1: \documentclass[11pt,a4paper,reqno,intlimits,sumlimits]{amsart}
2:
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{chicago}
6: \usepackage[mathscr]{euscript}
7: \usepackage{enumerate}
8: \usepackage{xspace}
9: \usepackage{color}
10: \usepackage{epsfig,rotating}
11:
12: \DeclareMathAlphabet{\mathpzc}{OT1}{pzc}{m}{it}
13:
14: \usepackage[bookmarksopen,pdfstartview=FitH]{hyperref}
15: \hypersetup{colorlinks,%
16: citecolor=black,%
17: filecolor=black,%
18: linkcolor=black,%
19: urlcolor=black}
20:
21:
22: \begin{document}
23:
24: \theoremstyle{plain}
25: \newtheorem{theorem}{Theorem}[section]
26: \newtheorem{lemma}[theorem]{Lemma}
27: \newtheorem{proposition}[theorem]{Proposition}
28: \newtheorem{corollary}[theorem]{Corollary}
29:
30: \theoremstyle{definition}
31: \newtheorem{remark}[theorem]{Remark}
32: \newtheorem{example}[theorem]{Example}
33: \newtheorem{assumption}[theorem]{Assumption}
34:
35: \newcommand{\Law}{\ensuremath{\mathop{\mathrm{Law}}}}
36: \newcommand{\loc}{{\mathrm{loc}}}
37: \newcommand{\Log}{\ensuremath{\mathop{\mathcal{L}\mathrm{og}}}}
38:
39: \let\SETMINUS\setminus
40: \renewcommand{\setminus}{\backslash}
41:
42: \def\stackrelboth#1#2#3{\mathrel{\mathop{#2}\limits^{#1}_{#3}}}
43:
44: \renewcommand{\theequation}{\thesection.\arabic{equation}}
45: \numberwithin{equation}{section}
46:
47: \newcommand{\prozess}[1][L]{{\ensuremath{#1=(#1_t)_{0\le t\le T}}}\xspace}
48: \newcommand{\prazess}[1][L]{{\ensuremath{#1=(#1_t)_{0\le t\le T^*}}}\xspace}
49: \newcommand{\scal}[2]{\ensuremath{\langle #1, #2 \rangle}}
50:
51: %% ================================================================ %%
52: \def\F{\ensuremath{\mathcal{F}}}
53: \def\R{\ensuremath{\mathbb{R}}}
54: \def\C{\ensuremath{\mathbb{C}}}
55: \def\bF{\mathbf{F}}
56: \def\V{\mathbb{V}}
57:
58: \def\Rmz{\R\setminus\{0\}}
59: \def\Rdmz{\R^d\setminus\{0\}}
60: \def\Rnmz{\R^n\setminus\{0\}}
61: \def\Rp{\mathbb{R}_+}
62:
63: \def\la{\ensuremath{L^1(\R)}}
64: \def\lad{\ensuremath{L^1(\R^d)}}
65: \def\lat{\ensuremath{L^2(\R)}}
66: \def\lap{\ensuremath{L^\infty(\R)}}
67: \def\labc{\ensuremath{L^1_{\text{bc}}(\R)}}
68:
69: \def\lev{L\'{e}vy\xspace}
70: \def\lk{L\'{e}vy--Khintchine\xspace}
71: \def\smmg{semimartingale\xspace}
72: \def\mg{martingale\xspace}
73: \def\tih{time-inhomogeneous\xspace}
74: \def\chartri{\ensuremath{(b,\sigma,\nu)}}
75: \def\num{num\'{e}raire\xspace}
76:
77: \def\eqlaw{\ensuremath{\stackrel{\mathrrefersm{d}}{=}}}
78:
79: \def\dsdx{\ensuremath{(\ud s, \ud x)}}
80: \def\dtdx{\ensuremath{(\ud t, \ud x)}}
81:
82: \def\intrr{\ensuremath{\int_{\R}}}
83:
84: \def\ud{\ensuremath{\mathrm{d}}}
85: \def\e{\mathrm{e}}
86: \def\dt{\ud t}
87: \def\ds{\ud s}
88: \def\dx{\ud x}
89: \def\dy{\ud y}
90: \def\dz{\ud z}
91: \def\du{\ud u}
92: \def\icc{\mathpzc{i}}
93: \def\ecc{\mathbf{e}_\mathpzc{i}}
94:
95: \def\EM{\ensuremath{(\mathbb{EM})}\xspace}
96: \def\ES{\ensuremath{(\mathbb{ES})}\xspace}
97: \def\AC{\ensuremath{(\mathbb{AC})}\xspace}
98:
99: \def\ott{{0\leq t\leq T}}
100:
101: \def\bg{\ensuremath{\bar g}}
102: \def\logs{\mathpzc s}
103: %% ================================================================ %%
104:
105:
106: \title[Analysis of valuation formulas]
107: {Analysis of fourier transform valuation formulas and applications}
108:
109: \author[E. Eberlein]{Ernst Eberlein}
110: \author[K. Glau]{Kathrin Glau}
111: \author[A. Papapantoleon]{Antonis Papapantoleon}
112:
113: \address{Department of Mathematical Stochastics, University of Freiburg,
114: Eckerstr. 1, 79104 Freiburg, Germany}
115: \email{eberlein@stochastik.uni-freiburg.de}
116:
117: \address{Department of Mathematical Stochastics, University of Freiburg,
118: Eckerstr. 1, 79104 Freiburg, Germany}
119: \email{glau@stochastik.uni-freiburg.de}
120:
121: \address{Institute of Mathematics, TU Berlin, Stra\ss e des 17. Juni 136,
122: 10623 Berlin, Germany \& Quantitative Products
123: Laboratory, Deutsche Bank AG, Alexanderstr. 5, 10178 Berlin,
124: Germany}
125: \email{papapan@math.tu-berlin.de}
126:
127: \keywords{option valuation; Fourier transform; semimartingales; \lev processes;
128: stochastic volatility models; options on several assets}
129:
130: \subjclass[2000]{91B28; 42B10; 60G48}
131:
132: \thanks{We would like to thank Friedrich Hubalek and Gabriel Maresch for
133: valuable discussions. K.~G. would like to thank the DFG for financial
134: support through project EB66/11-1, and the Austrian Science Fund (FWF)
135: for an invitation under grant P18022. A.~P. gratefully acknowledges the
136: financial support from the Austrian Science Fund (FWF grant Y328,
137: START Prize). \\\indent
138: We would like to thank the two anonymous referees for their careful
139: reading of the manuscript and their valuable suggestions that have
140: improved the paper.}
141:
142: \date{}
143: \maketitle
144:
145: \frenchspacing
146: \pagestyle{myheadings}
147:
148:
149:
150: \begin{abstract}
151: The aim of this article is to provide a systematic analysis of the conditions
152: such that Fourier transform valuation formulas are valid in a general framework;
153: i.e. when the option has an arbitrary payoff function and depends on the path
154: of the asset price process. An interplay between the conditions on the payoff
155: function and the process arises naturally. We also extend these results to the
156: multi-dimensional case, and discuss the calculation of Greeks by Fourier
157: transform methods. As an application, we price options on the minimum of two
158: assets in \lev and stochastic volatility models.
159: \end{abstract}
160:
161:
162:
163: \section{Introduction}
164:
165: Since the seminal work of \citeN{CarrMadan99} and \citeN{Raible00} on the
166: valuation of options with Fourier transform methods, there have been several
167: articles dealing with extensions and analysis of these
168: valuation formulas. This literature focuses on the extension of the method to
169: other situations, e.g. the pricing of exotic or multi-asset derivatives, or
170: on the analysis of the discretization error of the fast Fourier transform.
171:
172: The article of \citeN{BorovkovNovikov02} deals with the application of Fourier
173: transform valuation formulas for the pricing of some exotic options, while
174: \shortciteN{HubalekKallsenKrawczyk06} use similar techniques for hedging
175: purposes. Lee \citeyear{Lee04} provides an analysis of the discretization error in the
176: fast Fourier transform, while \citeN{Lord08} extends the method to the pricing
177: of options with early exercise features. Recently, \citeN{HubalekKallsen03},
178: Biagini et al.
179: \citeyear{BiaginiBregmanMeyerBrandis08} and \citeN{HurdZhou09} extend the
180: method to accommodate options on several assets, considering basket options,
181: spread options and catastrophe insurance derivatives.
182: \shortciteN{DufresneGarridoMorales09} also consider the valuation of payoffs
183: arising in insurance mathematics by Fourier methods. In addition, the books
184: of \citeN{ContTankov03}, Boyarchenko and Levendorski\v{\i}
185: \citeyear{BoyarchenkoLevendorskii02book} and
186: \citeN{Schoutens03} are also discussing Fourier transform methods for option
187: pricing. Let us point out that all these results are intimately related to
188: Parseval's formula, cf. \citeN[VI.2.2]{Katznelson04}.
189:
190: The aim of our article is to provide a systematic analysis of the conditions
191: required for the \textit{existence} of Fourier transform valuation formulas in
192: a general framework: i.e. when the underlying variable can depend on the path
193: of the price process and the payoff function can be discontinuous. Such an
194: analysis seems to be missing in the literature.
195:
196: In their work, \citeN{CarrMadan99}, \citeN{Raible00} and most others are
197: usually imposing a \emph{continuity} assumption, either on the payoff function
198: or on the random variable (i.e. existence of a Lebesgue density). However, when
199: considering e.g. a one-touch option on a \lev-driven asset, both assumptions
200: fail: the payoff function is clearly discontinuous, while a priori not much is
201: known about the existence of a density for the distribution of the supremum of
202: a \lev process. Analogous situations can also arise in higher dimensions.
203:
204: The key idea in Fourier transform methods for option pricing lies in the
205: separation of the \textit{underlying process} and the \textit{payoff function}.
206: We derive conditions on the moment generating function of the underlying random
207: variable and the Fourier transform of the payoff function such that
208: Fourier based valuation formulas hold true in one and several dimensions. An
209: interesting interplay between the continuity conditions imposed on the payoff
210: function and the random variable arises naturally. We also derive a result
211: that allows to easily verify the conditions on the payoff function (cf. Lemma
212: \ref{sobo}).
213:
214: The results of our analysis can be briefly summarized as follows: for general
215: continuous payoff functions or for variables, whose distribution has a Lebesgue
216: density, the valuation formulas using Fourier transforms are valid as Lebesgue
217: integrals, in one and several dimensions. When the payoff function is discontinuous
218: and the random variable might not possess a Lebesgue density then, in dimension
219: one, we get pointwise convergence of the valuation formulas under additional
220: assumptions, that are typically satisfied. In several dimensions pointwise
221: convergence fails, but we can deduce the valuation function as an $L^2$-limit.
222:
223: In addition, the structure of the valuation formulas allows us to derive easily
224: formulas for the sensitivities of the option price with respect to the various
225: parameters; otherwise, Malliavin calculus techniques or cubature formulas have
226: to be employed, cf. e.g. \shortciteN{FournieLasryLebuchouxLionsTouzi99},
227: \citeN{Teichmann06} and \citeN{KohatsuHigaYasuda08}. We discuss results
228: regarding the sensitivities with respect to the initial value, i.e. the delta
229: and the gamma. It turns out that the trade-off between continuity conditions on
230: the payoff function and the random variable established for the valuation
231: formulas, becomes now a trade-off between integrability and smoothness
232: conditions for the calculation of the sensitivities.
233:
234: The valuation formulas allow to compute prices of European options very fast,
235: hence they allow the efficient calibration of the model to market data for a
236: large variety of driving processes, such as \lev processes and affine stochastic
237: volatility models. Indeed, for \lev and affine processes the moment generating
238: function is usually known explicitly, hence these models are tailor-made for
239: Fourier transform pricing formulas.
240:
241: We also mention here that the Fourier transform based approach can be applied
242: for the efficient computation of prices in other frameworks as well.
243: An important area is the
244: valuation of interest rate derivatives in \lev driven models. \lev term
245: structure models were developed in a series of papers in the last ten years;
246: this development is surveyed in \citeN{EberleinKluge06}.
