Explicit constructions of martingales calibrated to given implied volatility smiles∗ Laurent Cousot BNP Paribas 10 Harewood Avenue London, NW1 6AA [email protected]

Peter Carr Courant Institute, NYU 251 Mercer Street New York, NY, 10012 [email protected]

First version: March 1, 2007 This version: July 10, 2010

Abstract The construction of martingales with given marginal distributions at given times is a recurrent problem in financial mathematics. From a theoretical point of view, this problem is well-known as necessary and sufficient conditions for the existence of such martingales have been described. Moreover several explicit constructions can even be derived from solutions to the Skorokhod embedding problem. However these solutions have not been adopted by practitioners, who still prefer to interpolate the implied volatility smiles and use the explicit constructions of calibrated (jump-) diffusions, available in the literature, when a continuum of marginal distributions is known. In this paper, we describe several new constructions of calibrated martingales, which do not rely on a potentially risky interpolation of the marginal distributions but are still intuitive. Indeed, the continuous-time versions of our constructions can be interpreted as time-changed (jump-) diffusions and all the techniques developed in the above papers are transposed to our situation. Moreover, we show that the valuation of claims, depending only on the values of the underlying process at maturities where the marginal distributions are known, can be extremely efficient in this setting. For example, path-independent claims of this type can be valued by solving a finite number of ordinary (integro-) differential equations. Finally, an example of calibration to the S&P 500 market is provided.

JEL classification: C02, C60, G12, G13. AMS classification codes: 60G42, 60J05, 60J35, 60J75, 91B25, 91B70. Key words: martingale, marginal distribution, diffusion, jump-diffusion, calibration, option, time change.



We are grateful to Jim Gatheral and Robert Kohn for insightful discussions and comments. We also thank David Hobson for drawing our attention to several papers about the Skorokhod embedding problem. All errors are of course our own.

1

1

Introduction

The construction of martingales with constrained marginal distributions is an omnipresent problem in mathematical finance. Indeed, the assumption of absence of arbitrage often translates into the fact that the underlying asset must be modeled by a martingale in some measure (see [22] and [23]). Moreover, when pricing exotic contingent claims, whose hedging requires the use of European options, one needs to have a model, which is consistent with the prices of these options. In this paper, we assume that European option prices are known at a finite number of maturities, or, equivalently, that the marginal distributions of the underlying asset are known at these maturities (see [9]) and we address the problem of constructing a martingale with these marginal distributions. This problem was first tackled by [33], who described necessary and sufficient conditions on the marginal distributions for the existence of a calibrated martingale. In the case of two marginal distributions - which is considered in this paper without loss of generality - Strassen’s theorem can be stated as follows: Theorem 1 (Strassen’s theorem (1965)) If two probability distributions µ1 and µ2 on R∗+ have the same finite first moment, X0 , and satisfy: Z +∞ Z +∞ f (x) µ1 (dx) 6 f (x) µ2 (dx) (1) 0

0

for every convex function f on R∗+ (where the integrals are possibly infinite), then there exists two random variables, X1 and X2 , defined on the same probability space, such that: • the probability distribution of Xi is µi for i ∈ {1, 2}, • the process (Xi ) is a martingale, i.e. E [ X2 | X1 ] = X1 .

(2)

Constructive proofs of this result have been known for a while, as all solutions to the Skorokhod embedding problem1 , which accept non trivial starting laws, correspond to constructions of calibrated martingales. Examples of such constructions include those of [14] or [3]. (We refer to [25] and the references therein for more examples and also for applications of solutions to the Skorokhod embedding problem in financial mathematics.) However, to the best of our knowledge, none of the so constructed processes are realistic enough to be used in the industry. Indeed, practitioners prefer to interpolate the given implied volatility smiles to obtain a whole implied volatility surface2 and then use the explicit constructions of calibrated continuous-time martingales available in the literature. We refer to [17] and [18] when the underlying process is a diffusion, to [2] and [11] when it is a jump-diffusion and, finally, to [12] when it is a deterministically time-changed L´evy process. The purpose of this paper is to describe several new constructions of calibrated martingales, (i) which do not require the construction of the whole implied volatility surface and, (ii) which are intuitive and realistic enough for practitioners. Our constructions can actually be interpreted as the discrete-time counterparts of local volatility models (Section 4), of local volatility models with jumps (Section 5.2), of finite activity local L´evy 1 2

We refer to [31] for a survey on this problem. See [1], [26] and the references therein for more details.

2

models (Section 5.3) and of localizable jump-diffusion models (Section 5.4). Indeed, they use transition distributions, which are those of (jump-) diffusions, sampled at independent and exponentially distributed random times (see Section 2) and all the techniques, which were developed in the above papers can be transposed to our discrete-time setting. Moreover, our constructions are not confined to discrete-time martingales. Indeed, if the distributions µ1 and µ2 correspond to given maturities T1 and T2 , such that T0 ≡ 0 < T1 < T2 , we also explain how to construct a continuous-time martingale (Xtc ), which starts at X0 and such that the distribution of XTc i is µi , for i ∈ {1, 2}. And, as one could expect, (Xtc ) can be understood as a time-changed (jump-) diffusion. Finally, as far as pricing is concerned, the only technique, which is missing is the valuation of claims on (Xtc ) using backward partial integro-differential equations, since this process is not a Markov process a priori. However, (Xi ) is a Markov process and the valuation of path-independent claims on (Xi ) can be performed by solving a finite number of ordinary (integro-) differential equations. Moreover Monte Carlo schemes, very similar to those used for jump-diffusions, are available for valuating complex claims on (Xi ), (Xtc ) or in a multinomial setting. This paper is organized as follows. Section 2 defines Sampled Jump-Diffusion (SJD) distributions, which will be used later as transition distributions of the process (Xi ). Section 3 describes a necessary condition linking the parameters of a SJD distribution and the marginal distributions µ1 and µ2 . Section 4 exhibits sufficient conditions for sampled diffusion transition distributions to define a calibrated martingale. Section 5 does the same, when the transition distributions are those of a sampled jump-diffusion, and transpose to the discrete-time setting the techniques used by [2], [12] and [11]. Section 6 addresses the problem of constructing a continuous-time martingale, calibrated to the µi . Section 7 explains how the valuation of European style path-independent claims can be reduced to the valuation of ordinary (integro-) differential equations. Section 8 gives an example of calibration to S&P 500 option quotes. Section 9 concludes.

2

Definition and properties of Sampled Jump-Diffusion distributions

To construct the distributions of X2 | X1 , we first consider a positive time-homogeneous Markov process (Ft ), whose generator G, is, for f ∈ C0∞ R∗+ – the space of infinitely differentiable functions with compact support on R∗+ : 1 Gf (x) ≡ x2 σ 2 (x) f ′′ (x) + λ (x) 2

Z

+∞

 f (y) − f (x) − f ′ (x) (y − x) γ (x, y) dy,

0

(3)

where xσ (x) corresponds to the diffusion coefficient, λ is the intensity of the counting process and γ (x, .) is the density of ( Ft | (Ft− = x) & (∆Ft 6= 0)). We assume in the following that: • σ is a positive, bounded and continuous function, • λ is a non-negative, bounded and continuous function, • If λ 6= 0, γ is a non-negative function such that: – γ (x, .) is a probability density on R∗+ for every x ∈ R∗+ , R +∞ y – 0 x γ (x, y) dy is bounded by above by a constant A and is continuous, 3



R

y ln( x ) Γ 1+(ln( y ))2 γ x

(x, y) dy is continuous in x for every Γ, Borel set of R∗+ .

The above assumptions ensure that the process (Ft ) exists and is a strong Markov process. Indeed the martingale problem associated to its logarithm is well-posed thanks to [34], who generalized to processes with L´evy generators the seminal results of [35] and [36]. (See Proposition 2.1 in [15] for the details.) Moreover, if we further assume, in the case λ 6= 0, that: Z +∞   y y (4) γ (x, y) dy 6 B < +∞, ln x x 0

then (Ft ) is a martingale since it satisfies the sufficient conditions for a positive local martingale to be a martingale described by [29]. (See Proposition 2.3 in [15] for the details.) We now have all the elements to construct X2 | X1 . Indeed, if (τx ) denotes a family of random times, which have an exponential distribution of parameter 1 and are independent of (Ft ), we define, as suggested in the introduction: law

( X2 | X1 = x1 ) ≡

 Fτx1 F0 = x1 .

(5)

The distributions of X2 | X1 will be called Sampled Jump-Diffusion (SJD) distributions in the following. Notice that this concept generalizes the one of generalized Laplace distributions, introduced in [11], which corresponds to the particular case where (Ft ) is a diffusion. In the above mentioned paper, the authors show in particular that generalized Laplace distributions are centered and the Green’s functions of linear second-order differential operators (see Appendix A for more details). As expected, SJD distributions have similar properties. Indeed, notice that E [ X2 | X1 = x1 ] = x1 , (6)  ∗ ∞ since τx1 is independent of (Ft ), which is a positive martingale. Furthermore, if P ∈ C0 R+ , the forward and backward Kolmogorov equations are respectively: ∂ (E [ P (Ft )| F0 ]) = E [ GP (Ft )| F0 ] , ∂t ∂ (E [ P (Ft )| F0 ]) = GE [ P (Ft )| F0 ] . ∂t

(7) (8)

Multiplying them by e−t and integrating over R∗+ gives, using integration by parts: E [ ([I − G] P ) (Fτ )| F0 ] = P (F0 ) ,

(9)

[I − G] E [ P (Fτ )| F0 ] = P (F0 ) ,

(10)

E [ ([I − G] P ) (X2 )| X1 ] = P (X1 ) ,

(11)

i.e.

[I − G] E [ P (X2 )| X1 ] = P (X1 ) .

