t t t t t SC C CrS rC E C t J S C t S t S S CE J S S V O O w w w ª º    ¬ ¼w w w w  w (4.20) where > @ 2 21 e 1 J JE J VP   and the implied volatility is given by 22 1 2 22 2 2 22ˆ .ʌ H H H Ht k t signt VV V V G V GG  § ·    *¨ ¸© ¹ (4.21) If 2Vˆ , Equation (4.20) becomes mathematically ill-posed. This occurs when 0*  and 0tG o . However, it is known that * is always positive for the simple European call and put options in the absence of transaction costs. If we postulate the same sign behaviour for * in the presence of transaction costs, Equation (4.20) becomes linear under such an assumption so that the Merton formula becomes applicable except that the modified volatility Vˆ should be used as the volatility parameter. Moreover, if 0* ! from Equation (4.20) we obtain 1 2 0 e , e , ! nT t r T t t t n T t C t S S d K d n O O I I c f   c  ª º ¬ ¼¦ where 2 1 2 1 ln 2 , , t n n n n S r T t T t Kd d d T t T t V VV § ·    ¨ ¸© ¹   2 2 2 22 ˆe , , J J J n nE J T t VP VO O O V Vc   2 2 2 2ln 1 e 1 , J J J J n n n E J r r E J r T t T t VP VP O O  § ·¨ ¸§ · © ¹¨ ¸      ¨ ¸ © ¹ Acta Wasaensia 73 F. Shokrollahi DOI: 10.4236/jmf.2018.84040 639 Journal of Mathematical Finance 22 1 2 22 2 2 22ˆ ,ʌ H H H Ht k t signt VV V V G V GG  § ·    *¨ ¸© ¹ sign * is the signum function of 2 2 t C S w w , tG is a small and fixed time-step, k is the transaction costs and .I is the cumulative normal distribution. 74 Acta Wasaensia