247: For the Fourier based formulas we mention the two papers by
248: Eberlein and Kluge \citeyear{EberleinKluge04,EberleinKluge05}, where caps,
249: floors, and swaptions as well as interest rate digital and range digital options
250: are discussed; furthermore Eberlein and Koval \citeyear{EberleinKoval06},
251: where cross currency derivatives are considered and
252: Eberlein, Kluge, and Sch\"onbucher \citeyear{EberleinKlugeSchoenbucher06}, where
253: pricing formulas for credit default swaptions are derived. Moreover, in the
254: framework of the `affine LIBOR' model (cf.
255: \shortciteNP{KellerResselPapapantoleonTeichmann09}) caps and swaptions can be
256: easily priced by Fourier based methods.
257:
258: This paper is organized as follows: in Section \ref{gval} we present valuation
259: formulas in the single asset case, and in Section \ref{Rdval} we deal with the
260: valuation of options on several assets. In Section \ref{greeks} we discuss
261: sensitivities. In Section \ref{ch3:payoffs} we review examples of commonly
262: used payoff functions, in dimension one and in multiple dimensions. In Section
263: \ref{LA} we review L\'evy and affine processes. Finally, in Section
264: \ref{ch3:sc5} we provide numerical examples for the valuation of options on
265: several assets in \lev and affine stochastic volatility models.
266:
267:
268:
269: \section{Option valuation: single asset}
270: \label{gval}
271:
272: \subsection*{1.}
273: Let $\mathscr B=(\Omega,\F,\bF,P)$ be a stochastic basis in the sense of Jacod and
274: Shiryaev \citeyear[I.1.3]{JacodShiryaev03}, where $\F=\F_T$ and $\bF
275: =(\F_t)_{0\le t\le T}$. We model the price process of a financial
276: asset, e.g. a stock or an FX rate, as an \textit{exponential
277: semimartingale} $S=(S_t)_{0\le t\le T}$, i.e. a stochastic process
278: with representation
279: \begin{equation}\label{ch3:eq1}
280: S_t = S_0 \e^{H_t},\qquad 0\le t\le T
281: \end{equation}
282: (shortly: $S=S_0\e^H$), where $H=(H_t)_{0\le t\le T}$ is a
283: semimartingale with $H_0=0$.
284:
285: Every semimartingale $H=(H_t)_{0\le t\le T}$
286: admits a \emph{canonical representation}
287: \begin{equation}\label{ch3:eq2}
288: H = B + H^c+h(x)*(\mu-\nu) + (x-h(x))*\mu,
289: \end{equation}
290: where $h=h(x)$ is a \emph{truncation function}, $B=(B_t)_{0\le t\le T}$
291: is a predictable process of bounded variation, $H^c=(H_t^c)_{0\le t\le T}$
292: is the continuous martingale part of $H$ with predictable quadratic
293: characteristic $\langle H^c\rangle=C$, and $\nu$ is the predictable
294: compensator of the random measure of jumps $\mu$ of $H$. Here $W*\mu$
295: denotes the integral process of $W$ with respect to $\mu$, and $W*(\mu-\nu)$
296: denotes the stochastic integral of $W$ with respect to the compensated random
297: measure $\mu-\nu$; cf. \citeN[Chapter II]{JacodShiryaev03}.
298:
299: Let $\mathcal M(P)$, resp. $\mathcal M_\loc(P)$, denote the class of
300: all martingales, resp. local martingales, on the given stochastic
301: basis $\mathscr B$.
302:
303: Subject to the assumption that the process $1_{\{x>1\}}\e^x*\nu$ has
304: bounded variation, we can deduce the martingale condition
305: \begin{equation}\label{ch3:eq3}
306: S = S_0\e^H\in\mathcal M_\loc(P)
307: \Leftrightarrow
308: B + \frac{C}{2} + (\e^x-1-h(x))*\nu = 0;
309: \end{equation}
310: cf. \shortciteN{EberleinPapapantoleonShiryaev06} for details. The martingale
311: condition can also be expressed in terms of the cumulant process $K$
312: associated to $(B,C,\nu)$, i.e. $K(1)=0$; for the cumulant process see
313: \citeN{JacodShiryaev03}.
314:
315: Throughout this work, we assume that $P$ is an (equivalent) martingale
316: measure for the asset $S$ and the martingale condition is in force; moreover,
317: for simplicity we assume that the interest rate and dividend yield are zero.
318: By no-arbitrage theory the price of an option on $S$ is calculated as its
319: discounted expected payoff.
320:
321:
322: \subsection*{2.}
323: Let $Y=(Y_t)_{0\le t\le T}$ be a stochastic process on the given basis. We
324: denote by $\overline{Y}=(\overline{Y}_t)_{0\le t\le T}$ and
325: $\underline{Y}=(\underline{Y}_t)_{0\le t\le T}$ the supremum and the infimum
326: processes of $Y$ respectively, i.e.
327: \begin{displaymath}
328: \overline{Y}_t = \sup_{0\leq u\leq t}Y_u
329: \quad \mathrm{ and } \quad
330: \underline{Y}_t = \inf_{0\leq u\leq t}Y_u.
331: \end{displaymath}
332:
333: Notice that since the exponential function is monotonically increasing,
334: the supremum processes of $S$ and $H$ are related via
335: \begin{align}\label{sup-SH}
336: \overline{S}_T & = \sup_{0\leq t\leq T}\left(S_0\e^{H_t}\right)
337: = S_0\e^{\sup_{0\leq t\leq T} H_t}
338: = S_0\e^{\overline{H}_T}.
339: \end{align}
340: Similarly, the infimum processes of $S$ and $H$ are related via
341: \begin{align*}
342: \underline{S}_T = S_0\e^{\underline{H}_T}.
343: \end{align*}
344:
345:
346: \subsection*{3.}
347: The aim of this work is to tackle the problem of efficient valuation for
348: plain vanilla options, such as \emph{European call} and \emph{put} options,
349: as well as for exotic path-dependent options, such as \emph{lookback}
350: and \emph{one-touch} options, in a unified framework. Therefore, we will
351: analyze and prove valuation formulas for options on an asset $S=S_0\e^{H}$
352: with a payoff at maturity $T$ that may depend on the whole path of $S$ up
353: to time $T$. These results, together with analyticity conditions on the
354: Wiener--Hopf factors, will be used in the companion paper
355: \cite{EberleinGlauPapapantoleon09} for the pricing of one-touch and lookback
356: options in \lev models.
357:
358:
359: The following example of a fixed strike lookback option will serve as
360: a guideline for our methodology; note that using \eqref{sup-SH} it
361: can be re-written as
362: \begin{align}\label{stdlookb}
363: \big(\overline{S}_T - K\big)^+ = \big(S_0\e^{\overline{H}_T} - K\big)^+\,.
364: \end{align}
365:
366: In order to incorporate both plain vanilla options and exotic options in a
367: single framework we separate the \emph{payoff function} from the
368: \emph{underlying process}, where:
369: \begin{enumerate}[(a)]
370: \item the \emph{underlying process} can be the log-asset price
371: process or the supremum/infimum of the log-asset price process
372: or an average of the log-asset price process. This process will always be
373: denoted by $X$ (i.e. $X=H$ or $X=\overline H$ or $X=\underline H$, etc.);
374: \item the \emph{payoff function} is an arbitrary function
375: $f:\R\rightarrow\Rp$, for example $f(x)=(\e^x-K)^+$ or
376: $f(x)=1_{\{\e^x>B\}}$, for $K,B\in\Rp$.
377: \end{enumerate}
378:
379: Clearly, we regard options as dependent on the \textit{underlying process} $X$,
380: i.e. on (some functional of) the logarithm of the asset price process $S$. The
381: main advantage is that the characteristic function of $X$ is easier to handle
382: than that of (some functional of) $S$; for example, for a \lev process $H=X$ it
383: is already known in advance.
384:
385: Moreover, we consider exactly those options where we can incorporate the
386: path-dependence of the option payoff into the underlying process $X$. European
387: vanilla options are a trivial example, as there is no path-dependence; a
388: non-trivial, example are options on the supremum, see again
389: \eqref{sup-SH} and \eqref{stdlookb}. Other examples are the geometric Asian
390: option and forward-start options.
391:
392: In addition, we will assume that the initial value of the underlying process $X$
393: is zero; this is the case in all natural examples in mathematical finance. The
394: initial value $S_0$ of the asset price process $S$ plays a particular role,
395: because it is convenient to consider the option price as a function of it, or
396: more specifically as a function of $\logs=-\log S_0$.
397:
398: Hence, we express a general payoff as
399: \begin{align}
400: \Phi\big(S_0\e^{H_t},\,0\le t\le T\big) = f(X_T-\logs)\,,
401: \end{align}
402: where $f$ is a payoff function and $X$ is the underlying process, i.e.
403: an adapted process, possibly depending on the full history of $H$,
404: with
405: $$X_t:= \Psi(H_s,\,0\le s\le t)\quad \text{ for } \,t\in[0,T],$$
406: and $\Psi$ a measurable functional. Therefore, the time-$0$ price of the option
407: is provided by the (discounted) expected payoff, i.e.
408: \begin{align}\label{genericprice}
409: \mathbb{V}_f(X;\logs)
410: = E\big[\Phi\big(S_t,\,0\le t\le T\big)\big]
411: = E\big[ f(X_T-\logs)\big]\,.
412: \end{align}
413:
414: Note that we consider `European style' options, in the sense that the holder
415: or writer do not have the right to exercise or terminate the option before
416: maturity.
417:
418: \begin{remark}
419: In case the interest rate $r$ and the dividend yield $\delta$ are non-zero,
420: then the martingale condition \eqref{ch3:eq3} reads
421: \begin{equation*}
422: (\delta - r)t + B_t + \frac{C_t}{2} + (\e^x-1-h(x))*\nu_t = 0
423: \end{equation*}
424: for all $t$, and the option price is given by
425: $\mathbb{V}_f(X;\logs)
426: = \e^{-rT}E\big[ f(X_T-\logs)\big]\,.$
427: \end{remark}
428:
429:
430: \subsection*{4.}
431: The first result focuses on options with continuous payoff functions, such as
432: European plain vanilla options, but also lookback options.
433:
434: Let $P_{X_T}$ denote the law, $M_{X_T}$ the moment generating function and
435: $\varphi_{X_T}$ the (extended) characteristic function of the random variable
436: $X_T$; that is
437: \begin{align*}
438: M_{X_T}(u)
439: = E\big[\e^{uX_T}\big]
440: = \varphi_{X_T}(-iu),
441: \end{align*}
442: for suitable $u\in\C$. For any payoff function $f$ let $g$ denote the \emph{dampened}
443: payoff function, defined via
444: \begin{align}\label{g-defin}
445: g(x)=\e^{-Rx}f(x)
446: \end{align}
447: for some $R\in\R$. Let $\widehat{g}$ denote the (extended) Fourier transform
448: of a function $g$, and $L^1_{\text{bc}}(\R)$ the space of bounded, continuous
449: functions in $L^1(\R)$.
450:
451: In order to derive a valuation formula for an option with an
452: arbitrary \emph{continuous} payoff function $f$, we will impose the
453: following conditions.
454: \begin{description}
455: \item[(C1)] Assume that $g\in L^1_{\text{bc}}(\R)$.
456: \item[(C2)] Assume that $M_{X_T}(R)$ exists.
457: \item[(C3)] Assume that $\widehat{g}\in L^1(\R)$.
458: \end{description}
459:
460: \begin{theorem}\label{valuation}
461: If the asset price process is modeled as an exponential
462: semimartingale process according to \eqref{ch3:eq1}--\eqref{ch3:eq3}
463: and conditions (C1)--(C3) are in force, then the time-0 price function
464: is given by
465: \begin{align}\label{value}
466: \mathbb V_f(X;\logs) =
467: \frac{\e^{-R\logs}}{2\pi}
468: \int_\R \e^{-iu\logs}\varphi_{X_T}(u-iR)\widehat{f}(iR-u)\ud u.
469: \end{align}
470: \end{theorem}
471: \begin{proof}
472: Using \eqref{genericprice} and \eqref{g-defin} we have
473: \begin{align}\label{ch3:eq4}
474: \V_f(X;\logs)
475: = \int_\Omega f(X_T-\logs)\ud P
476: = \e^{-R\logs}\int_\R \e^{Rx}g(x-\logs)P_{X_T}(\dx).