(12)

The above equations are based on the fact that the resolvent of the generator G is linked to the Laplace transform of the associated transition semigroup (see Proposition 2.1 in [19], p. 10) and will prove to be decisive in the following. Indeed, Equation (11) will be used in the next section to derive a necessary condition for (Xi ) to be calibrated and Equation (12) will be used in Section 7 for the pricing of path-independent claims. 4

3

A necessary condition for calibration

In this section, we derive heuristically a necessary condition that must be satisfied by σ, λ and γ if X1 and X2 are calibrated respectively to the call price functions C1 and C2 . Informally, in the case where P (x) = (x − K)+ and K > 0, Equation (11) implies:    E [I − G] (x − K)+ (X2 ) X1 = (X1 − K)+ . (13)

Moreover,

[I − G] (x − K)+ 1 = (x − K)+ − x2 σ 2 (x) δ (x − K) 2 Z +∞  − λ (x) (y − K)+ − (x − K)+ − H (x − K) (y − x) γ (x, y) dy,

(14)

0

where H is the Heaviside function. If we denote by Ψγ (x, K) the double tail associated to the density γ (x, .):    R +∞ R +∞  γ (x, v) dv du if K > x  u  K   R R R  R +∞ K u +∞ 1 Ψγ (x, K) ≡ (15) γ (K, v) dv du + 0 if K = x 2 u 0 γ (K, v) dv du K     RK Ru if K < x 0 0 γ (x, v) dv du then we obtain, using put-call parity:

1 [I − G] (x − K)+ = (x − K)+ − x2 σ 2 (x) δ (x − K) − λ (x) Ψγ (x, K) . 2

(16)

Therefore, Equation (13) can be rewritten as:   1 2 2 + E (X2 − K) − X2 σ (X2 ) δ (X2 − K) − λ (X2 ) Ψγ (X2 , K) X1 = (X1 − K)+ . 2

(17)

and integrating over X1 , we get:

1 C2 (K) − C1 (K) = K 2 σ 2 (K) P [X2 = K] + E [λ (X2 ) Ψγ (X2 , K)] . 2 If we further assume that X2 admits a density, we have, in the spirit of [9], P [X2 = K] = and the above equation becomes: Z +∞ 2 ∂ C2 ∂ 2 C2 1 (K) + (x) λ (x) Ψγ (x, K) dx. C2 (K) − C1 (K) = K 2 σ 2 (K) 2 ∂K 2 ∂x2 0

(18) ∂ 2 C2 ∂K 2

(K)

(19)

Notice that the problem of finding σ, λ and γ, which satisfy this equation, is very similar to the one of calibrating a (jump-) diffusion to an implied volatility surface. The only difference is that the time derivative is replaced by a time difference. For instance, when λ = 0, the discrete-time counterpart of the well-known Dupire’s formula (see [18]) is: σ 2 (K) =

C2 (K) − C1 (K) 1 2 ∂ 2 C2 2 K ∂K 2

5

(K)

.

(20)

4

Sufficient conditions in the case of generalized Laplace distributions

In this section, we derive sufficient conditions on the marginal distributions µ1 and µ2 , for the existence of centered transition densities of the generalized Laplace type (see Appendix A). As explained in the previous section, this construction can be interpreted as a discrete-time equivalent of local volatility models developed by [17] and [18]. Proposition 2 Let us denote by C1 (resp. C2 ) the call price function associated to the probability distribution µ1 (resp. µ2 ) on R∗+ : Ci (K) ≡

Z

+∞

0

(x − K)+ µi (dx) for i ∈ {1, 2} .

(21)

We assume that: • µ1 and µ2 have the same finite first moment X0 , • C1 (K) < C2 (K) for K ∈ R∗+ , • µ1 admits a continuous density f1 or is a Dirac function in X0 , • µ2 admits a continuous and positive density f2 , • the following equation:

σ 2 (K) =

C2 (K) − C1 (K) 1 2 2 K f2 (K)

(22)

defines a bounded function σ 2 on R∗+ . Then there exists a unique family of generalized Laplace densities (p (. |x1 )) associated to the function x2 σ 2 (x), i.e. of functions, which solves:   ∂ 2 x22 σ 2 (x2 ) p (x2 |x1 ) = δ (x2 − x1 ) , (23) I− 2 2 ∂x2   2 2 x2 σ (x2 ) ∂ p (x2 |x1 ) → 0 or +∞ 0, (24) ∂x2 2

2 for (x1 , x2 ) ∈ R∗+ . Moreover,

Z

Z

+∞

0 +∞

0

x2 p (x2 |x1 ) dx2 = x1 , p (x2 |x1 ) µ1 (dx1 ) = f2 (x2 ) .

(25) (26)

Proof. See Appendix B. In the above proposition, the assumption that σ 2 is a bounded function is not straightforward to test on the call price functions Ci . The next lemma describes sufficient conditions on f2 alone, which are based on Karamata’s theory. We refer to [7] for a clear and exhaustive description of this theory. (See Appendix C for a reminder of the main definitions.) 6

Lemma 3 With the notations of Proposition 2, if f2 has upper Matuszewska index α (f2 ) < −2, then: σ 2 (K) =0 O (1) .   If fe2 (x) ≡ f2 (1/x) has upper Matuszewska index α fe2 < 1, then: σ 2 (K) =+∞ O (1) .

Proof. See Appendix D. Remark 4 Note that the above assumptions are not very restrictive since f2 is a probability density with finite first moment. For instance, simple computations show that if f2 is a log-normal density,   e then α (f2 ) = α f2 = −∞ and the assumptions of the above lemma are therefore satisfied.

Some practitioners prefer to fit functionals to the implied volatility smiles instead of thinking in terms of risk-neutral densities. This is why, it could be interesting to find sufficient conditions on these functionals for the conclusion of the above lemma to be valid. Of course, several functionals are possible, however the results of [28], [5] and [4] give an insight on how the extrapolation should be performed at extreme strikes. These considerations led [21] to propose the so-called Stochastic Volatility Inspired (SVI) parametrization of the implied variance:   q 2 2 2 σimp (K) = a + b ρ (k − m) + (k − m) + σ . (27) with k ≡ ln



K X0

 . A little bit more generally, we assume that the implied cumulated variance is

2 given, at time T2 , by a functional σimp (k), with a slight abuse of notations, which satisfies: 2 σimp (k)

β±

=±∞ α± |k|



β± 1± +o |k|

 2 (k) ∼±∞ ±α± β± |k|β± −1 , ∂k σimp  2 (k) =±∞ o (1) , ∂kk σimp



1 |k|



,

(28) (29) (30)

with 0 < β± 6 1, α± > 0 and α± < 2 if β± = 1. Then, using classical expansions of the normal cumulative distribution function, it is possible to compute the lower and upper Matuszewska indices of f2 and fe2 : 1 − α2+ α (f2 ) = β (f2 ) = − 2 + 2α+

2 !

α (f2 ) = β (f2 ) = −∞, α− 2     1 + 2 < 1, α fe2 = β fe2 = 2 − 2α−     α fe2 = β fe2 = −∞,

< −2,

if β+ = 1,

(31)

if β+ < 1,

(32)

if β− < 1,

(33)

if β− = 1.

(34)

Consequently, under the above assumptions, σ 2 is bounded in a neighborhood of 0 or +∞, and therefore on R∗+ by continuity.

7

5

Sufficient conditions in the case of SJD distributions

In this section, we describe general sufficient conditions for SJD distributions to be suitable transition distributions for (Xi ). Furthermore, we adapt to the discrete-time setting the techniques, developed in [2], [12] and [11], to calibrate a jump-diffusion to a continuum of marginal densities.

5.1

General sufficient conditions

To describe sufficient conditions in the case of SJD transition distributions is more complicated than in the case of general Laplace distributions. Indeed, in the latter case, we knew sufficient conditions for the associated boundary value problems to have a unique solution and that generalized Laplace distributions admitted smooth enough densities. The situation is completely different here. First, to the best of our knowledge, Green’s functions associated to the operator [I − G] or its adjoint on the unbounded set R∗+ did not receive as much attention in the general case as in the case where G is a linear second-order differential operator. Moreover, sufficient conditions for SJD distributions to admit densities does not seem to be readily available in the literature. To tackle this problem, a possibility could have been to use the sufficient conditions for jump-diffusions to admit smooth densities. (We refer to [6], [37], [27] and [13].) However, all these results are valid for processes, which are strong solutions of stochastic differential equations with jumps, and as remarked in [6], to translate results in this setting to the setting of the martingale problem of Section 2 is not painless. Proposition 5 Let us denote by C1 (resp. C2 ) the call price function associated to the probability distribution µ1 (resp. µ2 ) on R∗+ : Ci (K) ≡

Z

0

+∞

(x − K)+ µi (dx) for i ∈ {1, 2} .

We assume (A1 ), that is: • µ1 and µ2 have the same finite first moment X0 , • C1 (K) < C2 (K) for K ∈ R∗+ , • µ1 admits a continuous density f1 or is a Dirac function in X0 , • µ2 admits a continuous and positive density f2 . We further assume (A2 ), i.e. • σ is a positive, bounded and continuous function on R∗+ , • λ is a non-negative, bounded and continuous functions on R∗+ , 2 • γ is a function from R∗+ to R+ such that: – γ (x, .) is a probability density for x ∈ R∗+ , R – R∗ xy γ (x, y) dy is continuous on R∗+ , +



R

y ln( x ) Γ 1+(ln( y ))2 γ x

(x, y) dy is continuous in x for every Γ, Borel set of R∗+ ,

8

(35)

– there exists A < +∞ such that, for x ∈ R∗+ , Z

+∞ 0

  y   y max 1, ln γ (x, y) dy 6 A, x x

(36)

• λ, γ and σ satisfy Equation (19), where Ψγ is defined by Equation (15). Then there exists a positive strong Markov martingale (Ft ), which is right-continuous with left limits, whose generator G is, for f ∈ C0∞ R∗+ : 1 Gf (x) ≡ x2 σ 2 (x) f ′′ (x) + λ (x) 2

Z

+∞

 f (y) − f (x) − f ′ (x) (y − x) γ (x, y) dy.

