Skip to main content

Convex Order for Path-Dependent Derivatives: A Dynamic Programming Approach

  • Chapter
  • First Online:
Séminaire de Probabilités XLVIII

Part of the book series: Lecture Notes in Mathematics ((SEMPROBAB,volume 2168))

Abstract

We explore the functional convex order for various classes of martingales: Brownian or Lévy driven diffusions with respect to their diffusion coefficient, stochastic integrals with respect to their integrand. Each result is bordered by counterexamples. Our approach combines the propagation of convexity results through (simulable) discrete time recursive dynamics relying on a backward dynamic programming principle and powerful functional limit theorems to transfer the results to continuous time models. In a second part, we extend this approach to optimal stopping theory, namely to the réduites of adapted functionals of (jump) martingale diffusions. Applications to various types of bounds for the pricing of pathwise dependent European and American options in local volatility models are detailed. Doing so, earlier results are retrieved in a unified way and new ones are proved. This systematic paradigm provides tractable numerical methods preserving functional convex order which may be crucial for applications, especially in Finance.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    A càdlàg \((\mathcal{F}_{t})_{t\in [0,T]}\) -adapted process X = (X t ) t∈[0,T] is quasi-left continuous with respect to the right continuous filtration \(\mathcal{F} = (\mathcal{F}_{t})_{t\in [0,T]}\) if, for every \(\mathcal{F}\) -stopping time τ having values in [0,T] ∪{ +∞} and every increasing sequence of \(\mathcal{F}\) -stopping times (τ k ) k≥1 with limit τ, \(\lim _{k}X_{\tau _{k}} = X_{\tau }\) on the event {τ < +∞} (see e.g. [17, Chap. I.2.25, p. 22]).

  2. 2.

    i.e. satisfying \(\mathbb{E}\,X_{\tau _{n}^{{\ast}}}^{n} = u_{0}^{n}\).

References

  1. J. Bergenthum, L. Rüschendorf, Comparison of option prices in semi-martingales models. Finance Stoch. 10 (2), 222–249 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  2. J. Bergenthum, L. Rüschendorf, Comparison of semi-martingales and Lévy processes. Ann. Probab. 35, 228–254 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  3. J. Bergenthum, L. Rüschendorf, Comparison results for path-dependent options. Stat. Decis. 26, 53–72 (2008)

    MathSciNet  MATH  Google Scholar 

  4. P. Billingsley, Convergence of Probability Measures. Wiley Series in Probability and Statistics: Probability and Statistics, 2nd edn. (Wiley, New York, 1999), 277pp.

    Google Scholar 

  5. N. Bouleau, D. Lépingle, Numerical Methods for Stochastic Processes. Wiley Series in Probability and Mathematical Statistics: Applied Probability and Statistics (Wiley, New York, 1994), 359pp.

    MATH  Google Scholar 

  6. P. Carr, C.-O. Ewald, Y. Xiao, On the qualitative effect of volatility and duration on prices of asian options. Finance Res. Lett. 5, 162–171 (2008)

    Article  Google Scholar 

  7. D. Dacunha-Castelle, M. Duflo, Probabilités et Statistique II, problèmes à temps mobile (Masson, Paris, 1982), xiv+286pp.

    Google Scholar 

  8. N. El Karoui, Les aspects probabilistes du contrôle stochastique (French) [The probabilistic aspects of stochastic control], in 9 th Saint Flour Probability Summer School, 1979 (Saint Flour, 1979). Lecture Notes in Mathematics, vol. 876 (Springer, Berlin/Heidelberg, 1981), pp. 73–238

    Google Scholar 

  9. N. El Karoui, M. Jeanblanc, S. Shreve, Robustness of the Black-Scholes formula. Math. Finance 8 (2), 93–126 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  10. B. Hajek, Mean stochastic comparison of diffusions. Probab. Relat. Fields 68 (3), 315–329 (1985)

    MathSciNet  MATH  Google Scholar 

  11. O. Hernández-Lerma, W.J. Runggaldier, Monotone approximations for convex stochastic control problems. J. Math. Syst. Estimation Control 4 (4), 99–140 (1994)

    MathSciNet  MATH  Google Scholar 

  12. F. Hirsch, M. Yor, Comparing Brownian stochastic integrals for the convex order, in Modern Stochastics and Applications. Springer Optimization and Its Applications, vol. 90 (Springer, Cham, 2014), pp. 3–19

    Google Scholar 

  13. F. Hirsch, C. Profeta, B. Roynette, M. Yor, Peacocks and Associated Martingales, with Explicit Constructions (Bocconi, Milan; Springer, Milan, 2011), 430pp.