477: \end{align}
478: By assumption (C1), $g\in\la$ and the Fourier transform of $g$,
479: \begin{align*}
480: \widehat{g}(u) = \int_\R \e^{iux} g(x)\dx,
481: \end{align*}
482: is well defined for every $u\in\R$ and is also continuous and
483: bounded. Additionally, using assumption (C3) we immediately
484: have that $\widehat{g}\in\labc$. Therefore, using the Inversion
485: Theorem (cf. \citeNP[Theorem 3.4.4]{Deitmar04}), $\widehat{g}$
486: can be inverted and $g$ can be represented, for \emph{all} $x\in\R$, as
487: \begin{align}\label{ch3:eq7}
488: g(x) = \frac1{2\pi} \int_\R \e^{-ixu}\widehat{g}(u)\ud u.
489: \end{align}
490:
491: Now, returning to the valuation problem \eqref{ch3:eq4} we get that
492: \begin{align}\label{ch3:eq8}
493: \V_f(X;\logs)
494: &= \e^{-R\logs}\int_\R \e^{Rx}
495: \Bigg(\frac1{2\pi}\int_\R \e^{-i(x-\logs)u}\widehat{g}(u)\ud u\Bigg)P_{X_T}(\dx)\nonumber\\
496: &= \frac{\e^{-R\logs}}{2\pi}\int_\R \e^{iu\logs}
497: \Bigg( \int_\R \e^{i(-u-iR)x}P_{X_T}(\dx)\Bigg)\widehat{g}(u)\ud u\nonumber\\
498: &= \frac{\e^{-R\logs}}{2\pi}\int_\R \e^{iu\logs}\varphi_{X_T}(-u-iR)\widehat{f}(u+iR)\ud u,
499: \end{align}
500: where for the second equality we have applied Fubini's theorem;
501: moreover, for the last equality we have
502: \begin{align*}%\label{ch3:FgFf}
503: \widehat{g}(u) = \int_\R \e^{iux}\e^{-Rx}f(x)\dx
504: = \widehat{f}(u+iR).
505: \end{align*}
506:
507: Finally, the application of Fubini's theorem is justified since
508: \begin{align*}
509: \int_\R\int_\R \e^{Rx}|\e^{-iu(x-\logs)}||\widehat{g}(u)|\ud u P_{X_T}(\dx)
510: &\le \int_\R\e^{Rx}\bigg(\int_\R |\widehat{g}(u)|\ud u \bigg)P_{X_T}(\dx)\\
511: &\le K M_{X_T}(R) < \infty,
512: \end{align*}
513: where we have used again that $\widehat{g}\in L^1(\R)$, and the finiteness
514: of $M_{X_T}(R)$ is given by Assumption (C2).
515: \end{proof}
516:
517: \begin{remark}
518: We could also replace assumptions (C1) and (C3) with the following conditions
519: \begin{center}
520: (C1$'$): $g\in L^1(\R)\quad$ and $\quad$
521: (C3$'$): $\widehat{\e^{Rx}P_{X_T}}\in L^1(\R)$.
522: \end{center}
523: Condition (C3$'$) yields that $\e^{Rx}P_{X_T}$ possesses a continuous
524: bounded Lebesgue density, say $\rho$; cf. \citeN[Theorem 8.39]{Breiman68}.
525: Then, we can identify $\rho$, instead of $g$, with the inverse of its Fourier
526: transform and the proof goes through with the obvious modifications. This
527: statement is almost identical to Theorem 3.2 in \citeN{Raible00}.
528: \end{remark}
529:
530: \begin{remark}[Numerical evaluation]
531: The option price represented as an integral of the form \eqref{value} can be
532: evaluated numerically very fast. The following simple observation can speed
533: up the computation of this expression even further: notice that for a fixed
534: maturity $T$, the characteristic function -- which is the computationally
535: expensive part -- should only be evaluated \emph{once} for \emph{all}
536: different strikes or initial values. The gain in computational time will be
537: significant when considering models where the characteristic function is not
538: known in closed form; e.g. in affine models where one might need to solve a
539: Riccati equation to obtain the characteristic function. This observation has
540: been termed `caching' by some authors (cf. \citeNP{Kilin07})
541: \end{remark}
542:
543:
544: \subsection*{5.}
545: Apart from (C3), the prerequisites of Theorem \ref{valuation} are quite easy
546: to check in specific cases. In general, it is also an interesting question
547: to know when the Fourier transform of an integrable function is integrable.
548: The problem is well understood for smooth ($C^2$ or $C^\infty$) functions, see
549: e.g. \citeN{Deitmar04}, but the functions we are dealing with are typically
550: not smooth. Hence, we will provide below an easy-to-check condition for a
551: non-smooth function to have an integrable Fourier transform.
552:
553: Let us consider the Sobolev space $H^1(\R)$, with
554: \begin{align*}
555: H^1(\R) = \Big\{g\in L^2(\R) \;\Big|\; \partial g \text{ exists and } \partial g\in L^2(\R)\Big\},
556: \end{align*}
557: where $\partial g$ denotes the \emph{weak} derivative of a function
558: $g$; see e.g. \citeANP{Sauvigny06} \citeyear{Sauvigny06}. Let $g\in H^1(\R)$, then from
559: Proposition 5.2.1 in \citeN{Zimmer90} we get that
560: \begin{align}\label{fou-der}
561: \widehat{\partial g}(u) = -iu\widehat{g}(u)
562: \end{align}
563: and $\widehat{g},\widehat{\partial g}\in L^2(\R)$.
564:
565: \begin{lemma}\label{sobo}
566: Let $g\in H^1(\R)$, then $\widehat{g}\in L^1(\R)$.
567: \end{lemma}
568: \begin{proof}
569: Using the above results, we have that
570: \begin{align}\label{fou-fdf}
571: \infty
572: > \int_\R \Big(\big|\widehat{g}(u)\big|^2 + \big|\widehat{\partial g}(u)\big|^2\Big)\ud u
573: = \int_\R \big|\widehat{g}(u)\big|^2\big(1+|u|^2\big)\ud u.
574: \end{align}
575: Now, by the H\"older inequality, using $(1+|u|)^2\leq3(1+|u|^2)$
576: and \eqref{fou-fdf}, we get that
577: \begin{align*}
578: \int_\R \big|\widehat{g}(u)\big|\ud u
579: &= \int_\R \big|\widehat{g}(u)\big|\frac{1+|u|}{1+|u|}\ud u\\
580: &\le \bigg(\int_\R \big|\widehat{g}(u)\big|^2(1+|u|)^2\ud u\bigg)^{\frac12}
581: \bigg(\int_\R \frac{1}{(1+|u|)^2}\ud u\bigg)^{\frac12}
582: < \infty
583: \end{align*}
584: and the result is proved.
585: \end{proof}
586:
587: \begin{remark}
588: A similar statement can be proved for functions in the Sobolev-Slobodeckij
589: space $H^s(\R)$, for $s>\frac12$.
590: \end{remark}
591:
592:
593: \subsection*{6.}
594: Next, we deal with the valuation formula for options whose payoff function can
595: be \textit{discontinuous}, while at the same time the measure $P_{X_T}$ does
596: \textit{not} necessarily possess a Lebesgue density. Such a situation arises
597: typically when pricing one-touch options in purely discontinuous \lev models.
598: Hence, we need to impose different conditions, and we derive the valuation formula
599: as a pointwise limit by generalizing the proof of Theorem 3.2 in \citeN{Raible00}.
600: A similar result (Theorem 1 in \shortciteNP{DufresneGarridoMorales09}) has been
601: pointed out to us by one of the referees.
602:
603: In this and the following sections we will make use of the following
604: notation; we define the function $\bg$ and the measure $\varrho$ as
605: follows
606: \begin{align*}
607: \bg(x):=g(-x)\quad \text{ and } \quad \varrho(\dx):=\e^{Rx}P_{X_T}(\dx).
608: \end{align*}
609: Moreover $\varrho(\R)=\int\varrho(\dx)$, while $\bg*\varrho$ denotes the
610: convolution of the function $\bg$ with the measure $\varrho$. In this case
611: we will use the following assumptions.
612:
613: \begin{description}
614: \item[(D1)] Assume that $g\in L^1(\R)$.
615: \item[(D2)] Assume that $M_{X_T}(R)$ exists
616: ($\Longleftrightarrow \varrho(\R)<\infty$).
617: \end{description}
618:
619: \begin{theorem}\label{valuation-dc}
620: Let the asset price process be modeled as an exponential
621: semimartingale process according to \eqref{ch3:eq1}--\eqref{ch3:eq3}
622: and conditions (D1)--(D2) be in force. The time-0 price function is
623: given by
624: \begin{align}\label{value-dc}
625: \mathbb V_f(X;\logs) =
626: \lim_{A\rightarrow\infty}\frac{\e^{-R\logs}}{2\pi}
627: \int_{-A}^A \e^{-iu\logs}\varphi_{X_T}(u-iR)\widehat{f}(iR-u)\ud u,
628: \end{align}
629: at the point $\logs\in\R$, if $\mathbb V_f(X;\cdot)$ is of bounded
630: variation in a neighborhood of $\logs$, and $\mathbb V_f(X;\cdot)$
631: is continuous at $\logs$.
632: \end{theorem}
633:
634: \begin{remark}
635: In Section \ref{ch3:payoffs} we will relate the conditions on the
636: valuation function $\mathbb V_f$ to properties of the measure $P_{X_T}$
637: for specific (dampened) payoff functions $g$. These properties are
638: easily checkable -- and typically satisfied -- in many models.
639: \end{remark}
640:
641: \begin{proof}
642: Starting from \eqref{ch3:eq4}, we can represent the option price function
643: as a convolution of $\bg$ and $\varrho$ as follows
644: \begin{align}\label{value-pw}
645: \V_f(X;\logs) &= \e^{-R\logs}\int_\R \e^{Rx}g(x-\logs)P_{X_T}(\dx)
646: = \e^{-R\logs}\int_\R \bg(\logs-x)\varrho(\dx) \nonumber\\
647: &= \e^{-R\logs}\bg*\varrho(\logs).
648: \end{align}
649: Using that $g\in L^1(\R)$, hence also $\bg\in L^1(\R)$, and $\varrho(\R)<\infty$
650: we get that $\bg*\varrho\in L^1(\R)$, since
651: \begin{align}\label{L1-conv}
652: \Vert\bg*\varrho\Vert_{L^1(\R)}
653: &\le \varrho(\R)\,\Vert\bg\Vert_{\la}<\infty;
654: \end{align}
655: compare with Young's inequality, cf. \citeN[IV.1.6]{Katznelson04}.
656: Therefore, the Fourier transform of the convolution is well defined
657: and we can deduce that, for all $u\in\R$,
658: \begin{align*}
659: \widehat{\bg*\varrho}(u)=\widehat{\bg}(u)\cdot \widehat{\varrho}(u);
660: \end{align*}
661: compare with Theorem 2.1.1 in \citeN{Bochner55}.
662:
663: By \eqref{L1-conv} we can apply the inversion theorem for the Fourier
664: transform, cf. Satz 4.2.1 in \citeN{Doetsch50}, and get
665: \begin{align}\label{ug-for}
666: \frac12 \big(\bg*\varrho(\logs+)+\bg*\varrho(\logs-)\big)
667: &= \frac{1}{2\pi} \lim_{A\rightarrow\infty}
668: \int_{-A}^A \e^{-iu\logs}\widehat{\varrho}(u)\widehat{\bg}(u)\du,
669: \end{align}
670: if there exists a neighborhood of $\logs$ where
671: $\logs\mapsto\bg*\varrho(\logs)$ is of bounded variation.
672:
673: We proceed as follows: first we show that the function
674: $\logs\mapsto\bg*\varrho(\logs)$ has bounded variation; then we show
675: that this map is also continuous, which yields that the left hand side
676: of \eqref{ug-for} equals $\bg*\varrho(\logs)$.