0

(37)

If we further assume (A3 ), i.e. the distribution of Fτx1 F0 = x1 – where (τx ) is a family of exponentially distributed random times with mean 1, independent of (Ft ) – admits a smooth enough density p (. |x1 ), such that the following limit conditions hold:   ∂ 1 2 2 p ( x2 | x1 ) x1 σ (x1 ) f2 (x1 ) → 0, (38) x1 →0 or +∞ ∂x1 2   1 2 2 ∂p ( x2 | x1 ) x σ (x1 ) f2 (x1 ) → 0, (39) x1 →0 or +∞ ∂x1 2 1 p ( x2 | x1 ) (x1 µ (x1 ) f2 (x1 )) → 0, (40) x1 →0 or +∞

then p is a suitable transition density between T1 and T2 : Z +∞ x2 p ( x2 | x1 ) dx2 = x1 , 0 Z +∞ p ( x2 | x1 ) µ1 (dx1 ) = f2 (x2 ) .

(41) (42)

0

Remark 6 An intuitive necessary condition implied by Equation (19) is: σ 2 (K) 6

C2 (K) − C1 (K) 1 2 ∂ 2 C2 2 K ∂K 2

(K)

,

(43)

where we recognize on the RHS the expression for σ 2 , which allows us to fit the marginal distributions when the transition densities are of the generalized Laplace type (see Section 4). Informally, the presence of jumps forces the diffusion component of the sampled process to be less volatile than the one of a sampled diffusion, which recovers the option prices. Proof. See Appendix E. Of course, there are many ways of choosing λ, γ and σ so that the assumptions (A2 ) – and Equation (19) in particular – are satisfied. In the following sections, we mimic, in a discrete-time setting, the strategies described respectively by [2], [12] and [11].

5.2

A solution ` a la Andersen and Andreasen (2000)

In [2], the authors calibrate a jump-diffusion process to an implied volatility surface through the diffusion coefficient once the jump intensity and the jump distributions have been specified. We 9

adopt a similar approach in this subsection by first fixing λ, γ, then defining σ via Equation (19), i.e. R +∞ ∂ 2 C2 (y) λ (y) Ψγ (y, x) dy C2 (x) − C1 (x) − 0 ∂y 2 2 σ (x) ≡ . (44) 2 1 2 ∂ C2 2 x ∂x2 (x) However, for this recipe to be valid, the above equation needs especially to define a positive quantity and it is clear that not all pairs (λ, γ) will do. The purpose of the next lemma is to give sufficient conditions on γ for the existence of a non-trivial λ and therefore σ satisfying the assumptions of Proposition 5. Lemma 7 With the notations of Proposition 5, we assume that: • µ1 , µ2 and γ satisfy the assumptions (A1 ) and (A2 ),  Z +∞ Z +∞ + (x − v) γ (u, v) dv f2 (u) du • P2,γ (x) ≡ 0

=0

• C2,γ (x) ≡

0

O (C2 (x) − C1 (x)) ,  Z +∞ Z +∞ + (v − x) γ (u, v) dv f2 (u) du 0

(45)

0

=+∞ O (C2 (x) − C1 (x)) ,

then there exist a positive, bounded and continuous function λ such that: Z +∞ 2 ∂ C2 (y) (C2 (x) − C1 (x)) > λ (y) Ψγ (y, x) dy. ∂y 2 0

(46)

(47)

If wefurther assume that f2 (resp. fe2 (x) ≡ f2 (1/x)) has upper Matuszewska index α (f2 ) < −2  (resp. α fe2 < 1), then the function σ, defined by Equation (44), λ and γ satisfy the assumptions (A2 ). Remark 8 Note that P2,γ (x) (resp. C2,γ (x)) is the price of a put (resp. call) option struck at x, when the risk-neutral distribution is the one of a random variable Γ, such that the density of Γ| X2 = x2 is γ (x2 , .) and the density of X2 is f2 . Proof. See Appendix F.

5.3

A solution ` a la Carr et al. (2004)

Another approach to calibrate a jump-diffusion to an implied volatility surface is the one of [12], where the authors consider a stochastic differential equation, which is not driven by a Brownian motion but by a general L´evy process and explain how to recover the local speed function from option prices. In our setting, their method would translate into fixing the diffusion coefficient σ, the jumpdensities γ and finding a function λ, which would satisfy Equation (19). However, for their method to apply, we need to specify a special structure for γ, i.e. 1 y  γ (x, y) ≡ g , (48) x x t where g is a probability density on R∗+ . Note that g is the density of FF− ∆Ft 6= 0 and that, in t this case: y  Ψγ (x, y) = xΨg , (49) x 10

where Ψg is the double tail associated to g, i.e.  R   +∞ R +∞  g (v) dv du if x > 1  u  x   R R R  R +∞ x u +∞ 1 Ψg (x) ≡ g (v) dv du + 0 0 g (v) dv du if x = 1 . u 1 2   R R   x u if x < 1 0 0 g (v) dv du

(50)

Consequently, Equation (19) becomes:

1 ∂ 2 C2 C2 (y) − C1 (y) − y 2 σ 2 (y) (y) = 2 ∂y 2

Z

0

+∞

y ∂ 2 C2 dx, (x) λ (x) xΨ g ∂x2 x

(51)

and changing variables to consider log-prices, we get:  1   2 2  ∂ 2 C2 C2 exp y ′ − C1 exp y ′ − exp y ′ exp y ′ σ exp y ′ 2 2 ∂y Z +∞ 2    2 ∂ C2 = exp x′ λ exp x′ exp x′ Ψg exp y ′ − x′ dx′ , 2 −∞ ∂x

(52)

with y = exp (y ′ ) and x = exp (x′ ). Therefore, the above equation becomes:

if A, B and C are defined by:

 A y′ =

Z

+∞

−∞

  B x′ C y ′ − x′ dx′ ,

   ∂ 2 C2 1 ′ exp 2y ′ σ 2 exp y ′ exp y , 2 ∂y 2    2 ∂ 2 C2 B x′ ≡ exp x′ λ exp x′ exp x′ , 2 ∂x   C z ′ ≡ Ψg exp z ′ . A y′



≡ C2 exp y ′



− C1 exp y ′





(53)

(54) (55) (56)

Since the RHS of Equation (53) is a convolution, we obtain, by taking the Fourier transform on both sides:     ′ −1 FA x′ , (57) B x =F FC where Ff (resp. F −1 f ) denotes the (resp. inverse) Fourier transform of f . Finally, we remind you that: B (ln (x)) . (58) λ (x) = ∂ 2 C x2 ∂x22 (x) using Equation (55). The above considerations result in the next proposition. Lemma 9 With the notations of Proposition 5, we assume (A1 ) and that: • f2 satisfies the following growth condition: Z +∞ x (ln (x))α2 f2 (x) dx < +∞ with α2 > 1, 1

11

(59)

• g is a continuous probability density on R∗+ such that: Z +∞ x (ln (x))αg g (x) dx < +∞ with αg > 1, 1 Z 1 x |ln (x)| g (x) dx < +∞ ,

(60) (61)

0

• σ is a positive, continuous and bounded function on R∗+ , such that: σ 2 (x) 6

C2 (x) − C1 (x) . 1 2 2 x f2 (x)

(62)

In that case, we define A (resp. C) via Equation (54) (resp. (56)) and denote by FA (resp. FC) its Fourier transform, which is well-defined. We further assume that: • FA/FC ∈ L1 (R), • B ≡ F −1 (FA/FC) ∈ L1 (R), • the following equation: λ (x) ≡

B (ln (x)) 2

x2 ∂∂xC22 (x)

(63)

defines a non-negative, bounded and continuous function.  Then, σ, λ and γ (x, .) ≡ x1 g x. satisfies the assumptions (A2 ) of Proposition 5.

Proof. See Appendix G.

5.4

A solution ` a la Carr and Cousot (2007a)

Another recipe to calibrate a jump-diffusion to an implied volatility surface is the one described in [11], where the authors explain how using jump-densities of the generalized Laplace type greatly simplifies both pricing and calibration. Indeed, in this case, the forward and backward Kolmogorov equations can be transformed into partial differential equations, what makes the calibration and the pricing of the same order of complexity as in a diffusion setting. As expected, it is possible to translate this technique to the discrete-time framework. Indeed, if γ are generalized Laplace densities associated to a given function a2 , then, according to Lemma 16 in Appendix A: a2 (y) a2 (x) γ (y, x) = γ (x, y) , (64) Ψγ (x, y) = 2 2 and Equation (19) becomes: Z +∞ 2 a2 (x) 1 2 2 ∂ 2 C2 ∂ C2 (x) λ (x) γ (y, x) dx = C (y) − C (y) − y σ (y) (y) . (65) 2 1 ∂x2 2 2 ∂y 2 0 h i 2 ∂2 Consequently, applying the operator I − a 2(y) ∂y yields: 2    a2 (y) a2 (y) ∂ 2 y 2 σ 2 (y) ∂ 2 C2 ∂ 2 C2 (y) λ (y) = I− (y) , C2 (y) − C1 (y) − ∂y 2 2 2 ∂y 2 2 ∂y 2 12

(66)

h since I −

a2 (y) ∂ 2 2 ∂y 2

i

γ (y, x) = δ (y − x). This equation can be rewritten as:

a2 (y) = 2

2

∂2 ∂y 2

or λ (y) =

h

C2 (y) − C1 (y) − 21 y 2 σ 2 (y) ∂∂yC22 (y)   , 2 2 C2 (y) − C1 (y) − 21 y 2 σ 2 (y) ∂∂yC22 (y) + ∂∂yC22 (y) λ (y)

I−

a2 (y) ∂ 2 2 ∂y 2

i

 2 C2 (y) − C1 (y) − 21 y 2 σ 2 (y) ∂∂yC22 (y) a2 (y) ∂ 2 C2 2 ∂y 2

(y)

.