    Book  MATH  Google Scholar 

  14. D.G. Hobson, Robust hedging of the loopback option. Finance Stoch. 2 (4), 329–347 (1998)

    Article  MATH  Google Scholar 

  15. J. Jacod, The Euler scheme for Lévy driven stochastic differential equations: limit theorems. Ann. Probab. 32 (3), 1830–1872 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  16. J. Jacod, A.N. Shiryaev, Limit Theorems for Stochastic Processes. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 288 (Springer, Berlin, 1987), 601pp.

    Google Scholar 

  17. J. Jacod, A.N. Shiryaev, Limit Theorems for Stochastic Processes. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 288, 2nd edn. (Springer, Berlin, 2003), 661pp.

    Google Scholar 

  18. A. Jakubowski, J. Mémin, G. Pagès, Convergence en loi des suites d’intégrales stochastiques sur l’espace \(\mathbb{D}^{1}\) de Skorokhod (French) [Convergence in law of sequences of stochastic integrals on the Skorokhod space \(\mathbb{D}^{1}\)]. Probab. Theory Relat. Fields 81 (1), 111–137 (1989)

    Google Scholar 

  19. I. Karatzas, On the pricing of American options. Appl. Math. Optim. 17 (1), 37–60 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  20. I. Karatzas, S.E. Shreve, Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics, vol. 113, 2nd edn. (Springer, New York, 1991), 470pp.

    Google Scholar 

  21. H.G. Kellerer, Markov-Komposition und eine Anwendung auf Martingale (German). Math. Ann. 198, 99–122 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  22. P.E. Kloeden, E. Platen, Numerical Solution of Stochastic Differential Equations. Applications of Mathematics, vol. 23 (Springer, Berlin, 1992), xxxvi+632pp.

    Google Scholar 

  23. T. Kurtz, P. Protter, Weak limit theorems for stochastic integrals and stochastic differential equations. Ann. Probab. 19, 1035–1070 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  24. D. Lamberton, Optimal stopping and American options (2009), Ljubljana Summer School on Financial Mathematics, http://www.fmf.uni-lj.si/finmath09/ShortCourseAmericanOptions.pdf

  25. D. Lamberton, G. Pagès, Sur la convergence des réduites. Ann. de l’IHP, série B 26 (2), 331–35 (1990)

    Google Scholar 

  26. R. Lucchetti, Convexity and Well-Posed Problems. CMS Books in Mathematics (Springer, Berlin, 2006), 305pp.

    Google Scholar 

  27. J. Neveu, Martingales à temps discret (Masson, Paris, 1972), 218pp. English translation: Discrete-Parameter Martingales (North-Holland, New York, 1975), 236pp.

    Google Scholar 

  28. G. Pagès, Functional co-monotony of processes with an application to peacocks, in Séminaire de Probabilités XLV, ed. by C. Donati, A. Lejay, A. Rouault. Lecture Notes in Mathematics, vol. 2078 (Springer, Berlin, 2013), pp. 365–400

    Google Scholar 

  29. G. Pagès, Introduction to Numerical Probability and Applications to Finance, coll. Universitext, Springer (2014, to appear)

    Google Scholar 

  30. P. Protter, Stochastic Integration and Differential Equation. Stochastic Modeling and Applied Probability, vol. 21, 2nd edn. (3rd corrected printing) (Springer, New York, 2006), 419pp.

    Google Scholar 

  31. D. Revuz, M. Yor, Continuous Martingales and Brownian Motion. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 293, 3rd edn. (Springer, Berlin, 1999), 560pp.