677:
678: For that purpose, we re-write \eqref{value-pw} as
679: \begin{align*}
680: \bg*\varrho(\logs) = \e^{R\logs}\,\mathbb V_f(X;\logs);
681: \end{align*}
682: then, $\bg*\varrho$ is of bounded variation on a compact interval
683: $[a,b]$ if and only if $\mathbb V_f(X;\cdot)\in BV([a,b])$;
684: this holds because the map $\logs\mapsto\e^{R\logs}$ is of bounded
685: variation on any bounded interval on $\R$, and the fact that the
686: space $BV([a,b])$ forms an algebra; cf. Satz 91.3 in \citeN{Heuser01}.
687: Moreover, $\logs$ is a continuity point of $\bg*\varrho$ if and only
688: if $\mathbb V_f(X;\cdot)$ is continuous at $\logs$.
689:
690: In addition, we have that
691: \begin{align}\label{dc-val-3}
692: \widehat{\bg}(u)
693: &= \int_\R \e^{-iux}\e^{-Rx}f(x)\dx
694: = \widehat{f}(iR-u)
695: \end{align}
696: and
697: \begin{align}\label{dc-val-4}
698: \widehat{\varrho}(u)
699: = \int_\R \e^{iux}\e^{Rx}P_{X_T}(\dx)
700: = \varphi_{X_T}(u-iR).
701: \end{align}
702: Hence, \eqref{ug-for} together with \eqref{dc-val-3}, \eqref{dc-val-4}
703: and the considerations regarding the continuity and bounded variation
704: properties of the value function yield the required result.
705: \end{proof}
706:
707:
708:
709: \section{Option valuation: multiple assets}
710: \label{Rdval}
711:
712: \subsection*{1.}
713: We would like to establish valuation formulas for options that depend
714: on several assets or on multiple functionals of one asset.
715: Typical examples of options on several assets are basket options and
716: options on the minimum or maximum of several assets, with payoff
717: \begin{displaymath}
718: (S_T^1\wedge\cdots\wedge S_T^d-K)^+,
719: \end{displaymath}
720: where $x\wedge y=\min\{x,y\}$. Typical examples of options on functionals
721: of a single asset are \emph{barrier} options, with payoff
722: \begin{displaymath}
723: (S_T-K)^+1_{\{\overline S_T>B\}},
724: \end{displaymath}
725: and \emph{slide-in} or \emph{corridor} options, with payoff
726: \begin{displaymath}
727: (S_T-K)^+ \sum_{i=1}^N1_{\{L<S_{T_i}<H\}},
728: \end{displaymath}
729: at maturity $T$, where $0=T_0<T_1<\dots<T_N=T$.
730:
731: In the previous section we proved that the valuation formulas for a single
732: underlying is still valid -- at least as a pointwise limit, under reasonable
733: additional assumptions -- even if the underlying distribution does not possess
734: a Lebesgue density and the payoff is discontinuous.
735:
736: In the present section we will generalize the valuation formulas to the case
737: of several underlyings. Once again, if either the joint distribution possesses
738: a Lebesgue density or the payoff function is continuous, the formula is valid
739: as a Lebesgue integral. In case both assumptions fail, we will encounter situations
740: that are apparently of harmless nature, but where the pointwise convergence
741: will fail. In this case we will establish the valuation formulas as an $L^2$-limit;
742: however, with respect to numerical evaluation, a stronger notion of convergence
743: would be preferable.
744:
745: Analogously to the single asset case we assume that the asset prices evolve
746: as exponential semimartingales. Let the driving process be an $\R^d$-valued
747: semimartingale $H=(H^1,\dots,H^d)^\top$ and $S=(S^1,\dots,S^d)^\top$ be the
748: vector of asset price processes; then each component $S^\icc$ of $S$ is
749: modeled as an exponential semimartingale, i.e.
750: \begin{align}\label{asset-Rd}
751: S^\icc_{t}=S^\icc_0\exp H^\icc_{t}, \qquad \ott, \;1\le \icc\le d,
752: \end{align}
753: where $H^\icc$ is an $\R$-valued
754: semimartingale with canonical representation
755: \begin{equation}\label{Hi-canon}
756: H^\icc = H^\icc_0 + B^\icc + H^{\icc,c} + h^\icc(x)*(\mu-\nu) + (x^\icc-h^\icc(x))*\mu,
757: \end{equation}
758: with $h^\icc(x)=\e_\icc^\top h(x)$. The martingale condition can be given
759: as in eq. (3.3) in \citeN{EberleinPapapantoleonShiryaev08}.
760:
761:
762: \subsection*{2.}
763: In the sequel, we will price options with payoff $f(X_T-\logs)$
764: at maturity $T$, where $X_T$ is an $\F_T$-measurable $\R^d$-valued
765: random variable, possibly dependent of the history of the $d$
766: driving processes, i.e.
767: \begin{align*}
768: X_T = \Psi\big(H_t,\,0\le t\le T\big),
769: \end{align*}
770: where $\Psi$ is an $\R^d$-valued measurable functional. Further $f$ is a
771: measurable function $f:\R^d\rightarrow\R_+$, and
772: $\logs=(\logs^1,\dots,\logs^d)\in\R^d$ with $\logs^\icc=-\log S_0^\icc$.
773:
774: Analogously to the single asset case, we use the dampened payoff function
775: \begin{align*}
776: g(x):= \e^{-\langle R,x\rangle}f(x) \quad\text{ for } x\in\R^d,
777: \end{align*}
778: and denote by $\varrho$ the measure defined by
779: \begin{align*}
780: \varrho(\dx):= \e^{\langle R,x\rangle} P_{X_T}(\dx),
781: \end{align*}
782: where $R\in\R^d$ serves as a dampening coefficient. Here $\langle\cdot,\cdot\rangle$
783: denotes the Euclidian scalar product in $\R^d$. The scalar product is extended to
784: $\C^d$ as follows: for $u,v\in\C^d$, set $\langle u,v\rangle = \sum_iu_iv_i$, i.e.
785: we do not use the Hermitian inner product. Moreover, $M_{X_T}$ and
786: $\varphi_{X_T}$ denote the moment generating, resp. characteristic, function
787: of the random vector $X_T$.
788:
789: To establish our results we will make use of the following assumptions.
790:
791: \begin{description}
792: \item[(A1)] Assume that $g\in L^1(\R^d)$.
793: \item[(A2)] Assume that $M_{X_T}(R)$ exists.
794: \item[(A3)] Assume that $\widehat\varrho\in L^1(\R^d)$.
795: \end{description}
796:
797: \begin{remark}
798: We can also replace Assumptions (A1) and (A3) with the following
799: assumption
800: \begin{description}
801: \item[(A1$'$)] Assume that $g\in L_{\text{bc}}^1(\R^d)$ and $\widehat{g}\in L^1(\R^d)$;
802: \end{description}
803: this shows again the interplay between the continuity properties of the payoff
804: function and the underlying distribution.
805: \end{remark}
806:
807: \begin{theorem}\label{valuation-Rd}
808: If the asset price processes are modeled as exponential semimartingale
809: processes according to \eqref{asset-Rd}--\eqref{Hi-canon} and conditions
810: (A1)--(A3) are in force, then the time-$0$ price function is given by
811: \begin{align}\label{value-Rd}
812: \mathbb V_{f}(X;\logs)
813: &= \frac{\e^{-\langle R,\logs\rangle}}{(2\pi)^d}
814: \int_{\R^d} \e^{-i\langle u,\logs\rangle} M_{X_T}(R+iu) \widehat{f}(iR-u) \ud u.
815: \end{align}
816: \end{theorem}
817: \begin{proof}
818: Similarly to the one-dimensional case we have that
819: \begin{align}\label{value-map}
820: \V_f(X;\logs) = \e^{-\langle R,\logs\rangle} \bar{g}*\varrho(\logs).
821: \end{align}
822: Since $g\in L^1(\R^d)$ and $\varrho(\R^d)<\infty$, we get that
823: $\bar{g}*\varrho\in L^1(\R^d)$; therefore
824: $\widehat{\bg*\varrho}(u)=\widehat{\bg}(u)\cdot\widehat{\varrho}(u)$ for all
825: $u\in\R^d$. By assumption we know that $\widehat{\varrho}\in L^1(\R^d)$; moreover
826: $\widehat{\bg}\in L^\infty(\R^d)$ since
827: $|\widehat{\bg}|\le\Vert g\Vert_{L^1(\R^d)}<\infty$.
828: These considerations yield that $\widehat{\bg*\varrho}\in L^1(\R^d)$, again by
829: using Young's inequality.
830:
831: Hence, applying the formula for the Fourier inversion, cf. Corollary
832: 1.21 in \citeN{SteinWeiss71}, we conclude that
833: \begin{align*}
834: \V_f(X;\logs)
835: &= \frac{\e^{-\langle R,\logs\rangle}}{(2\pi)^d}
836: \int_{\R^d}\e^{-i\langle u,\logs\rangle}\widehat{\bg}(u)\widehat{\varrho}(u)\du \nonumber\\
837: &= \frac{\e^{-\langle R,\logs\rangle}}{(2\pi)^d}
838: \int_{\R^d} \e^{-i\langle u,\logs\rangle} M_{X_T}(R+iu) \widehat{f}(iR-u) \ud u,
839: \end{align*}
840: for a.e. $\logs\in\R^d$.
841:
842: Moreover, if $\logs\mapsto\V_f(X;\logs)$ is continuous, then the
843: equality holds pointwise for \textit{all} $\logs\in\R^d$. The mapping
844: \eqref{value-map} is continuous if the mapping $\logs\mapsto\bg*\varrho(\logs)$
845: is continuous. Using Assumption (A3) we have that $\varrho$ possesses
846: a bounded continuous Lebesgue density $\rho\in L^1(\R^d)$; cf. Proposition
847: 2.5 (xii) in Sato \citeyear{Sato99}. Then $\bg*\varrho=\bg*\rho$ and
848: \begin{align}\label{Rd-cont}
849: \lim_{|x|\rightarrow0} \bg*\varrho(\logs+x)
850: &= \lim_{|x|\rightarrow0}\int \bg(\logs+x-z)\rho(z)\dz\nonumber\\
851: &= \int\lim_{|x|\rightarrow0} \bg(\logs+y)\rho(x-y)\dy
852: = \bg*\varrho(\logs)
853: \end{align}
854: yielding the continuity of the map. Note that we have used the continuity of
855: $\rho$; additionally, we can interchange integration and limit using the dominated
856: convergence theorem, with majorant $\bg(\cdot)\max_x\rho(x)$.
857: \end{proof}
858:
859: \begin{remark}
860: The proof using Assumption (A1$'$) follows analogously, with the obvious
861: modifications for \eqref{Rd-cont}.
862: \end{remark}
863:
864:
865: \subsection*{3.}
866: Next, we consider the valuation of options on several assets when the
867: payoff function is \emph{discontinuous} and the driving process does
868: \textit{not} necessarily possess a Lebesgue density.
869:
870: The main difference to the analogous situation in dimension one is that
871: the pointwise convergence of capped Fourier integrals -- as is the case
872: in Satz 4.2.1 in \citeN{Doetsch50} -- cannot be generalized to the
873: multidimensional case. M. Pinsky gives the following astonishing example
874: to illustrate this fact, see section 4.1 in \citeN{Pinsky93}; let $f$ be
875: the indicator function of the unit ball in $\R^3$, then
876: \begin{align}\label{pin-couex}
877: \frac{1}{(2\pi)^{3}}\int_{|x|\le A}\e^{-i\langle u,x\rangle}\widehat{f}(x)\dx\Big|_{u=0}
878: = 1 - \frac{2}{\pi} \sin(A) + o(1),
879: \end{align}
880: for $A\uparrow \infty$. Extrapolating the convergence results from the
881: one-dimensional case to $\R^3$, we would expect pointwise convergence
882: of the spherical sum to the indicator function, at least in the interior
883: of the ball; on the contrary, the right hand side of \eqref{pin-couex}
884: is even \textit{divergent}.
885:
886: As a consequence, we only derive an $L^2$-limit for the valuation function.
887:
888: The setting is similar to the previous sections, and we need to impose the
889: following conditions.
890: \begin{description}
891: \item[(G1)] Assume that $g\in L^1(\R^d)\cap L^2(\R^d)$.
892: \item[(G2)] Assume that $M_{X_T}(R)$ exists.