(67)

(68)

The above equations suggest two strategies: • the first one (henceforth Strategy A) would be to choose first σ and λ, and take a2 according to Equation (67); • the second one (henceforth Strategy Λ) would be to take λ via Equation (68) once a2 and σ have been chosen. But, before attempting to find sufficient conditions for these strategies in the following sections, we need to ensure that the generalized Laplace distributions γ satisfy the assumptions of Proposition 5. This is the aim of the following lemma. Lemma 10 If a2 is a continuous function on R∗+ , which satisfies: m2 6

a2 (x) 6 M 2, x2

(69)

for given m, M ∈ R∗+ , then there exist a unique family of generalized Laplace distributions γ associated to the function a2 . Moreover, the assumptions (A2 ) of Proposition 5, which only concern γ, are satisfied. Proof. This lemma corresponds to the particular case of Lemma 2.5 in [15], where a2 is timehomogeneous. 5.4.1

Strategy A

In this section, we focus on Strategy A, which is the discrete-time equivalent of the strategy by the same name developed in [11]. Lemma 11 With the notation of Proposition 5, we assume (A1 ) and that: • µ1 admits a continuous density f1 , • σ is a positive, bounded and continuous function on R∗+ , such that: σ 2 (x) <

C2 (x) − C1 (x) 1 2 ∂ 2 C2 2 x ∂x2

(x)

• λ is a non-negative, bounded and continuous function on R∗+ ,

13

(70)

• the following equation defines a continuous and positive function a2 : a2 (x) = 2

2

∂2 ∂x2

C2 (x) − C1 (x) − 12 x2 σ 2 (x) ∂∂xC22 (x)   , 2 2 C2 (x) − C1 (x) − 12 x2 σ 2 (x) ∂∂xC22 (x) + λ (x) ∂∂xC22 (x)

(71)

which satisfies the growth conditions of Equation (69).

Then, γ, the unique family of generalized Laplace distributions associated to a2 , σ and λ satisfy the assumptions (A2 ) of Proposition 5. Proof. See Appendix H. Note that a possible specification of σ, which ensures that Equation (70) is satisfied, is: σ 2 (x) = α2

C2 (x) − C1 (x) 1 2 2 x f2 (x)

(72)

with 0 < α2 < 1. Since we recognize on the RHS, the expression for σ 2 , which allows one to fit option prices using a sampled diffusion (see Section 4), the parameter α2 allows one to control how much of the smile should be explained by the diffusion component of (Ft ). Moreover, in this case, Equation (71) greatly simplifies: (C2 (x) − C1 (x)) a2 (x) = 2 (f2 (x) − f1 (x)) + β 2 λ (x) f2 (x)

(73)

 with β 2 = 1/ 1 − α2 . The next lemma, which uses concepts belonging to Karamata’s theory (see Appendix C), describes sufficient conditions for these specifications to be valid. Lemma 12 With the notations of Proposition 5, we assume (A1 ) and that: • µ1 admits a continuous density f1 , • f2 has bounded decrease and upper Matuszewska index α (f2 ) < −2,

  • fe2 (x) ≡ f2 (1/x) has bounded decrease and upper Matuszewska index α fe2 < 1, • There exists M > 1, such that f2 (x) > M f1 (x) in a neighborhood of 0 and +∞.

Then there exist a positive function λ, such that σ, defined by Equation (72), a2 , defined by Equation (73) and λ satisfy the assumptions of Lemma 11. Proof. See Appendix I. 5.4.2

Strategy Λ

In this section, we describe sufficient conditions for the feasibility of Strategy Λ, which is the discrete-time counterpart of the strategy by the same name introduced in [11]. Lemma 13 With the notations of Proposition 5, we assume (A1 ) and that: • µ1 admits a continuous density f1 , 14

• σ is a positive, bounded and continuous function such that, for x ∈ R∗+ : C2 (x) − C1 (x)

σ 2 (x) 6

1 2 ∂ 2 C2 2 x ∂x2

(x)

,

(74)

• a2 is a continuous function, which satisfies: m2 6

a2 (x) 6 M2 x2

(75)

for x ∈ R∗+ , with m, M > 0, • the following equation λ (x) =

h

I−

a2 (x) ∂ 2 2 ∂x2

i  2 C2 (x) − C1 (x) − 12 x2 σ 2 (x) ∂∂xC22 (x) a2 (x) ∂ 2 C2 2 ∂x2

(x)

(76)

defines a continuous, non-negative and bounded function. Then, γ, the unique generalized Laplace distributions associated to a2 , σ and λ satisfy the assumptions (A2 ) of Proposition 5. Proof. This proof is in every point similar to the one of Lemma 11 (see Appendix H) and is left to the reader. Note that a specification of σ and a2 , which greatly simplifies Equation (76), is: σ 2 (x) = α2

C2 (x) − C1 (x) , 1 2 2 x f2 (x)

(77)

a2 (x) 2

C2 (x) − C1 (x) , f2 (x)

(78)

= β2

with α2 , β 2 ∈ (0, 1). Indeed, in this case, Equation (76) reduces to:     f1 (x) 1 2 −1 + . λ (x) = 1 − α β2 f2 (x)

(79)

Moreover the parameters α2 and β 2 have a clear interpretation: α2 allows one to explain how much of the smile should be explained by the diffusion component of (Ft ) and β 2 allows one to control the size of its jumps. The next lemma describes sufficient conditions for this simplified strategy to work. Lemma 14 With the notations of Proposition 5, we assume (A1 ) and that: • µ1 admits a continuous density f1 , • f2 has bounded decrease and upper Matuszewska index α (f2 ) < −2,

  • fe2 (x) ≡ f2 (1/x) has bounded decrease and upper Matuszewska index α fe2 < 1, • There exists M > 1, such that f2 (x) > M f1 (x) in a neighborhood of 0 and +∞.

Then Equations (77), (78) and (79) define respectively functions σ, a2 and λ, which satisfy the assumptions of Lemma 13. Proof. See Appendix J. 15

6

Construction of continuous-time consistent martingales

Since the transition distributions of the process (Xi ) are those of continuous-time processes sampled at random times, one could naturally imagine that it is possible to construct a non-trivial continuous-time process, with the same transition distributions, using a proper time change. The purpose of this section is to make this intuition more precise. We assume that: • µ0 ≡ δ (x − x0 ) with x0 ∈ R∗+ , • µ1 (resp. µ2 ) admits a positive and continuous density f1 (resp. f2 ) and has a first moment equal to x0 , • For x > 0, +

(x0 − x) <

Z

+∞ 0

+

(y − x) f1 (y) dy <

Z

0

+∞

(y − x)+ f2 (y) dy.

(80)

Using Sections 4 or 5, we know  sufficient conditions on f1 and f2 , for the existence of positive i strong Markov martingales Ft i∈{1,2} , such that the process (Xi ) defined by: X0 ≡ x0 ,   X1 | X0 ≡ Fτ11,X F01 = X0 , 0   2 X2 | X1 ≡ Fτ2,X F02 = X1 , 1

(81)

(82) (83)

where (τ1,x ) and (τ2,x ) are two families of exponentially distributed random variables of mean 1, independent respectively of Ft1 and Ft2 , is a martingale, which is calibrated to the distributions (µi )06i62 . Consequently, if we want to construct a continuous-time martingale (Xtc ), which starts at x0 in T0 ≡ 0 and has marginal density f1 (resp. f2 ) at time T1 > 0 (resp. T2 > T1 ), then a possibility is: XTc 0 ≡x0 , Xtc ≡ FT1 1 Xtc ≡ 

 1 F = Xc for t ∈ (T0 , T1 ] , 0 0 t−T0 ,X c 0 ! 2 FT2 2 for t > T1 , F0 = XTc 1 c t−T1 ,X T

(84)

1

 i where Tt,x is an adapted and increasing process, independent of Fti , starting at 0, and, whose marginal distribution at time Ti − Ti−1 is an exponential distribution of parameter 1, for  i ∈ {1, 2}. Note that the process (Xtc ) is indeed a martingale, because the processes Fti i∈{1,2} are  martingales and the time changes Tti i∈{1,2} are independent. Furthermore, by construction, law XTc i XTc i−1 = Xi | Xi−1 ,

(85)

for i ∈ {1, 2}. Consequently, the fact that (Xi ) is calibrated to the marginal distributions (µi ) implies that (Xtc ) is too.

16

 i A simple candidate for Tt,x is a Gamma process (Γtµi ,νi ), independent of all the other processes. We remind you that (Γtµi ,νi ) is an increasing pure-jump L´evy process, whose L´evy measure is defined by:   µ2i exp − µνii x kµi ,νi (x) dx = 1{x>0} dx, (86) νi x where µi (resp. νi ) is the mean (resp. variance) rate. Its marginal density at time t is a Gamma distribution of mean µi t and variance γi t. Since an exponential distribution is a special case of a Gamma distribution, we just need to chose the parameters according to our needs, that is: 1 . Ti − Ti−1

(87)

t τi,x . Ti − Ti−1

(88)

µ i = νi =

 i A second simple candidate for Tt,x is a linear interpolation of the family of random times (τi,x ): i Tt,x ≡

For more elaborated constructions of increasing processes - or time changes - with given marginal distributions, we refer to [15]. Finally, notice that the process (Xtc ) is a priori not Markov but the process (Xi ) is. As a  c consequence, the valuation of contingent claims depending only on XTi is much simpler, as explained in the following section. Remark 15 In the case where there is only one maturity, this section, along with Section 4, explains how to time-change a driftless diffusion to match a given marginal distribution at a given time. Consequently, using the classical Dambis-Dunbins-Schwarz theorem, we know how to sample a Brownian motion to match a given marginal distribution - what is exactly the purpose of the Skorokhod embedding problem (see [31]). We refer to [10] for a more detailed description of this solution. See also the recent working paper [16] on the same subject.