    Google Scholar 

  32. K.I. Sato, Lévy Distributions and Infinitely Divisible Distributions. Cambridge Studies in Advanced Mathematics (Cambridge University Press, Cambridge, 1999), 486pp.

    Google Scholar 

  33. A.N. Shiryaev, Optimal Stopping Rules (Translated from the 1976 Russian 2nd edition by A.B. Aries). Stochastic Modeling and Applied Probability, vol. 8 (Springer, Berlin, 2008), xii+216pp. Reprint of the 1978 translation

    Google Scholar 

Download references

Acknowledgements

I am indebted to Marc Yor for helpful discussions and comments on a very preliminary version of this work. I also thank the referee for a careful reading of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gilles Pagès .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Euler Scheme for Brownian Martingale Diffusions

Proposition 10

Let \((\bar{X}_{t}^{n})_{t\in [0,T]}\) be the genuine Euler scheme of step \(\frac{T} {n}\) of the SDE ≡ dXt = σ(t,Xt)dWt, X0 = x defined as the solution to

$$\displaystyle{d\bar{X}_{t}^{n} =\sigma (\underline{t}_{ n},\bar{X}_{\underline{t}_{n}}^{n})dW_{ t},\;\bar{X}_{0}^{n} = x,\;t\! \in [0,T].}$$

If \(\sigma: [0,T] \times \mathbb{R} \rightarrow \mathbb{R}\) is continuous and satisfies the linear growth assumption

$$\displaystyle{\forall \,t\! \in [0,T],\;\forall \,x\! \in \mathbb{R},\quad \vert \sigma (t,x)\vert \leq C_{\sigma }(1 + \vert x\vert ),}$$

then the sequence \((\bar{X}^{n})_{n\geq 1}\) is C-tight on \(\mathcal{C}([0,T], \mathbb{R})\) and any of its limiting distributions is a weak solution to the above SDE. In particular if a weak uniqueness assumption holds, then \(\bar{X}^{n}\stackrel{\mathcal{L}(\|\,.\,\|_{\sup })}{\longrightarrow }X\).

Following e.g. [5] (Lemma B.1.2, p. 275, see also [22, 29]), we first show that, owing to the linear growth assumption, the non-decreasing function \(\varphi _{p,n}(t) = \mathbb{E}\big(\sup _{s\in [0,t]}\vert \bar{X}_{s}^{n}\vert ^{p}\big)\), p​ ∈ [1, +), is finite for every t​ ∈ [0, T]. Using Doob’s Inequality and Gronwall’s Lemma, it follows that

$$\displaystyle{\varphi _{p,n}(t) \leq \varphi _{p}(t):= Ce^{Ct}(1 + \vert x\vert ^{p})}$$

for a real constant C = C p, σ  > 0. Consequently, it follows from the L p-B.D.G. and Hölder inequalities, applied successively that, for every for p​ ∈ (2, +) and every s, t​ ∈ [0, T], s ≤ t,

$$\displaystyle{\mathbb{E}\vert \bar{X}_{t}^{n} -\bar{ X}_{ s}^{n}\vert ^{p} \leq c_{ p}^{p}\mathbb{E}\left (\int _{ s}^{t}\vert \sigma (\underline{u}_{ n},\bar{X}_{\underline{u}_{n}}^{n})\vert ^{2}du\right )^{\frac{p} {2} } \leq c_{p}^{p}\vert t - s\vert ^{\frac{p} {2} }\big(1 +\varphi _{p}(T)\big).}$$