893: \end{description}
894:
895: \begin{theorem}\label{L2-valuation}
896: If the asset price process is modeled as an exponential semimartingale
897: process according to \eqref{asset-Rd}--\eqref{Hi-canon} and conditions
898: (G1)--(G2) are in force, then the time-$0$ price function satisfies
899: \begin{align}\label{L2-value}
900: \mathbb V_f(X;\cdot) =
901: \frac{\e^{-\langle R,\cdot\rangle}}{(2\pi)^d}
902: \mathop{L^2\hbox{-}\lim}_{A\rightarrow\infty}
903: \int_{[-A,A]^d}\e^{-i\langle u,\cdot\rangle}\varphi_{X_T}(u-iR)\widehat{f}(iR-u)\ud u.
904: \end{align}
905: \end{theorem}
906: \begin{proof}
907: Similarly to the previous section, we have that
908: \begin{align}\label{ch3:sec5.1}
909: \V_f(X;\logs) = \e^{-\langle R,\logs\rangle}\bg*\varrho(\logs),
910: \end{align}
911: and, for all $u\in\R^d$
912: \begin{align}\label{L2-fouval}
913: \widehat{\bg*\varrho}(u)=\widehat{\bg}(u)\cdot\widehat{\varrho}(u).
914: \end{align}
915: Now, since $\bg\in L^1(\R^d)\cap L^2(\R^d)$, we get that
916: $\widehat{\bg}\in L^2(\R^d)$ and
917: $\Vert\bg\Vert_{L^2(\R^d)}=\Vert\widehat{\bg}\Vert_{L^2(\R^d)}$; the
918: proofs are analogous to Theorem 9.13 in \citeN{Rudin87}.
919: Moreover, we have that $\bg*\varrho\in L^2(\R^d)$, because
920: \begin{align*}
921: \Vert\bg*\varrho\Vert^2_{L^2(\R^d)}
922: &\le \varrho(\R^d)^2\,\Vert\bg\Vert^2_{L^2(\R^d)}<\infty.
923: \end{align*}
924: Therefore, since also $\bg*\varrho\in L^1(\R^d)$ we get that
925: $\bg*\varrho\in L^1(\R^d)\cap L^2(\R^d)$
926: and, analogously again to Theorem 9.13 in \citeN{Rudin87}, we get
927: that $\widehat{\bg*\varrho}\in L^2(\R^d)$
928: and $\Vert\bg*\varrho\Vert_{L^2(\R^d)}=\Vert\widehat{\bg*\varrho}\Vert_{L^2(\R^d)}$.
929:
930: Therefore, the Fourier transform in \eqref{L2-fouval} can be inverted
931: and the inversion is given as an $L^2$-limit; more precisely, we have
932: \begin{align}\label{dc-val-1}
933: \Vert\bg*\varrho-\psi_A\Vert_{L^2(\R^d)}\rightarrow0
934: \qquad (A\rightarrow\infty)
935: \end{align}
936: where
937: \begin{align}\label{dc-val-2}
938: \psi_A(\logs)
939: &= \frac{1}{(2\pi)^d}\int_{[-A,A]^d}
940: \!\e^{-i\langle u,\logs\rangle}\widehat{\bg*\varrho}(u)\ud u\nonumber\\
941: &= \frac{1}{(2\pi)^d}\int_{[-A,A]^d}\!\e^{-i\langle u,\logs\rangle}
942: \widehat{f}(iR-u)\varphi_{X_T}(u-iR)\ud u.
943: \end{align}
944: Finally, \eqref{ch3:sec5.1} and \eqref{dc-val-1}--\eqref{dc-val-2}
945: yield the option price function.
946: \end{proof}
947:
948: \begin{remark}
949: The problem becomes significantly simpler when dealing with the product
950: $f_1(X_T)f_2(Y_T)$ of a continuous payoff function $f_1$ for the variable
951: $X$ and a discontinuous payoff function $f_2$ for the other variable $Y$,
952: even in the absence of Lebesgue densities. A typical example of this situation
953: is the barrier option payoff, where $f_1(x)=(\e^{x}-K)^+$ and
954: $f_2(y)=1_{\{\e^{y}>B\}}$. Then, one can make a measure change using the
955: (normalized) continuous payoff as the Radon--Nikodym derivative, apply Theorem
956: \ref{valuation} and then Theorem \ref{valuation-dc}; this leads to pointwise
957: convergence of the valuation function. The measure change argument is outlined
958: in \citeN{BorovkovNovikov02} and \citeN[Theorem 3.5]{Papapantoleon06}.
959: \end{remark}
960:
961:
962:
963: \section{Sensitivities -- Greeks}
964: \label{greeks}
965:
966: The structure of the asset price model as an exponential semimartingale,
967: and the resulting structure of the option price function, allows us to
968: easily derive general formulas for the sensitivities of the option price
969: with respect to model parameters. In this section we will focus on the
970: sensitivities with respect to the initial value, i.e. delta and gamma,
971: while sensitivities with respect to other parameters can be derived
972: analogously.
973:
974: Let us rewrite the option price function as a function of the initial
975: value, using that $S_0=\e^{-\logs}$, as follows:
976: \begin{align}\label{price}
977: \mathbb{V}_f(X;S_0)
978: &= \frac{1}{2\pi}\int_{\R} S_0^{R-iu} M_{X_T}(R-iu) \widehat{f}(u+iR) \ud u.
979: \end{align}
980: The delta of an option is the partial derivative of
981: the price with respect to the initial value. For a generic option with payoff
982: $f$, we have that
983: \begin{align}\label{delta}
984: \Delta_f(X;S_0)
985: &= \frac{\partial\mathbb{V}_f(X;S_0)}{\partial S_0}\nonumber\\
986: &= \frac{1}{2\pi}\int_{\R} \frac{\partial}{\partial S_0}S_0^{R-iu}
987: M_{X_T}(R-iu) \widehat{f}(u+iR) \ud u \nonumber\\
988: &= \frac{1}{2\pi}\int_{\R} S_0^{R-1-iu} M_{X_T}(R-iu)
989: \frac{\widehat{f}(u+iR)}{(R-iu)^{-1}}\ud u.
990: \end{align}
991: The gamma of an option is the partial derivative of the delta with respect
992: to the initial value. For a generic option with payoff $f$, we get
993: \begin{align}\label{gamma}
994: \Gamma_f(X;S_0)
995: &= \frac{\partial\Delta_f(X;S_0)}{\partial S_0}
996: = \frac{\partial^2\mathbb{V}_f(X;S_0)}{\partial^2 S_0}\nonumber\\
997: &= \frac{1}{2\pi}\int_{\R} S_0^{R-2-iu}
998: \frac{M_{X_T}(R-iu)\widehat{f}(u+iR)}{(R-1-iu)^{-1}(R-iu)^{-1}}\ud u.
999: \end{align}
1000:
1001: In the above equations we have taken \textit{for granted} that we can exchange
1002: integration and differentiation; however, this is the crucial step and we will
1003: provide sufficient conditions when we are allowed to do that. Using Satz
1004: IV.5.7 in \citeN{Elstrodt99} and the elementary inequality
1005: $|\text{Im} f|+|\text{Re} f|\le 2|f|$, we get that we can differentiate under
1006: the integral sign if there exists an integrable function $\wp$ such
1007: that for all $u\in\R$ and all $S_0>0$
1008: \begin{align*}
1009: \Big|\frac{\partial}{\partial S_0}F(u,S_0)\Big|\le \wp(u),
1010: \end{align*}
1011: where
1012: \begin{align*}
1013: F(u,S_0) = S_0^{R-iu} M_{X_T}(R-iu) \widehat{f}(u+iR).
1014: \end{align*}
1015: Now we can estimate the partial derivative of the function $F$:
1016: \begin{align}\label{D-est}
1017: \Big|\frac{\partial}{\partial S_0}F(u,S_0)\Big|
1018: &= |\e^{(R-1-iu)\log S_0}||R-iu| |M_{X_T}(R-iu) \widehat{f}(u+iR)|\nonumber\\
1019: &\le \mathpzc{c}(1+|u|)|M_{X_T}(R-iu)||\widehat{f}(u+iR)|
1020: =:\wp(u).
1021: \end{align}
1022: Analogously we can estimate for the second derivative of $F$:
1023: \begin{align}\label{G-est}
1024: \Big|\frac{\partial^2}{\partial S_0^2}F(u,S_0)\Big|
1025: &\le \mathpzc{c}'(1+|u|^2)|M_{X_T}(R-iu)||\widehat{f}(u+iR)|
1026: =:\wp'(u).
1027: \end{align}
1028:
1029: Sufficient conditions for the function $\wp$ in \eqref{D-est}, resp.
1030: $\wp'$ in \eqref{G-est}, to be integrable are that
1031: $|u||M_{X_T}(R-iu)|$, resp. $|u|^2|M_{X_T}(R-iu)|$, is integrable and
1032: $\widehat{f}(\cdot+iR)$ is bounded; the first condition dictates in particular
1033: that the measure $P_{X_T}$ -- equivalently $\varrho$ -- has a density of class
1034: $C^1$, resp. $C^2$; see Proposition 28.1 in \citeN{Sato99}. Alternatively, a
1035: sufficient condition is that the function $|u||\widehat{f}(u+iR)|$, resp.
1036: $|u|^2|\widehat{f}(u+iR)|$, is integrable and $M_{X_T}(R-i\cdot)$ is bounded,
1037: highlighting once again the interplay between the properties of the measure
1038: and the payoff function.
1039:
1040:
1041:
1042: \section{Examples of payoff functions}
1043: \label{ch3:payoffs}
1044:
1045: \subsection*{1.}
1046: Here we list some representative examples of payoff functions used in
1047: finance, together with their Fourier transforms and comment on whether
1048: they satisfy some of the required assumptions for option pricing. The
1049: calculations for the call option are provided explicitly and for other
1050: options they follow along the same lines.
1051:
1052: \begin{example}[Call and put option]\label{ch3:ex-call}
1053: The payoff of the standard call option with strike $K\in\Rp$ is $f(x)=(\e^x-K)^+$.
1054: Let $z\in\C$ with $\Im z\in(1,\infty)$, then the
1055: Fourier transform of the payoff function of the call option is
1056: \begin{align}\label{ch3:eqop:1}
1057: \widehat{f}(z)
1058: &= \int_\R\e^{izx}(\e^x-K)^+\dx
1059: = \int_{\ln K}^\infty \e^{(1+iz)x}\dx -K\int_{\ln K}^\infty \e^{izx}\dx\nonumber\\
1060: % &= -\frac{1}{1+iz}\e^{(1+iz)\ln K} + \frac{K}{iz}\e^{iz\ln K}\\
1061: &= -K^{1+iz}\frac{1}{1+iz} + K^{iz}\frac{K}{iz}
1062: = \frac{K^{1+iz}}{iz(1+iz)}.
1063: \end{align}
1064: Now, regarding the dampened payoff function of the call option, we
1065: easily get for $R\in(1,\infty)$ that $g\in\labc\cap L^2(\R)$. The weak
1066: derivative of $g$ is
1067: \begin{align}
1068: \partial g(x) =
1069: \left\{%
1070: \begin{array}{ll}
1071: 0, & \hbox{if $x<\ln K$,} \\
1072: \e^{-Rx}(\e^x-R\e^x+RK), & \hbox{if $x>\ln K$.} \\
1073: \end{array}%
1074: \right.
1075: \end{align}
1076: Again, we have that $\partial g\in L^2(\R)$. Therefore, $g\in H^1(\R)$ and
1077: using Lemma \ref{sobo} we can conclude that $\widehat{g}\in\la$. Summarizing,
1078: conditions (C1) and (C3) of Theorem \ref{valuation} are fulfilled for the
1079: payoff function of the call option.
1080:
1081: Similarly, for a put option, where $f(x)=(K-\e^x)^+$, we
1082: have that
1083: \begin{align}\label{ch3:eqop:2}
1084: \widehat{f}(z) = \frac{K^{1+iz}}{iz(1+iz)},
1085: \qquad \Im z\in (-\infty,0).
1086: \end{align}
1087: Analogously to the case of the call option, we can conclude for the dampened
1088: payoff function of the put option that $g\in\labc$ and $g\in H^1(\R)$ for $R<0$,
1089: yielding $\widehat{g}\in\la$. Hence, conditions (C1) and (C3) of Theorem
1090: \ref{valuation} are also fulfilled for the payoff function of the put option.