7

Valuation of path-independent claims

In this section, we focus on the valuation of path-independent claims on (Xi ), using ordinary (integro-) differential equations (O(I)DEs). Actually, for the sake of simplicity, we only describe the valuation of European style claims, but a similar treatment is possible for American style ones.  With the notations of the previous section, we denote by Gi the generator of the process Fti , for i ∈ {1, 2}: Z +∞  1 2 2 f (y) − f (x) − f ′ (x) (y − x) γi (x, y) dy. (89) x σi (x) f ′′ (x) + λi (x) Gi f (x) ≡ 2 0

Then, if Vif denotes the prices, at time Ti , of a claim expiring at time T2 and, whose payoff is a given function f , we have: Vif (x) ≡ E [f (X2 )| Xi = x] , (90)

and the purpose of this section is to show that V0f can be computed by solving a finite number of O(I)DEs. First, V2f clearly satisfies: V2f (x) = f (x). Second, by conditioning, we have: i h f (91) Vi−1 (x) = E [ E [ f (X2 )| Xi ]| Xi−1 = x] = E Vif (Xi ) Xi−1 = x . 17

Therefore, applying the backward operator [I − Gi ] on the above equation, we obtain, heuristically, that: f [I − Gi ] Vi−1 (x) = Vif (x) , (92) using Equation (12). f , if the transition distributions are of the general Note that Equation (92) is a linear ODE in Vi−1 Laplace type: f ∂ 2 Vi−1 1 2 2 f Vi−1 (x) − x σi (x) (x) = Vif (x) , (93) 2 ∂x2 but an OIDE if the transition distributions are general SJD distributions: Vif

(x) =

f Vi−1 (x) −

−λi (x)

Z

0

f ∂ 2 Vi−1 1 2 2 x σi (x) (x) 2 ∂x2

+∞

f f Vi−1 (y) − Vi−1 (x) −

(94) !

f ∂Vi−1 (x) (y − x) γi (x, y) dy. ∂x

However, if the jump-densities γi are of the generalized Laplace type and are associated to a function a2i , as in Section 5.4, then this OIDE can be transformed into an ODE. Indeed, since these distributions are centered, Equation (94) can be transformed into: Z

∂2V f

f V f (x) − Vif (x) − 12 x2 σi2 (x) ∂xi−1 (x) + λi (x) Vi−1 (x) 2 f (y)γi (x, y) dy = i−1 Vi−1 , λi (x) 0 i h a2 (x) ∂ 2 on the above equation yields: if λi > 0. And applying the operator I − i 2 ∂x 2 +∞

    ∂2V f f f f (x) a2i (x) ∂ 2 Vi−1 a2i (x) ∂ 2  Vi−1 (x) − Vi (x) − 12 x2 σi2 (x) ∂xi−1 2 , (x) = I − 2 ∂x2 2 ∂x2 λi (x)

(95)

(96)

i h a2 (x) ∂ 2 after a simple simplification, since I − i 2 ∂x 2 γi (x, y) = δ (x − y). Consequently, the valuation of European path-independent claims can be performed, at least heuristically, by solving, backward in time, a finite number of inhomogeneous ODEs (resp. OIDEs) if the transition distributions are of the generalized Laplace (resp. SJD) type. Moreover, to use generalized Laplace distributions as jump densities of general SJD distributions not only simplifies the calibration procedure, as we saw in Section 5.4, but also the valuation of claims, as the general valuation OIDE of Equation (94) was transformed, in this special case, into the simpler ODE of Equation (96).

8

A numerical example: SPX quotes

In this section, we give an example of calibration to S&P 500 option quotes, using the model developed in Section 4, where the transition distributions are of the generalized Laplace type. The market data are SPX quotes as of the close on September 15, 2005. To construct marginal distributions, which are (mostly) consistent with these quotes, we fit the different volatility smiles using the Stochastic Volatility Inspired (SVI) parametrization described in [21], p. 37, and evoked

18

in Section 4. The SVI fits to the SPX quotes3 are displayed in Figure 1. The corresponding marginal distributions (in the forward measure) are displayed in Figure 2.

09/16/05 0.6 0.5 0.4 0.3 0.2 0.1 0 0.24 0.2 0.16 0.12 0.08 0.2 0.175 0.15 0.125 0.1 0.2 0.175 0.15 0.125 0.1

10/21/05 0.24 0.18 0.12

1150

1200 11/18/05

0.06

1250

0.24 0.2 0.16 0.12 0.08

1100 1150 1200 1250 1300 1350 03/17/06

0.2 0.175 0.15 0.125 0.1

1100 1150 1200 1250 1300 1350 12/15/06

0.2 0.175 0.15 0.125 0.1

1100 1200 1300 1400 1500 1600 SVI

Bid

1100 1150 1200 1250 1300 1350 12/16/05

1100

1100

1200 1300 06/16/06

1200 1300 06/15/07

1400

1400

1100 1200 1300 1400 1500 1600 Ask

Figure 1: SVI fits to SPX quotes as of the close on September 15, 2005. Now that we have at our disposal marginal distributions, we need to ensure that they satisfy the sufficient conditions of Proposition 2. First, the cumulated implied variances are increasing with maturity. As a consequence, the corresponding call price functions are increasing with maturity for any positive strike. Second, an expansion of the normal cumulative distribution function allows us to show that all these marginal distributions have first moments equal to the index price. And third, the Matuszewska indices of these distributions are known as a function of the SVI parameters (see Section 4) and a simple check shows that our parametrizations lead to densities, which satisfy the assumptions of Lemma 3. Therefore, Proposition 2 applies and valid transition distributions between the maturities Ti−1 and Ti are generalized Laplace distributions associated to the function σi2 defined by: Ci (x) − Ci−1 (x) σi2 (x) = . (97) 1 2 ∂ 2 Ci 2 x ∂x2 (x) Notice that we did not really pay attention to the time dimension so far. Indeed, to ease notations, we decided to sample (jump-) diffusions at exponential random times of mean 1, but 3

We would like to thank Jim Gatheral for communicating the SPX quotes and the corresponding SVI fits he used in his book (see [21]).

19

0.013 T1 T2 T3 T4 T5 T6 T7 T8

0.012 0.011 0.01 0.009 0.008 0.007 0.006 0.005 0.004 0.003 0.002 0.001 0

800

900

1000

1100

1200 1300 Spot

1400

1500

1600

1700

Figure 2: Marginal densities. this mean is arbitrary and can be set to Ti −Ti−1 . In this case, the only change is that the associated generator Gi (see Equation (3)) should be replaced by Gi ≡ Gi / (Ti − Ti−1 ). These considerations lead us to consider the following rescaled quantities: σi (x) ≡ √

σi (x) . Ti − Ti−1

(98)

0.55 T1−>T2 T2−>T3 T3−>T4 T4−>T5 T5−>T6 T6−>T7 T7−>T8

0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0

250

500

750

1000

1250 Spot

1500

1750

2000

2250

2500

Figure 3: The functions σi . Since we have analytical expressions for the call price and for the density at every maturities, computing the above functions σ i is straightforward. These functions are displayed in Figure 3 between the 10−5 quantiles of the Ti density. Note that this choice is only esthetic and that these functions can be computed until much smaller quantiles without facing any numerical instability. 20

0.015 T1 −> T2 T2 −> T3 T3 −> T4 T4 −> T5 T5 −> T6 T6 −> T7 T7 −> T8

0.0125

0.01

0.0075

0.005

0.0025

0 900

950

1000 1050 1100 1150 1200 1250 1300 1350 1400 1450 1500 1550 1600 Spot

Figure 4: At-the-money forward transition densities. A first reassuring remark is that the functions σ i are of the same order of magnitude as the SPX implied volatilities and look like admissible local volatility functions. To have an idea of the behavior of the model, we draw the at-the-money forward transition distributions4 in Figure 4. These transitions densities have been obtained by using the definition of generalized Laplace distributions as the Green’s functions of a differential operator and the algorithm described in [32] to numerically solve linear second-order boundary value problem on infinite intervals. 0.28 T1−>T2 T2−>T3 T3−>T4 T4−>T5 T5−>T6 T6−>T7 T7−>T8

0.26 0.24 0.22 0.2 0.18 0.16 0.14 0.12 0.1 0.08 1050

1100

1150

1200

1250

1300

1350

1400

Strike

Figure 5: At-the-money forward smiles. The corresponding forward smiles are displayed in Figure 5. Note that the smiles correspond4 Note that these transition densities do not need to have a particular shape since generalized Laplace densities are structurally able to match a very large class of densities – see Section 4. In particular, the shape of the transition distribution from T1 to T2 displayed in Figure 4 is not surprising since the function σ2 decreases a lot at the right of the forward.

21

ing to the same time period are close to each other (and look like the initial smiles). This is a consequence of the fact that the corresponding functions σ i are similar around the index price.

1800 1700 1600

Spot

1500 1400 1300 1200 1100 1000 900 0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

Time

Figure 6: Different paths of the continuous-time process. Finally, we draw some paths of the continuous-time process, with the same transition distributions, obtained by linear interpolation of independent sampling times (see Section 6). The paths are displayed in Figure 6 and the corresponding time changes in Figure 7. 3.5

3

Time change

2.5

2

1.5

1

0.5

0

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

Time

Figure 7: Time changes corresponding to the paths of Figure 6.

9

Conclusion

In this paper, we described several constructions of discrete-time martingales consistent with a finite set of marginal distributions. Moreover, we explained how to extend these constructions to 22

continuous-time martingales (with a solution to the Skorokhod embedding problem as a by-product – see Remark 15). The advantages of these constructions over those of [17], [18], [2] and [11] are three-fold. First, once the marginal distributions at traded maturities are known, there is no need to interpolate them in time to get the whole implied volatility surface. One could actually see our constructions as a way of interpolating in an arbitrage-free and robust way given implied volatility smiles. Second, the constructed models are local stochastic volatility models with jumps and, to the best of our knowledge, are the first models of this kind to be structurally able to fit given implied volatility smiles. Third, as explained in Section 6, there is some room in the specification of the time changes when constructing the continuous-time martingales. This freedom should be studied further but could potentially be exploited to calibrate volatility products, when available. For instance, a lead that we did not pursue here is to introduce some correlation between the time change and the underlying process by making the mean of the exponentially distributed time change depend on the initial value of the corresponding (jump-) diffusion process. Finally, we focused in this paper on sampling (jump-) diffusions but the same recipe can be applied to other processes. For instance, to obtain transition distributions, which are the Green’s functions of fourth-order differential operators, one could consider a driftless diffusion with a Brownian time change (see Burdzy (1993), Allouba and Zheng (2001) and Allouba (2002) and the references therein). Another application is the construction of increasing processes calibrated to (a finite number of) marginal distributions described in [15] – where the techniques described in this paper and in [11] are also applied to the construction of increasing processes.