1091: \end{example}
1092:
1093: \begin{example}[Digital option]\label{ch3:ex-dig}
1094: The payoff of a digital call option with barrier $B\in\Rp$ is
1095: $1_{\{\e^x>B\}}$. Let $z\in\C$ with $\Im z\in(0,\infty)$, then
1096: the Fourier transform of the payoff function of the digital call
1097: option is
1098: \begin{align}\label{ch3:eqop:3}
1099: \widehat{f}(z)
1100: = -\frac{B^{iz}}{iz}.
1101: \end{align}
1102: Similarly, for a digital put option, where $f(x)=1_{\{\e^x<B\}}$,
1103: we have that
1104: \begin{align}\label{ch3:eqop:4}
1105: \widehat{f}(z) = \frac{B^{iz}}{iz},
1106: \qquad \Im z\in (-\infty,0).
1107: \end{align}
1108: For the dampened payoff function of the digital call and put option,
1109: we can easily check that $g\in\la$ for $R\in(0,\infty)$
1110: and $R\in(-\infty,0)$.
1111:
1112: Regarding the continuity and bounded variation properties of the value
1113: function, we have that
1114: \begin{align*}
1115: \mathbb{V}_f(X,\logs)
1116: &= E\big[1_{\{\e^{X_T-\logs}>B\}}\big]
1117: = P\big(X_T>\log(B)+\logs\big)
1118: = 1-F_{X_T}\big(\log(B)+\logs\big),
1119: \end{align*}
1120: where $F_{X_T}$ denotes the cumulative distribution function of $X_T$.
1121: Therefore, $\logs\mapsto\mathbb{V}_f(X,\logs)$ is monotonically decreasing,
1122: hence it has locally bounded variation. Moreover, we can conclude
1123: that $\logs\mapsto\mathbb{V}_f(X,\logs)$ is continuous if the
1124: measure $P_{X_T}$ is \textit{atomless}.
1125:
1126: Summarizing, condition (D1) is always satisfied for the payoff function of the
1127: digital option, while the prerequisites of Theorem \ref{valuation-dc} on
1128: continuity and bounded variation are satisfied if the measure $P_{X_T}$ does
1129: not have atoms.
1130: \end{example}
1131:
1132: \begin{example}
1133: A variant of the digital option is the so-called
1134: \emph{asset-or-nothing digital}, where the option holder receives
1135: one unit of the \emph{asset}, instead of \emph{currency}, depending
1136: on whether the underlying reaches some barrier or not. The payoff of
1137: the asset-or-nothing digital call option with barrier $B\in\Rp$ is
1138: $f(x)=\e^x1_{\{\e^x>B\}}$, and the Fourier transform, for $z\in\C$
1139: with $\Im z\in(1,\infty)$, is
1140: \begin{align}
1141: \widehat{f}(z)
1142: = -\frac{B^{1+iz}}{1+iz}.
1143: \end{align}
1144: Arguing analogously to the previous example, we can deduce that condition
1145: (D1) is always satisfied for the payoff function of the asset-or-nothing
1146: digital option, while the prerequisites of Theorem \ref{valuation-dc} are
1147: satisfied if the measure $P_{X_T}$ does not have atoms.
1148: \end{example}
1149:
1150: \begin{example}[Double digital option]\label{ch3:ddig}
1151: The payoff of the double digital call option with barriers
1152: $\underline{B},\overline{B}>0$ is
1153: $1_{\{\underline{B}<\e^x<\overline{B}\}}$. Let $z\in\C\setminus\lbrace0\rbrace$,
1154: then the Fourier transform of the payoff function is
1155: \begin{align}
1156: \widehat{f}(z)
1157: = \frac{1}{iz}\left(\overline{B}^{iz}-\underline{B}^{iz}\right).
1158: \end{align}
1159: The dampened payoff function of the double digital option
1160: satisfies $g\in\la$ for all $R\in\R$.
1161:
1162: Moreover, we can decompose the value function of the double
1163: digital option as
1164: \begin{align*}
1165: \mathbb{V}_f(X,\logs)
1166: &= \mathbb{V}_{f_1}(X,\logs)-\mathbb{V}_{f_2}(X,\logs),
1167: \end{align*}
1168: where $f_1(x)=1_{\{\e^{x}<\overline{B}\}}$ and
1169: $f_2(x)=1_{\{\underline{B}\le\e^{x}\}}$. Hence, by the results of Example
1170: \ref{ch3:ex-dig}, we get that condition (D1) is always satisfied for the
1171: payoff function of the double digital option, while the prerequisites of
1172: Theorem \ref{valuation-dc} are satisfied if the measure $P_{X_T}$ does not
1173: have atoms.
1174: \end{example}
1175:
1176: \begin{example}[Self-quanto and power options]
1177: The payoff of a self-quanto call option with strike $K\in\Rp$ is
1178: $f(x)=\e^x(\e^x-K)^+$. The Fourier transform of the payoff function of the
1179: self-quanto call option, for $z\in\C$ with $\Im z\in(2,\infty)$, is
1180: \begin{align}
1181: \widehat{f}(z)
1182: &= \frac{K^{2+iz}}{(1+iz)(2+iz)}.
1183: \end{align}
1184: The payoff of a power call option with strike $K\in\Rp$ and power $2$ is
1185: $f(x)=[(\e^x-K)^+]^2$; for $z\in\C$ with $\Im z\in(2,\infty)$, the Fourier
1186: transform is
1187: \begin{align}
1188: \widehat{f}(z)
1189: &= -\frac{2K^{2+iz}}{iz(1+iz)(2+iz)}.
1190: \end{align}
1191: The payoff functions for the respective put options are defined in the
1192: obvious way, while the Fourier transforms are identical, with the range
1193: for the imaginary part of $z$ being respectively $(-\infty,1)$ and
1194: $(-\infty,0)$.
1195:
1196: Analogously to Example \ref{ch3:ex-call}, we can deduce that conditions
1197: (C1) and (C3) of Theorem \ref{valuation} are fulfilled for the payoff
1198: function of the self-quanto and the power option.
1199: \end{example}
1200:
1201: \begin{remark}
1202: For power options of higher order we refer to Raible \citeyear[Chapter 3]{Raible00}.
1203: \end{remark}
1204:
1205:
1206: \subsection*{2.}
1207: Next we present some examples of payoff functions for options on
1208: several assets and for options on multiple functionals of one asset,
1209: together with their corresponding Fourier transforms.
1210:
1211: \begin{example}[Option on the minimum/maximum]\label{min-two-asset}
1212: The payoff function of a call option on the minimum of $d$ assets is
1213: \begin{align*}
1214: f(x)=(\e^{x_1}\wedge\dots\wedge\e^{x_d}-K)^+,
1215: \end{align*}
1216: for $x\in\R^d$. The Fourier transform of this payoff function is
1217: \begin{align}\label{Fou-min-d}
1218: \widehat{f}(z)
1219: = -\frac{K^{1+i\sum_{k=1}^d z_k}}{(-1)^d(1+i\sum_{k=1}^d z_k)\prod_{k=1}^d (iz_k)},
1220: \end{align}
1221: where $z\in\C^d$ with $\Im z_k>0$ for $1\le k\le d$ and
1222: $\Im(\sum_{k=1}^d z_k)>1$; for more details we refer to Appendix
1223: \ref{fou-min}. Then, we can easily deduce for the dampened payoff function
1224: that $g\in L^1_{\text{bc}}(\R^d)$.
1225:
1226: Moreover, for the put option on the maximum of $d$ assets, the payoff
1227: function is
1228: \begin{align*}
1229: f(x)=(K-\e^{x_1}\vee\dots\vee\e^{x_d})^+,
1230: \end{align*}
1231: for $x\in\R^d$, where $a\vee b=\max\{a,b\}$. The Fourier transform is
1232: \begin{align}\label{Fou-max-d}
1233: \widehat{f}(z)
1234: = \frac{K^{1+i\sum_{k=1}^d z_k}}{(1+i\sum_{k=1}^d z_k)\prod_{k=1}^d (iz_k)},
1235: \end{align}
1236: with the restriction now being $\Im z_k<0$ for all $1\le k\le d$. Again, we can
1237: easily deduce that the dampened payoff function satisfies
1238: $g\in L^1_{\text{bc}}(\R^d)$. Therefore, condition (A1) of Theorem
1239: \ref{valuation-Rd} is satisfied.
1240: \end{example}
1241:
1242: \begin{example}
1243: A natural example of multi-asset payoff functions are products of single
1244: asset payoff functions. These payoff functions have the form
1245: \begin{align*}
1246: f(x) = \prod_{\icc=1}^d f_\icc(x_\icc),
1247: \end{align*}
1248: for $x\in\R^d$, where $x_\icc\in\R$ and $f_\icc:\R\rightarrow\R_+$, for all
1249: $1\le\icc\le d$; for example, one can consider $f_1(x_1)=(\e^{x_1}-K)^+$
1250: and $f_2(x_2)=1_{\{\e^{x_2}>B\}}$.
1251:
1252: The Fourier transform of these payoff functions is simply the product
1253: of the Fourier transform of the `marginal' payoff functions, since
1254: \begin{align*}
1255: \widehat{f}(z)
1256: &= \int_{\R^d} \e^{i\langle z,x\rangle}\prod_{\icc=1}^d f_\icc(x_\icc)\dx
1257: = \prod_{\icc=1}^d\int_{\R} \e^{iz_\icc x_\icc}f_\icc(x_\icc)\ud x_\icc
1258: = \prod_{\icc=1}^d\widehat{f}_\icc(z_\icc),
1259: \end{align*}
1260: for $z\in\C^d$ and $z_\icc\in\C$, with $\Im z$ in an appropriate range such
1261: that $\widehat{g}\in L^1(\R^d)$. This range, as well as other properties of
1262: $\widehat{f}$, are dictated by the corresponding properties of the Fourier
1263: transforms $\widehat{f}_\icc$ of the marginal payoff functions $f_\icc$.
1264: \end{example}
1265:
1266: \begin{remark}
1267: Further examples of multiple asset payoff functions, such as basket
1268: and spread options, and their Fourier transforms can be found in
1269: \citeN{HubalekKallsen03}.
1270: \end{remark}
1271:
1272:
1273: \subsection*{3.}
1274: We add a short remark on the rate of decay of the Fourier transform of
1275: the various payoff functions and its consequence for numerical implementations.
1276:
1277: Consider the standard call option, where the Fourier transform of
1278: the dampened payoff function has the form, cf. \eqref{ch3:eqop:1},
1279: \begin{align*}
1280: \widehat{g}(u) = \frac{K^{1-R}\e^{iu\log{K}}}{(R-iu)(R-1-iu)},
1281: \qquad u\in\R.
1282: \end{align*}
1283: Then, we have that
1284: \begin{align*}
1285: |\widehat{g}(u)|
1286: \leq \frac{K^{1-R}}{\sqrt{R^2+u^2}\sqrt{(R-1)^2+u^2}}
1287: \leq \frac{K^{1-R}}{(R-1)^2+u^2},
1288: \end{align*}
1289: which shows that $\widehat{g}(u)$ behaves like $\frac{1}{u^2}$ for
1290: $|u|>1$. On the other hand, a similar calculation for the digital
1291: option shows that the Fourier transform of the dampened digital
1292: payoff behaves like $\frac{1}{u}$ for $|u|>1$.
1293:
1294: Therefore, splitting a call option into the difference of an
1295: asset-or-nothing digital and a digital option, as many authors have
1296: proposed in the literature (cf. e.g. \citeNP{Heston93}), is not only
1297: `conceptually' sub-optimal, as can be seen by Theorems \ref{valuation}
1298: and \ref{valuation-dc}. More importantly, it is also not optimal from
1299: the numerical perspective, since the rate of decay for the digital
1300: option is much slower than for the call option, leading to slower
1301: numerical evaluation of the corresponding option prices.
1302:
1303: Indeed, we have calculated the prices of call options corresponding
1304: to 11 strikes and 10 maturities, first using the formula for the call
1305: option, and then representing the call option as the difference of two
1306: digital options. The numerical calculation using the second method
1307: lasts twice as long (6 secs compared to less than 3 secs) in a
1308: standard Matlab implementation.