23

Appendix A

Reminders about generalized Laplace distributions

The Laplace distribution, whose density is defined, for λ > 0, by:   1 |x| pλ (x) ≡ exp − , (99) 2λ λ √  2λWτ , where (Wt ) is a Brownian motion and τ is an can be interpreted as the distribution of independent random time with an exponential distribution of mean 1. To define generalized Laplace distributions, we sample a general driftless diffusion (St ), satisfying: dSt = a (St ) dWt ,

(100)

instead of the multiple of a Brownian motion. The lemma below, along with its proof, can be found in [11] and in [15] and are recalled here for the sake of completeness. Lemma 16 If a2 is a continuous and positive function on R∗+ , which satisfies:  Z 1 Z 1 2 J0 ≡ dy dx = +∞, 2 0 x a (y) then, for every y ∈ R∗+ , there exists a unique function x → G (x, y), which satisfies:   a2 (x) ∂ 2 a2 (y) δ (x − y) , I− G (x, y) = 2 ∂x2 2

(101)

(102)

as well as the following boundary conditions: ∂G (x, y) → 0. x→0 or +∞ ∂x

(103)

Moreover, x → G (x, y) is positive and continuous on R∗+ as well as convex, twice differentiable and increasing (resp. decreasing) on (0, y] (resp. [y, +∞)). It also possesses the following symmetry: G (x, y) = G (y, x) .

(104)

, then y → g (x, y) is a probability density and g Furthermore, if we define g (x, y) ≡ 2G(x,y) a2 (y) satisfies:   ∂ 2 a2 (y) g (x, y) = δ (y − x) , (105) I− 2 ∂y 2

as well as



 a2 (x) ∂ 2 I− g (x, y) = δ (x − y) . 2 ∂x2

(106)

Finally, if we further assume that:

1

x dx = +∞, 2 0 a (x) Z +∞ x dx ≡ = +∞, 2 (x) a 1

K0 ≡ K+∞

Z

24

(107) (108)

then: G (x, y) x



x→0 or +∞

0,

∂G (x, y) → 0, x→0 or +∞ ∂x

(109) (110)

and y → g (x, y) is a probability density of mean x. Proof. First, since a2 is positive and continuous, we clearly have:  Z +∞ Z x dy J+∞ ≡ dx = +∞. 2 1 1 a (y)

(111)

Therefore, the results concerning G are mainly consequences of results, which can be found in [20] and [30]. Indeed, in this case, two solutions u1 and u2 of:   a2 (x) ∂ 2 f (x) = 0, (112) I− 2 ∂x2 exist and satisfy: u′1 (x) → 0,

(113)

x→0

u′2 (x)



x→+∞

0.

(114)

Moreover, u1 (resp. u2 ) is positive, convex and increasing (resp. decreasing). Besides, u1 and u2 are independent solutions and their Wronskian – denoted by W [ui ] – is a positive constant. Consequently, it allows us to define: G (x, y) ≡

u1 (min(x, y)) u2 (max(x, y)) , W [ui ]

(115)

and all the properties of G follow easily from this definition. Now let us focus on the properties of g. Equation (105) is a direct consequence of the definition of g and of Equation (102). Equation (106) is proved by using the definition of g – Equation (102) – and the symmetry property of G – Equation (104). To prove that g integrates to 1, we just need to integrate Equation (105) over R∗+ and use the boundary conditions of Equation (103). If we further assume Conditions (107) and (108), then: u1 (x) → 0,

(116)

x→0

u2 (x)



x→+∞

0,

(117)

and Equation (109) follows. To prove Equation (110), we obtain an estimate of u′2 in terms of u2 , using Theorem 1 in [24]. Indeed, u2 is positive, decreasing and satisfies Equation (112), therefore:   ′ xu2 (x) 6 8u2 x , (118) 2

for x big enough and the conclusion follows, since the equivalent result for u1 is a consequence of Equation (113). 25

Finally, to prove that the mean of y → g (x, y) is x, we multiply Equation (105) by y, integrate over R∗+ to obtain: Z Z ∂2G y 2 (x, y) dy = x, yg (x, y) dy − R∗+ ∂y R∗+ and we use the boundary conditions of Equations (109) and (110) to show that the second integral is zero by using two integrations by parts.

Appendix B

Proof of Proposition 2

The existence and the properties of the generalized Laplace densities (p (. |x1 )) are a simple consequence of the fact that a2 (x) ≡ x2 σ 2 (x) satisfies the assumptions of Lemma 16 in Appendix A. The only exception is Equation (26). Note that Equation (22) can be transformed into:   a2 (K) ∂ 2 C1 (K) = I − C2 (K) 2 ∂K 2 because f2 (K) =

∂ 2 C2 ∂K 2

(K). Differentiating this equation twice, we obtain:   ∂ 2 a2 (K) ∂ 2 C1 (K) = I − f2 (K) . ∂K 2 ∂K 2 2

Finally, we have: f2 (K) = ∂ since ∂K problem.



Z

0

a2 (K) 2

+∞

p (K |x1 )

∂ 2 C1 (x1 ) dx1 ∂x21

 f2 (K) →0 or +∞ 0 and using the uniqueness of the associated boundary value

Remark 17 In Proposition 2, the assumption ’σ 2 is bounded’ could have been replaced by the less constraining conditions:  Z 1 Z 1 f2 (y) dy dx = +∞, (119) 0 x C2 (y) − C1 (y) Z 1 xf2 (x) dx = +∞, (120) 0 C2 (x) − C1 (x) Z +∞ xf2 (x) dx = +∞, (121) C (x) − C1 (x) 2 1 and the above proof would have remained valid. However, we ensure with the assumptions of Proposition 2 that the generalized Laplace densities are those of a sampled diffusion, which is a martingale. And this will have its importance in Section 6 when we will explain how to construct continuous-time martingales with marginal distributions µ1 and µ2 .

Appendix C

Reminders about Karamata’s theory

For the sake of completeness, we state here some definitions and properties taken from [7] about Karamata’s theory. 26

Definition 18 Let f be a positive function defined in a neighborhood of +∞. Its upper Matuszewska index α (f ) is the infimum of those α for which there exists a constant C = C (α) such that for each Λ > 1, f (λx) 6 C (1 + o (1)) λα when x → +∞, uniformly in λ ∈ [1, Λ] ; f (x) Its lower Matuszewska index β (f ) is the supremum of those β for which, for some D = D (β) > 0 and all Λ > 1, f (λx) > D (1 + o (1)) λβ when x → +∞, uniformly in λ ∈ [1, Λ] . f (x) If α (f ) < +∞ (resp. β (f ) > −∞), we say that f has bounded increase (resp. decrease). A way to compute α (f ), β (f ) or to check if f has bounded increase (resp. decrease) is given by the following lemma, which is a simple rewriting of Theorem 2.1.5 and Corollary 2.1.6 in [7]. We will need the following notations: f (µx) , µ∈[1,λ] f (x)

Ψ (f, λ) ≡ lim sup sup x→+∞

f ∗ (λ) ≡ lim sup

f (λx) , f (x)

f∗ (λ) ≡ lim inf

f (λx) . f (x)

x→+∞

Ψ− (f, λ) ≡ Ψ (1/f, λ) ,

x→+∞

Lemma 19 Let f be a positive function defined in a neighborhood of +∞. We have the following equivalence: f has bounded increase ⇔ Ψ (f, .) < +∞

f has bounded decrease ⇔ Ψ− (f, .) < +∞ Moreover, if Ψ (f, .) < +∞ or Ψ− (f, .) < +∞, then     log (f ∗ (λ)) log (f ∗ (λ)) = inf , α (f ) = lim λ>1 λ→+∞ log (λ) log (λ) and β (f ) = lim

λ→+∞

Appendix D



log (f∗ (λ)) log (λ)



= inf

λ>1



log (f∗ (λ)) log (λ)



.

Proof of Lemma 3

Using the Potter-type bounds of Proposition 2.2.1. in [7], p. 72, ∃ε0 , A0 , M0 > 0, s.t. ∀y > x ≥ M0 , f2 (y) 6 A0 f2 (x)

 2+ε0 x . y

Consequently, σ 2 (x) 6

C2 (x) 1 2 2 x f2 (x)

=2

R +∞ R +∞ x

y

f2 (z) f2 (x) x2

 dz dy

6 2A0

27

R +∞ R +∞ x

y

 x 2+ε0 z

x2

 dz dy

=

2A0 . ε0 (1 + ε0 )

Likewise, using the Potter-type bounds, we have: ∃ε∞ , A∞ , M∞ > 0, s.t. ∀y > x > M∞ ,

Therefore, ∀y 6 x 6 1/M∞ ,

 y 1−ε∞ fe2 (y) 6 A∞ . x fe2 (x)

f2 (y) 6 A∞ f2 (x)

 1−ε∞ x . y

Consequently, if Pi denotes the put price function associated to the distribution µi , we have, using put-call parity: R x R y f2 (z)  0 f2 (x) dz dy 0 P2 (x) − P1 (x) A∞ P2 (x) σ 2 (x) = 6 . 6 = 2 1 2 1 2 2 x ε∞ (ε∞ + 1) 2 x f2 (x) 2 x f2 (x)

Appendix E

Proof of Proposition 5

The existence and the properties of (Ft ) are a consequence of Propositions 2.1 and 2.3 in [15] as discussed in Section 2. Since (Ft ) is a martingale, Equation (41) is satisfied. As far as Equation (42) is concerned, we have: ∂ 2 C1 (x1 ) = [I − G ∗ ] f2 (x1 ) , ∂x21 where G ∗ is the adjoint of G, by differentiating Equation (19) twice. Consequently, integrating by parts and using the limit conditions of Equations (38), (39) and (40), we obtain: Z

+∞ 0

Z

∂ 2 C1 p ( x2 | x1 ) (x1 ) dx1 = ∂x21

+∞

0

Z

=

0

p ( x2 | x1 ) (I − G ∗ ) f2 (x1 ) dx1

+∞

(I − G) p ( x2 | x1 ) f2 (x1 ) dx1 .

But remember that the probability distribution of (Ft ) also satisfies the backward Kolmogorov equation, which yields (I − G) p ( x2 | x1 ) = δ (x2 − x1 ) once the Carson-Laplace transform has been taken on both sides (see Equation (12)). Equation (42) follows.