1309:
1310:
1311:
1312: \section{Examples of driving processes}
1313: \label{LA}
1314:
1315: The application of Fourier transform valuation formulas in practice requires
1316: the explicit knowledge of the moment generating function of the underlying
1317: random variable. As such, Fourier methods are tailor-made for pricing European
1318: options in \lev and affine models, since in these models one typically knows
1319: the moment generating function explicitly (at least up to the solution of a
1320: Riccati equation). In order to give a flavor, we present here an overview of
1321: \lev and affine processes, referring to the literature for specific formulas
1322: and proofs.
1323:
1324: In \lev processes, the moment generating function of the random variable is
1325: described by the celebrated \lk formula; for a \lev process \prozess[H] with
1326: triplet ($b,c,\lambda$) we have:
1327: \begin{align}
1328: E\big[\e^{\scal{u}{H_t}}\big] = \exp\left(\kappa(u)\cdot t\right),
1329: \end{align}
1330: for suitable $u\in\R^d$, where the cumulant generating function is
1331: \begin{align}\label{lk-mgf-general}
1332: \kappa(u) = \scal{b}{u} + \frac{1}{2}\scal{u}{cu}
1333: + \int_{\R^d}\left(\e^{\scal{u}{x}}-1-\scal{u}{h(x)}\right)\lambda(\dx);
1334: \end{align}
1335: here $h$ denotes a suitable truncation function. The most popular \lev models
1336: are the VG and CGMY processes (cf. \citeNP{MadanSeneta90},
1337: Carr et al. \citeyearNP{Carretal02}), the hyperbolic, NIG and GH processes
1338: (cf. Eberlein and Keller \citeyearNP{EberleinKeller95},
1339: \citeNP{Barndorff-Nielsen98}, \citeNP{Eberlein01a}), and the Meixner model (cf.
1340: \citeNP{SchoutensTeugels98}).
1341:
1342: In affine processes, the moment generating functions are described by the
1343: very definition of these processes. Let \prozess[X] be an affine process
1344: on the state space $D=\R^m\times\Rp^n\subseteq\R^d$, starting from $x\in D$;
1345: i.e., under suitable conditions, there exist functions
1346: $\phi:[0,T]\times\mathcal{I}\to\R$ and
1347: $\psi:[0,T]\times\mathcal{I}\to\R^d$ such that
1348: \begin{align}
1349: E_x\big[\e^{\scal{u}{X_t}}\big] = \exp\left(\phi_t(u) + \scal{\psi_t(u)}{x} \right),
1350: \end{align}
1351: for all $(t,u,x)\in[0,T]\times\mathcal{I}\times D$, $\mathcal{I}\subseteq\R^d$.
1352: The functions $\phi$ and $\psi$ satisfy generalized Riccati equations, while
1353: their time derivatives
1354: \begin{align*}
1355: F(u) = \frac{\partial}{\partial t}\big|_{t=0}\phi_t(u)
1356: \quad\text{ and }\quad
1357: R(u) = \frac{\partial}{\partial t}\big|_{t=0}\psi_t(u),
1358: \end{align*}
1359: are of \lk form \eqref{lk-mgf-general}; we refer to
1360: \shortciteN{DuffieFilipoviSchachermayer03} and
1361: Keller-Ressel \citeyear{KellerRessel08} for
1362: comprehensive expositions and the necessary details. The class of affine
1363: processes contains as special cases -- among others -- many stochastic
1364: volatility models, such as the \citeN{Heston93} model, the BNS model (cf.
1365: Barndorff-Nielsen and Shephard
1366: \citeyearNP{Barndorff-NielsenShephard01}, \citeNP{NicolatoVenardos03}), and
1367: time-changed \lev models (cf. \shortciteNP{Carretal03}, \citeNP{Kallsen06}).
1368:
1369:
1370:
1371: \section{Numerical illustration}
1372: \label{ch3:sc5}
1373:
1374: As an illustration of the applicability of Fourier-based valuation formulas
1375: even for the valuation of options on several assets, we present a numerical
1376: example on the pricing of an option on the minimum of two assets. As driving
1377: motions we consider a 2d normal inverse Gaussian (NIG) \lev process and a
1378: 2d affine stochastic volatility model.
1379:
1380: Let $H$ denote a 2d NIG random variable,
1381: i.e. \[H=(H^1,H^2)\sim\text{NIG}_2(\alpha,\beta,\delta,\mu,\Delta),\]
1382: where the parameters have the following domain of definition:
1383: $\alpha,\delta\in\R_+$, $\beta,\mu\in\R^2$, and
1384: $\Delta\in\R^{2\times2}$ is a symmetric, positive-definite matrix; w.l.o.g.
1385: we can assume that $\text{det}(\Delta)=1$; in addition,
1386: $\alpha^2>\scal{\beta}{\Delta\beta}$. Then, the moment generating function of
1387: $H$, for $u\in\R^2$ with $\alpha^2-\langle \beta+u,\Delta(\beta+u)\rangle\ge0$,
1388: is
1389: \begin{align}\label{2d-NIG-MGF}
1390: M_{H}(u)
1391: &= \exp\left( \langle u,\mu\rangle
1392: + \delta\left(\sqrt{\alpha^2-\langle \beta,\Delta\beta\rangle}
1393: -\sqrt{\alpha^2-\langle \beta+u,\Delta(\beta+u)\rangle}\right)\right).
1394: \end{align}
1395: In the $\text{NIG}_2$ model, we specify the parameters $\alpha,\beta,\delta$
1396: and $\Delta$, and the drift vector $\mu$ is determined by the martingale
1397: condition. Note that the marginals $H^\icc$ are also NIG
1398: distributed (cf. \citeNP[Theorem 1]{Blaesild81}), hence the drift vector can be
1399: easily evaluated from the cumulant of the univariate NIG law. The covariance
1400: matrix corresponding to the $\text{NIG}_2$-distributed random variable $H$ is
1401: \begin{align*}
1402: \Sigma_{\text{NIG}}
1403: &= \delta \left( \alpha^2-\scal{\beta}{\Delta\beta} \right)^{-\frac{1}{2}}
1404: \left( \Delta + \left( \alpha^2-\scal{\beta}{\Delta\beta} \right)^{-1}
1405: \Delta\beta\beta^{\top}\Delta\right),
1406: \end{align*}
1407: cf. \citeN[eq. (4.15)]{Prause99}. A comprehensive exposition of the
1408: multivariate generalized hyperbolic distributions can be found in
1409: \citeN{Blaesild81}; cf. also \citeN{Prause99}.
1410:
1411: We will also consider the following affine stochastic volatility model
1412: introduced by \citeN{DempsterHong02}, that extends the framework of Heston
1413: \citeyear{Heston93} to the multi-asset case. Let $H=(H^1,H^2)$ denote the logarithm
1414: of the asset price processes $S=(S^1,S^2)$, i.e. $H^\icc=\log S^\icc$; then,
1415: $H^\icc$, $\icc=1,2$ satisfy the following SDEs:
1416: \begin{align*}
1417: \ud H^1_t &= -\frac{1}{2}\sigma_1^2v_t\dt + \sigma_1\sqrt{v_t}\ud W_t^1 \nonumber\\
1418: \ud H^2_t &= -\frac{1}{2}\sigma_2^2v_t\dt + \sigma_2\sqrt{v_t}\ud W_t^2\\
1419: \ud v_t &= \kappa(\mu-v_t)\dt + \sigma_3\sqrt{v_t}\ud W_t^3, \nonumber
1420: \end{align*}
1421: with initial values $H_0^1,H_0^2,v_0>0$. The parameters have the
1422: following domain of definition: $\sigma_1,\sigma_2,\sigma_3>0$ and
1423: $\mu,\kappa>0$. Here $W=(W^1,W^2,W^3)$ denotes a 3-dimensional
1424: Brownian motion with correlation coefficients
1425: \begin{align*}
1426: \scal{W^1}{W^2}=\rho_{12},\quad
1427: \scal{W^1}{W^3}=\rho_{13},\quad\text{ and }\quad
1428: \scal{W^2}{W^3}=\rho_{23}.
1429: \end{align*}
1430: The moment generating function of the vector $H=(H^1,H^2)$ has been calculated
1431: by \citeN{DempsterHong02}; for $u=(u_1,u_2)\in\R^2$ we have
1432: \begin{align*}
1433: M_{H_t}(u)
1434: &= \exp\bigg( \scal{u}{H_0}
1435: + \frac{2\zeta(1-\e^{-\theta t})}{2\theta - (\theta-\gamma)(1-\e^{-\theta t})}\cdot v_0
1436: \nonumber\\
1437: &\qquad\qquad - \frac{\kappa\mu}{\sigma_3^2}
1438: \Big[ 2\cdot\log\Big(\frac{2\theta -
1439: (\theta-\gamma)(1-\e^{-\theta t})}{2\theta}\Big) + (\theta-\gamma)t \Big] \bigg),
1440: \end{align*}
1441: where $\zeta=\zeta(u)$, $\gamma=\gamma(u)$, and $\theta=\theta(u)$ are
1442: \begin{align*}
1443: \zeta &=
1444: \frac{1}{2}\Big(\sigma_1^2u_1^2 + \sigma_2^2u_2^2 + 2\rho_{12}\sigma_1\sigma_2u_1u_2
1445: - \sigma_1^2u_1 - \sigma_2^2u_2\Big),\\
1446: \gamma &=
1447: \kappa - \rho_{13}\sigma_1\sigma_3u_1 - \rho_{23}\sigma_2\sigma_3u_2,\\
1448: \theta &=
1449: \sqrt{\gamma^2 - 2\sigma_3^2\zeta}.
1450: \end{align*}
1451:
1452: We can deduce that all three models satisfy conditions (A2) and (A3) of Theorem
1453: \ref{valuation-Rd} for certain values of $R$. Explicit calculations for the 2d
1454: NIG model are deferred to Appendix \ref{regularity-2D-NIG}; analogous calculations
1455: yield the results for the other models.
1456:
1457: The Fourier transform of the payoff function
1458: $f(x)=(\e^{x_1}\wedge\e^{x_2}-K)^+$, $x\in\R^2$, corresponding to the option on
1459: the minimum of two assets is given by \eqref{Fou-min-d} for $d=2$, and we get
1460: that condition (A1) of Theorem \ref{valuation-Rd} is satisfied for $R_1,R_2>0$
1461: such that $R_1+R_2>1$.
1462:
1463: Therefore, applying Theorem \ref{valuation-Rd}, the price of an option on the
1464: minimum of two assets is given by
1465: \begin{align*}%\label{ch3:tac}
1466: \mathbb{MTA}_T(S^1,S^2;K)
1467: &= \frac{1}{4\pi^2}\int_{\R^2}
1468: (S_0^1)^{R_1+iu_1}(S_0^2)^{R_2+iu_2}
1469: M_{H_T}(R_1+iu_1,R_2+iu_2)\nonumber\\
1470: & \qquad \times
1471: \frac{K^{1-R_1-R_2-iu_1-iu_2}}{(R_1+iu_1)(R_2+iu_2)(R_1+R_2-1+iu_1+iu_2)}\ud u,
1472: \end{align*}
1473: where $M_{H_T}$ denotes the moment generating function of the random vector
1474: $H_T$, and $R_1,R_2$ are suitably chosen.
1475:
1476: In the numerical illustrations, we consider the following parameters: strikes
1477: \[
1478: K=\left\{85,90,92.5,95,97.5,100,102.5,105,107.5,110,115\right\}
1479: \]
1480: and times to maturity
1481: \[
1482: T = \left\{\tfrac{1}{12},\tfrac{2}{12},0.25,0.50,0.75,1.00\right\}.
1483: \]
1484: In the 2d NIG model, we consider some typical parameters, e.g.