Appendix F

Proof of Lemma 7

Let us first consider a positive, bounded and continuous function l. Then Z +∞ 2 ∂ C2 (y) l (y) Ψγ (y, x) dy ∂y 2 0 Z x 2 Z +∞ 2 ∂ C2 (y) ∂ C2 (y) = l (y) Ψγ (y, x) dy + l (y) Ψγ (y, x) dy 2 ∂y ∂y 2 0 x  Z Z x 2 +∞ ∂ C2 (y) (u − x)+ γ (y, u) du dy l (y) = 2 ∂y 0 0  Z +∞ Z +∞ 2 ∂ C2 (y) + + (x − u) γ (y, u) du dy, l (y) ∂y 2 x 0 28

since Ψγ (y, x) represents the price of an out-of-the-money option, struck at x, when the risk-neutral density is γ (y, .). Moreover,  Z +∞ Z x 2 Z +∞ 2 ∂ C2 (y) ∂ C2 (y) uγ (y, u) du dy + (sup l) P2,γ (x) l (y) Ψγ (y, x) dy 6 l (y) ∂y 2 ∂y 2 0 0 0 Z x 2 ∂ C2 (y) 6 A l (y) y dy + (sup l) P2,γ (x) ∂y 2 0 ≡ A0 + B0 , since the price of a call is always less than or equal to the forward price and l is bounded. Therefore, by taking l small enough in a neighborhood of 0, it is possible to obtain A0 =0 o (B0 ) and therefore: Z

0

+∞

∂ 2 C2 (y) l (y) Ψγ (y, x) dy =0 O (C2 (x) − C1 (x)) , ∂y 2

(122)

using Equation (45). R  +∞ Likewise, since 0 (x − u)+ γ (y, u) du 6 x, it is possible to take l small enough in a neighborhood of +∞, so that: Z +∞ 2 ∂ C2 (y) l (y) Ψγ (y, x) dy =+∞ O (C2 (x) − C1 (x)) , (123) ∂y 2 0 using Equation (46). R +∞ ∂ 2 C2 (y) We are now in a position to conclude. Indeed, C2 − C1 (resp. 0 l (y) Ψγ (y, .) dy) is ∂y 2 a continuous function on R∗+ and admits a minimum (resp. maximum) on any compact interval. Therefore, using Equations (122) and (123), λ = αl with α > 0 small enough satisfies Equation (47). The second part of the lemma is a simple consequence of the fact that: C2 (x) − C1 (x) −

R +∞

∂ 2 C2 (y) ∂y 2 λ (y) Ψγ 2 1 2 ∂ C2 2 x ∂x2 (x) 0

(y, x) dy 6

C2 (x) − C1 (x) 1 2 ∂ 2 C2 2 x ∂x2

(x)

and of Lemma 3.

Appendix G

Proof of Lemma 9

This proof has three main steps. Step 1: We show that the Fourier transforms of A and C are well-defined.  Ψ (z) First C ∈ L1 (R), since z → gz ∈ L1 R∗+ . Indeed,   if z < 1  1R Ψg (z)  0+∞ yg(y) dy if z ∈ (1, 2) 6 R +∞ z  αg z  g(y) dy y(ln(y))  1 if z > 2 αg z(ln(z))

29

,

using Equation (60). Likewise, C2 (z) − C1 (z) 6 C2 (z) − (X0 − z)+  R +∞ f (y) (z − y)+ R0+∞ 2 = f2 (y) (y − z)+  0   Rz +∞ yf2 (y) dy 6 R0   1+∞ f2 (y)y(ln(y))α2 dy (ln(z))α2

dy if z < X0 dy if z > X0 if z < X0 if z ∈ (X0 , max (X0 , 2)) if z > max (X0 , 2)

 1 (z) ∈ L1 R∗+ and, as a result, A ∈ L1 (R). using Equation (59). Therefore C2 (z)−C z Step 2: We show that  the assumptions of (A2 ), which only concern γ, are satisfied. 1 . First γ (x, .) ≡ x g x is clearly a probability density since g is one. Second, Z +∞ Z +∞ Z +∞ y  y  dy y ug (u) du γ (x, y) dy = g = x x x x 0 0 0

is indeed a bounded and continuous function since g has a finite first moment. Third, Z +∞   Z +∞ y y u |ln (u)| g (u) du < +∞ γ (x, y) dy = ln x x 0 0

using Equations (60) and (61). Finally, if Γ is a Borel set of R∗+ , let us consider a sequence (xn ), which goes to x as n goes to +∞. Then, we have:   Z y Z y ln xn ln x γ (x , y) dy γ (x, y) dy −    2 0 Γ 1 + ln y 2 y Γ 1 + ln x xn Z Z ln (u) ln (u) = g (u) du − g (u) du Γ 1 + (ln (u))2 Γ 1 + (ln (u))2 x xn Z |ln (u)| g (u) du 6 Γ Γ 1 + (ln (u))2 ∆x x n Z 1 6 g (u) du, 2 Γ∆ Γ x

Γ x

y

xn



s.t. y ∈ Γ and A∆B ≡ (A ∪ B) ∩ (A ∩ B). Finally, the facts that 1 Γ ∆ Γ (u) → 0 x xn  as n → +∞ and that g ∈ L1 R∗+ allows us to conclude that this integral is continuous using the dominated convergence theorem. Step 3: We show that Equation (19) is satisfied. Since B, FA/FC ∈ L1 (R), we have (see Theorem 8 in [8], p. 19):

where



x

FA = FBFC Using the theorems about the Fourier transform of a convolution (see Theorem 2 in [8], p. 6.) and the above theorem about inversion of the Fourier transform, we get: A = B ∗ C. Finally, reproducing backward the computations of the beginning of this section gives the desired conclusion. 30

Appendix H

Proof of Lemma 11

Thanks to Lemma 10, we only need to prove that Equation (19) is valid. This can be done by rewriting Equation (71), as follows:    ∂ 2 C2 ∂ 2 C2 a2 (x) 1 2 2 a2 (x) ∂ 2 x σ (x) λ (x) (x) = (x) . C (x) − C (x) − I− 2 1 2 ∂x2 2 ∂x2 2 ∂x2   2 Since C2 (x) − C1 (x) − 12 x2 σ 2 (x) ∂∂xC22 (x) →0 or +∞ 0, we have, by unicity of the associated value problem: Z +∞ 1 2 2 ∂ 2 C2 ∂ 2 C2 a2 (y) C2 (x) − C1 (x) − x σ (x) λ (y) (x) = (y) γ (x, y) dy 2 ∂x2 ∂x2 2 0 and the conclusion follows using Equation (64).

Proof of Lemma 12

Appendix I

The proof has five main steps. Step 1: We show that σ satisfies the assumptions of Lemma 11. The fact that σ is a positive, bounded and continuous function on R∗+ is a consequence of Lemma 3. Step 2: We introduce a candidate for the function λ. f1 f2 is bounded since it is a continuous function, bounded in a neighborhood of 0 and +∞. Consequently, if l is a positive, bounded and continuous function, such that:   ε + sup ff12 (x) (x) − 1 , l (x) > β2 where ε is a positive constant, then the function: A2l (x) ≡

1 (C2 (x) − C1 (x)) , 2 x (f2 (x) − f1 (x)) + β 2 l (x) f2 (x)

is a well-defined, positive and continuous function on R∗+ . Step 3: We show that A2l is bounded by two positive constants in a neighborhood of +∞. First, by reducing this neighborhood, we can assume that f2 (x) > M f1 (x). Therefore,   1 C2 (x) , C2 (x) > (C2 (x) − C1 (x)) > 1 − M and 2

2

(1 + β (sup l))f2 (x) > (f2 (x) − f1 (x)) + β l (x) f2 (x) > Consequently,



1 1− M

 1 1− M 1 C2 (x) 1 C2 (x) M ≤ A2l (x) 6 2 2 (1 + β (sup l)) x f2 (x) (M − 1) x2 f2 (x)



f2 (x) .

(x) Second x12 Cf22(x) is bounded by above in a neighborhood of +∞, since α (f2 ) < −2 (see the proof of Lemma 3 in Appendix D). Third, this quantity is also bounded by below by a positive

31

constant because f2 has bounded decrease. Indeed, using Proposition 2.2.1 in [7], p. 72, there exists ε∞ < −2, A∞ , M∞ > 0, s.t. ∀v > x > M∞ ,  v  ε∞ f2 (v) . > A∞ f2 (x) x

Therefore, integrating twice in v, we get:  Z +∞ Z +∞  ε A∞ v ∞ C2 (x) dv du = > A∞ x2 . f2 (x) x (ε + 1) (ε + 2) ∞ ∞ x u Step 4: We show that A2l is bounded by two positive constants in a neighborhood of 0. This step is actually left to the reader   as it is entirely similar to Step 3. As one could expect, its proof is based on the facts that α fe2 < 1 and that fe2 has bounded decrease. Step 5: We conclude. Since A2l is a positive, continuous function, which is bounded, by below and above, by positive constants in neighborhoods of 0 or +∞, there exist two positive constants ml , Ml > 0, such that: ml 6 A2l (x) 6 Ml . As a consequence, taking λ = l is sufficient.

Appendix J Since

f1 f2

Proof of Lemma 14

is continuous, bounded in a neighborhood of 0 and +∞, it is bounded on R∗+ . Consequently,

λ is bounded and continuous. Moreover, it is positive since α2 , β 2 ∈ (0, 1). Finally, bounded by two positive constants using the proof of Lemma 12 in Appendix I.

32

C2 (x)−C1 (x) x2 f2 (x)

is

References [1] Y. Achdou and O. Pironneau, Volatility smile by multilevel least square, International Journal of Theoretical and Applied Finance, 5 (2002), pp. 619–643. [2] L. Andersen and J. Andreasen, Jump-diffusion processes: Volatility smile fitting and numerical methods for option pricing, Review of Derivatives Research, 4 (2000), pp. 231–262. [3] J. Azema and M. Yor, Une solution simple au probleme de Skorokhod, in S´eminaire de Probabilit´es, XIII, vol. 721 of Lecture Notes in Math., Springer-Verlag, 1979, pp. 90–115. [4] S. Benaim and P. Friz, Smile asymptotics II: Models with known moment generating functions, J. Appl. Probab., 45 (2008), pp. 16–32. [5]

, Regular variation and smile asymptotics, Math. Finance, 19 (2009), pp. 1–12.