1485: $S_0^1=100$, $S_0^2=95$, $\alpha=6.20$, $\beta_1=-3.80$, $\beta_2=-2.50$ and
1486: $\delta = 0.150$; we consider two matrices
1487: $\Delta^+=\bigl(\begin{smallmatrix}1&0\\0&1\end{smallmatrix}\bigr)$ and
1488: $\Delta^-=\bigl(\begin{smallmatrix}1&-1\\-1&2\end{smallmatrix}\bigr)$, which
1489: give positive and negative correlations respectively; indeed we get that
1490: \begin{align*}
1491: \Sigma_{\text{NIG}}^+ =
1492: \begin{pmatrix}
1493: 0.0646 & 0.0191\\
1494: 0.0191 & 0.0481
1495: \end{pmatrix}
1496: \quad \text{and} \quad
1497: \Sigma_{\text{NIG}}^- =
1498: \begin{pmatrix}
1499: 0.0287 & -0.0258\\
1500: -0.0258 & 0.0556
1501: \end{pmatrix}.
1502: \end{align*}
1503: The option prices in these two cases are exhibited in Figure \ref{Fig:2NIG}.
1504:
1505: \begin{figure}
1506: \begin{center}
1507: \includegraphics[width=6.250cm,keepaspectratio=true]{2NIG_MTA_PosCorr.eps}
1508: \includegraphics[width=6.250cm,keepaspectratio=true]{2NIG_MTA_NegCorr.eps}
1509: \caption{Option prices in the 2d NIG model with positive (left) and
1510: negative (right) correlation.}
1511: \label{Fig:2NIG}
1512: \end{center}
1513: \end{figure}
1514:
1515: Finally, in the stochastic volatility model we consider the parameters used in
1516: \citeN{DempsterHong02}, that is $S_0^1=96$, $S_0^2=100$, $\sigma_1=0.5$,
1517: $\sigma_2=1.0$, $\sigma_3=0.05$, $\rho_{12}=0.5$, $\rho_{13}=0.25$,
1518: $\rho_{23}=-0.5$, $v_0=0.04$, $\kappa=1.0$ and $\mu=0.04$; the option prices are
1519: shown in Figure \ref{Fig:2SV}.
1520:
1521: \begin{figure}
1522: \begin{center}
1523: \includegraphics[width=8.00cm,keepaspectratio=true]{2SV_MTA.eps}
1524: \caption{Option prices in the 2d stochastic volatility model.}
1525: \label{Fig:2SV}
1526: \end{center}
1527: \end{figure}
1528:
1529:
1530:
1531: \appendix
1532: \section{Fourier transforms of multi-asset options}
1533: \label{fou-min}
1534:
1535: In this appendix we outline the derivation of the Fourier transform
1536: corresponding to the payoff function of an option on the minimum of
1537: several assets; the derivation for the maximum is completely analogous
1538: and therefore omitted.
1539:
1540: The payoff of a (call) option on the minimum of $d$ assets is
1541: \begin{align*}
1542: (S^1\wedge S^2\wedge\dots\wedge S^d -K)^+.
1543: \end{align*}
1544: The payoff function $f$ corresponding to this option is given,
1545: for $x\in\R^d$, by
1546: \begin{align*}
1547: f(x)
1548: &=(\e^{x_1}\wedge\e^{x_2}\wedge\dots\wedge\e^{x_d} -K)^+
1549: = (\e^{x_1\wedge x_2\wedge\dots\wedge x_d}-K)^+. %\nonumber\\
1550: \end{align*}
1551: The following decomposition holds, if $x_i\neq x_j$ for $i\neq j$,
1552: $1\le i,j\le d$
1553: \begin{align*}
1554: f(x)
1555: = \sum_{i=1}^d(\e^{x_i}-K)^+1_{\{x_i\le x_j, \forall j\}}
1556: = \sum_{i=1}^d(\e^{x_i}-K)
1557: \prod_{\substack{j=1\\j\neq i}}^d1_{\{k<x_i<x_j\}},
1558: \end{align*}
1559: where $k=\log K$. Define also the auxiliary functions $f_i$, $1\le i\le d$, where
1560: \begin{align*}
1561: f_i(x)= (\e^{x_i}-K)\prod_{\substack{j=1\\j\neq i}}^d1_{\{k<x_i<x_j\}}.
1562: \end{align*}
1563:
1564: The dampened payoff function is $g(x)=\e^{-\langle R,x\rangle}f(x)$,
1565: where $R\in\R^d$; we define analogously the dampened $f_i$-functions,
1566: i.e. $g_i(x)=\e^{-\langle R,x\rangle}f_i(x)$. For simplicity, we first
1567: calculate the Fourier transform of the dampened $f_1$-function; for
1568: $u\in\R^d$ we get
1569: \begin{align}
1570: \widehat{g}_1(u)
1571: &= \int_{\R^d}\e^{\langle iu-R,x\rangle}(\e^{x_1}-K)\prod_{j=2}^d1_{\{k<x_1\le x_j\}}\dx\nonumber\\
1572: &= \int_k^{\infty}\int_{x_1}^{\infty}\dots\int_{x_1}^{\infty}
1573: \e^{\langle iu-R,x\rangle}(\e^{x_1}-K)\dx_d\dots\dx_1 \nonumber\\
1574: &= \int_k^{\infty}\e^{(iu_1-R_1)x_1}(\e^{x_1}-K)
1575: \left(\prod_{j=2}^d\int_{x_1}^{\infty}\e^{(iu_j-R_j)x_j}\dx_j\right)\dx_1 \nonumber\\
1576: &= \int_k^{\infty}\e^{(iu_1-R_1)x_1}(\e^{x_1}-K)
1577: \prod_{j=2}^d\left(-\frac{\e^{(iu_j-R_j)x_1}}{iu_j-R_j}\right)\dx_1\nonumber\\
1578: &= \frac{1}{\prod_{j=2}^d(R_j-iu_j)}
1579: \int_k^{\infty}\e^{\sum_{j=1}^d(iu_j-R_j)x_1}(\e^{x_1}-K)\dx_1 \nonumber\\
1580: &= \frac{1}{\prod_{j=2}^d(R_j-iu_j)}\left(
1581: -\frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}{1+\sum_{j=1}^d(iu_j-R_j)}
1582: +\frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}{\sum_{j=1}^d(iu_j-R_j)}\right)\nonumber\\
1583: &= \frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}
1584: {\prod_{j=2}^d(R_j-iu_j)\times\left(1+\sum_{j=1}^d(iu_j-R_j)\right)
1585: \times\left(\sum_{j=1}^d(iu_j-R_j)\right)} \nonumber,
1586: \end{align}
1587: subject to the conditions $R_j>0$ for all $j\ge2$ and
1588: $\sum_{j=1}^dR_j>1$.
1589:
1590: Hence, in general we have that
1591: \begin{align*}
1592: \widehat{g}_l(u)
1593: &= \frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}
1594: {\prod_{\substack{j=1\\j\neq l}}^d(R_j-iu_j)
1595: \times\left(1+\sum_{j=1}^d(iu_j-R_j)\right)\times\left(\sum_{j=1}^d(iu_j-R_j)\right)},
1596: \end{align*}
1597: subject to the conditions $R_j>0$ for all $1\le j\le d$ and
1598: $\sum_{j=1}^dR_j>1$.
1599:
1600: Now, we recall that $f(x)=\sum_{l=1}^df_l(x)$, hence $g(x)=\sum_{l=1}^dg_l(x)$
1601: which yields $\widehat{g}(u)=\sum_{l=1}^d\widehat{g}_l(u)$; therefore
1602: \begin{align*}
1603: \widehat{g}(u)
1604: &= \sum_{l=1}^d\frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}
1605: {\prod_{\substack{j=1\\j\neq l}}^d(R_j-iu_j)
1606: \times\left(1+\sum_{j=1}^d(iu_j-R_j)\right)\times\left(\sum_{j=1}^d(iu_j-R_j)\right)} \nonumber\\
1607: &= \frac{K^{1+\sum_{j=1}^d(iu_j-R_j)}}
1608: {\left(1+\sum_{j=1}^d(iu_j-R_j)\right)\times\left(\sum_{j=1}^d(iu_j-R_j)\right)}
1609: \;\sum_{l=1}^d\frac{R_l-iu_l}{\prod_{j=1}^d(R_j-iu_j)}\nonumber\\
1610: &= \frac{-K^{1+\sum_{j=1}^d(iu_j-R_j)}}
1611: {(-1)^d\prod_{j=1}^d(iu_j-R_j)\left(1+\sum_{j=1}^d(iu_j-R_j)\right)}.
1612: \end{align*}
1613: This we can also rewrite as
1614: \begin{align}
1615: \widehat{f}(z)
1616: &= - \frac{K^{1+i\sum_{j=1}^d z_j}}
1617: {(-1)^d\prod_{j=1}^d(iz_j)\left(1+i\sum_{j=1}^dz_j\right)},
1618: \end{align}
1619: subject to the conditions $\Im z_j>0$ for all $1\le j\le d$ and
1620: $\sum_{j=1}^d \Im z_j>1$.
1621:
1622:
1623:
1624: \section{Calculations for the 2d NIG model}
1625: \label{regularity-2D-NIG}
1626:
1627: By the moment generating function of the 2d NIG process, cf. \eqref{2d-NIG-MGF},
1628: it is evident that assumption (A2) is satisfied for $R\in\R^2$ with
1629: $\alpha^2-\langle\beta+R,\Delta(\beta+R)\rangle\ge0$. In order to verify
1630: condition (A3) we have to show that the function $u\mapsto M_H(R+iu)$ is
1631: integrable; it suffices to show that the real part of the exponent of
1632: $M_H(R+iu)$ decays like $-|u|$. We have
1633: \begin{align*}
1634: \log\big(M_H(R+iu)\big)
1635: &= i \langle \mu,u \rangle + \langle \mu,R\rangle + \delta \sqrt{ \alpha^2 - \langle \beta,\Delta \beta \rangle }\\
1636: &\quad - \delta \sqrt{ \alpha^2 - \langle \beta+R+iu, \Delta(\beta + R +iu)\rangle}\,.
1637: \end{align*}
1638: Recall that the product $\langle\cdot,\cdot\rangle$ over $\C^d$ is defined
1639: as follows: for $u,v\in\C^d$ set $\langle u,v\rangle = \sum_iu_iv_i$. Then
1640: \begin{multline*}
1641: \langle \beta+R+iu, \Delta(\beta + R +iu)\rangle \\ =
1642: \langle \beta+R, \Delta(\beta + R )\rangle - \langle u, \Delta u\rangle + 2 i \langle \beta+R, \Delta u\rangle
1643: \end{multline*}
1644: and since
1645: $\sqrt{z} = \sqrt{\frac{1}{2}(|z| + \Re(z))} + i \frac{\Im(z)}{|\Im(z)|}\sqrt{\frac{1}{2}(|z| - \Re(z))}$,
1646: we get
1647: \begin{align*}
1648: \lefteqn{ \Re\big( \log\big( M_H(R+iu)\big)\big)}\\
1649: &=
1650: \langle \mu,R\rangle + \delta \sqrt{ \alpha^2 - \langle
1651: \beta,\Delta \beta \rangle }
1652: -\frac{\delta}{\sqrt{2}}\Big\{
1653: \big| \alpha^2 - \langle \beta+R+iu, \Delta(\beta + R +iu)\rangle \big| \\
1654: &\quad+ \alpha^2 - \langle \beta+R,\Delta( \beta+R) \rangle + \langle u,
1655: \Delta u\rangle \Big\}^{1/2} \\
1656: &\le
1657: \langle \mu,R\rangle + \delta \sqrt{ \alpha^2 - \langle \beta,\Delta
1658: \beta \rangle }
1659: - \delta \sqrt{ \alpha^2 - \langle \beta+R,\Delta (\beta +R)\rangle +
1660: \langle u, \Delta u\rangle } \\
1661: &\le
1662: \langle \mu,R\rangle + \delta \sqrt{ \alpha^2 - \langle \beta,\Delta
1663: \beta \rangle }
1664: - \delta \sqrt{\lambda_{\min}} |u| \,,
1665: \end{align*}
1666: where $\lambda_{\min}$ denotes the smallest eigenvalue of the matrix $\Delta$.
1667:
1668:
1669:
1670: \bibliographystyle{chicago}
1671: %\bibliography{C:/WorkFiles/Papers/references}
1672: % \bibliography{H:/Papers/references}
1673: %\bibliography{/home/famuser/papapan/Papers/references}
1674: \bibliography{references}
1675:
1676:
1677:
1678: \end{document}