[6] K. Bichteler and J. Jacod, Calcul de Malliavin pour les diffusions avec sauts: existence d’une densit´e dans le cas unidimensionnel, S´eminaire de probabilit´es de Strasbourg, 17 (1983), pp. 132–157. [7] N. Bingham, C. Goldie, and J. Teugels, Regular Variation, Cambridge University Press, 1987. [8] S. Bochner and K. Chandrasekharan, Fourier transforms, Princeton University Press, 1949. [9] D. Breeden and R. Litzenberger, Prices of state-contingent claims implicit in options prices, J. Bus, 51 (1978), pp. 621–651. [10] P. Carr and L. Cousot, A diffusion-based solution to the Skorokhod embedding problem, work in progress, New York University, 2007. [11]

, A PDE approach to jump diffusions, preprint, New York University, 2007.

[12] P. Carr, H. Geman, D. Madan, and M. Yor, From local volatility to local L´evy models, Quant. Finance, 4 (2004), pp. 581–588. [13] T. Cass, Smooth densities for solutions to stochastic differential equations with jumps, Stochastic Process. Appl., 119 (2009), pp. 1416–1435. [14] R. Chacon and J. Walsh, One-dimensional potential embedding, in S´eminaire de Probabilit´es, X, vol. 511 of Lecture Notes in Math., Springer-Verlag, 1976, pp. 19–23. [15] L. Cousot, Constructions of martingales and of increasing processes with constrained marginal distributions., PhD thesis, New York University, 2008. ´ j, Time-homogeneous diffusions with a given marginal at [16] A. Cox, D. Hobson, and J. Oblo a random time, preprint, University of Bath, 2009. [17] E. Derman and I. Kani, Riding on a smile, Risk Magazine, 7 (1994), pp. 32–39. [18] B. Dupire, Pricing with a smile, Risk Magazine, 7 (1994), pp. 18–20. [19] S. Ethier and T. Kurtz, Markov processes: Characterization and convergence, Wiley Series in Probability and Statistics, John Wiley and Sons, second ed., 2005. 33

[20] W. Feller, The parabolic differential equations and the associated semi-groups of transformations, Ann. Math., 55 (1952), pp. 468–519. [21] J. Gatheral, The volatility surface: A practitioner’s guide, John Wiley and Sons, 2006. [22] J. Harrison and D. Kreps, Martingales and arbitrage in multiperiod securities markets, J. Econ. Theory, 20 (1979), pp. 381–408. [23] J. Harrison and S. Pliska, Martingales and stochastic integrals in the theory of continuous trading, Stochastic Process. Appl., 11 (1981), pp. 215–260. [24] P. Hartman, On the derivatives of solutions of linear, second-order, ordinary differential equations, Amer. J. Math., 75 (1953), pp. 173–177. [25] D. Hobson, The Skorokhod embedding problem and model-independent bounds for option prices, in Paris-Princeton Lectures on Mathematical Finance 2010, R. Carmona, E. Cinlar, I. Ekeland, E. Jouini, J. Scheinkman, and N. Touzi, eds., Lecture Notes in Math., SpringerVerlag, 2010. [26] L. Jiang, Q. Chen, L. Wang, and J. Zhang, A new well-posed algorithm to recover implied local volatility, Quant. Finance, 3 (2003), pp. 451–457. [27] T. Komatsu and A. Takeuchi, On the smoothness of pdf of solutions to SDE of jump type, Int. J. Differ. Equ. Appl., 2 (2001), pp. 141–197. [28] R. Lee, The moment formula for implied volatility at extreme strikes, Math. Finance, 14 (2004), pp. 469–480. ´min, Sur l’int´egrabilit´e uniforme des martingales exponentielles, Z. [29] D. Lepingle and J. Me Wahr. verw. Geb., 42 (1978), pp. 175–203. [30] H. McKean, Elementary solutions for certain parabolic partial differential equations, Trans. Amer. Math. Soc., 82 (1956), pp. 519–548. ´ j, The Skorokhod embedding problem and its offspring, Probab. Surv., 1 (2004), [31] J. Oblo pp. 321–392. [32] T. Robertson, The linear two-point boundary-value problem on an infinite interval, Math. Comp., 25 (1971), pp. 475–481. [33] V. Strassen, The existence of probability measures with given marginals, Ann. Math. Statist., 36 (1965), pp. 423–439. [34] D. Stroock, Diffusion processes associated with L´evy generators, Z. Wahr. verw. Geb., 32 (1975), pp. 209–244. [35] D. Stroock and S. Varadhan, Diffusion processes with continuous coefficients I, Comm. Pure and App. Math., 22 (1969), pp. 345–400. [36]

, Diffusion processes with continuous coefficients II, Comm. Pure and App. Math., 22 (1969), pp. 479–530.

[37] A. Takeuchi, The Malliavin calculus for SDE with jumps and the partially hypoelliptic problem, Osaka J. Math., 39 (2002), pp. 523–559.

34

Explicit constructions of martingales calibrated to given ...

prices, J. Bus, 51 (1978), pp. 621–651. [10] P. Carr ... prices, in Paris-Princeton Lectures on Mathematical Finance 2010, R. Carmona, E. Cinlar,. I. Ekeland, E.

811KB Sizes 2 Downloads 150 Views

Recommend Documents

Construction_Disabilities-Constructions-Renovations-Regulations.pdf
or elements, and changes or rearrangement in the plan configuration of walls and ... Redwood Trust, LLC, the defendant owner had recently remodeled and ...

Spatially explicit approach to estimation of total ...
examined an extensive data set of 1068 passerine birds in sub-Saharan ..... Instead, we apply an approximated pdf of the Thomas process ..... Encyclopedia of.

The Revenge of the Given
fluii they occupy is a philosophical enterprise within the meaning of the act, and that ..... То a first approximation, systems of representation are committed to .... All sorts of things follow: the phone books of big cities are generally bigger l

Comparative Constructions
18 Jun 2005 - (which some have claimed involves a type 〈1,1,1,1〉 quantifier operating on the four sets of students, attending-party, teachers, and attending-reception), would have a similar entry to (25) except that it would expect a complement w

Empirically Calibrated Model of Hepatitis C Virus Infection in the ...
year cumulative probability of cirrhosis to range from 6 to 71 percent in males ..... versus interferon alpha2b plus placebo for 48 weeks for treat- ment of chronic ...

Universal Constructions for Hopf Monoids
We are going to analyze under which conditions on a monoidally closed category C (being locally presentable as a category) the catgegory of Hopf monoids ...

Some Geometric Constructions
Dec 18, 2006 - Abstract. We solve some problems of geometric construction. Some of them cannot be solved with ruler and compass only and require the drawing of a rect- angular hyperbola: (i) construction of the Simson lines passing through a given po

Explicit Meaning Transmission
Agents develop individual, distinct meaning structures, ..... In Proceed- ings of the AISB Symposium: Starting from Society – the application of social analogies to ...

On the Complexity of Explicit Modal Logics
Specification (CS) for the logic L. Namely, for the logics LP(K), LP(D), LP(T ) .... We describe the algorithm in details for the case of LP(S4) = LP and then point out the .... Of course, if Γ ∩ ∆ = ∅ or ⊥ ∈ Γ then the counter-model in q

Some Constructions Related to the Kiepert Hyperbola
Dec 28, 2006 - SC + t. ) , t ∈ R ∪ {∞}, we study triangles with vertices on K having K as their own Kiepert hyperbolas. We shall work with homogeneous barycentric coordinates and make use of stan- dard notations of triangle geometry as in [2].

The Intellectual Given
May 26, 2015 - relevant doxastic attitudes, in the absence of putative reason to with- hold assent). To say that ..... Frey's strategy for proving Fermat's Last Theorem—could be salvaged by reverting to the. Horizontal ...... 59 Earlier drafts of t

THE NUMBER OF B3-SETS OF A GIVEN ...
Feb 7, 2014 - In view of the fact that Fh(n) = Θ(n1/h), one sees that c0hn1/h log ...... (55). It follows by (23) and Proposition 3.6 that. 360. eC. (A). eeCG0. (A).

Simple Constructions of the Incircle of an Arbelos
Sep 18, 2001 - In this note we give several simple constructions of the incircle of the arbelos. The elegant Construction 1 below was given by Leon Bankoff [2].

Integral representation of martingales and endogenous ...
May 17, 2011 - While the drift b = b(t, x) and the volatility σ = σ(t, x) coefficients for X need to have only minimal regularity properties with respect to x, they are ...

On the decompositions of T-quasi-martingales on Riesz spaces
On the decompositions of T-quasi-martingales on Riesz spaces. ∗. Jessica J. Vardy. Bruce A. Watson. †. School of Mathematics. University of the Witwatersrand. Private Bag 3, P O WITS 2050, South Africa. July 11, 2013. Abstract. The concept of a q

82 Year Old Community Beach Given To Adjacent Neighbors
sued King County and Seattle for ownership. By exploiting a ... domain: that is, acquisition of property by a government agency for the public good. I know this ...

Estimating the Yield Curve Using Calibrated Radial ...
Pohang University of Science and Technology. Pohang ... A (generalized) radial ba- sis function .... Lecture Notes in Computer Science, Springer-Verlag,.

pdf-1456\constructions-of-deviance-social-power-context-and ...
... apps below to open or edit this item. pdf-1456\constructions-of-deviance-social-power-context ... 7th-seventh-edition-by-peter-adler-patricia-a-adler.pdf.

The probabilistic structure of the distance between tributaries of given ...
May 16, 2007 - Indeed if at i sites one collects the areas Ai and must enforce the con- straint ...... See color version of this figure in the HTML. Figure 2. Sample ...

The comparative syntax of ditransitive constructions in Japanese ...
The present paper seeks to add a further perspective on this universalist-particularist ... on goal/theme ordering, (b) binding and reconstruction phenomena, ...