Kolmogorov’s criterion (see [4, Theorem 12.3, p. 95]) implies that the sequence \(M_{n} = (W_{t},\bar{X}_{t}^{n})_{t\in [0,T]}\) is C-tight, i.e. tight on \((\mathcal{C}([0,T], \mathbb{R}^{2}),\|\,.\,\|_{\sup })\). From now on, we mainly rely on the results established in [18]. Let n′ be a subsequence such that \((\bar{X}^{n'},W)\) functionally weakly converges to a probability \(\mathbb{Q}\) on \((\mathcal{C}([0,T], \mathbb{R}^{2}),\|\,.\,\|_{\sup })\); hence it satisfies the U. T. (for Uniform Tightness) assumption (see Proposition 3.2 in [18]). The function σ being continuous on \([0,T] \times \mathbb{R}\), the sequence of càdlàg processes \((\sigma (\underline{t}_{n},\bar{X}_{\underline{t}_{n}}^{n}))_{n\geq 1}\) is C-tight on the Skorokhod space since \(\big((\underline{t}_{n},\bar{X}_{\underline{t}_{n}}^{n})_{t\in [0,T]}\big)_{n\geq 1}\) clearly is. One derives that, up to a new extraction still denoted (n′), we may assume that \(\big((\sigma (\underline{t}_{n'},\bar{X}_{\underline{t}_{n'}}^{n'}))_{t\in [0,T]},\bar{X}^{n'},W\big)_{n\geq 1}\) functionally converges toward a probability \(\mathbb{P}\) on \(\mathbb{D}([0,T], \mathbb{R}^{3})\). By Theorem 2.6 from [18]—the functional weak convergence of stochastic integrals theorem—we know that

$$\displaystyle{\left (\sigma (\underline{t}_{n'},\bar{X}_{\underline{t}_{n'}}^{n'}),(\bar{X}_{ t}^{n'},W_{ t}),\int _{0}^{t}\sigma (\underline{s}_{ n'},\bar{X}_{\underline{s}_{n'}}^{n'})dW_{ s}\right )_{t\in [0,T]}\stackrel{\mathcal{L}(Sk)}{\longrightarrow }\mathbb{Q}\quad \mbox{ as }\quad n \rightarrow +\infty }$$

where \(\mathbb{Q}\) is a probability distribution on \(\mathbb{D}([0,T], \mathbb{R}^{4})\) such that the canonical process Y = (Y i) i = 1: 4 satisfies \(Y \stackrel{\mathcal{L}}{\sim }\big(Y ^{1},(Y ^{2},B),\int _{0}^{.}Y _{s}^{2}dB_{s})\) where B: = Y 3 is a standard \(\mathbb{Q}\)-Brownian motion with respect to the \(\mathbb{Q}\)-completed right continuous canonical filtration \((\mathcal{D}_{t}^{4})_{t\in [0,T]}\) on \(\mathbb{D}([0,T], \mathbb{R}^{4})\). Furthermore, we know that Y 1 = σ(. , Y 2) \(\mathbb{Q}\)-a. s. since \(\sup _{t\in [0,T]}\vert \sigma (\underline{t}_{n'},\bar{X}_{\underline{t}_{n'}}^{n'}) -\sigma (t,\bar{X}_{t}^{n'})\vert \) converges to 0 in probability. The former claim follows from the facts that \(\sup _{t\in [0,T]}\vert \bar{X}_{t}^{n}\vert \) is tight and σ(t, ξ) is uniformly continuous on every compact set of \([0,T] \times \mathbb{R}\), with linear growth in ξ uniformly in t​ ∈ [0, T]. On the other hand, we know that \(\bar{X}^{n'} = x +\int _{ 0}^{.}\sigma (\underline{s}_{n'},\bar{X}_{\underline{s}_{n'}}^{n'})dW_{s}\), which in turn implies that Y .  2 = x + 0 .  σ(s, Y s 2)dW s . This shows the existence of a weak solution to the SDEX t  = x + 0 t σ(s, X s ) dW s , t​ ∈ [0, T].

Under the weak uniqueness assumption, this distribution is unique, hence is the only functional weak limiting distribution for the tight sequence \((\bar{X}^{n})_{n\geq 1}\). The convergence in distribution on \(\mathcal{C}([0,T], \mathbb{R})\) follows.

Remark 10

If the original SDE has a unique strong solution, the same proof leads to establish the convergence in probability of the Euler scheme toward X. One just has to add the process X itself to the sequence \(\big((\sigma (\underline{t}_{n},\bar{X}_{\underline{t}_{n}}^{n}))_{t\in [0,T]},\bar{X}^{n},W\big)_{n\geq 1}\).

Appendix 2: Euler Scheme for a Lévy Driven Martingale Diffusion

We consider the following SDE driven by a martingale Lévy process Z with Lévy measure ν:

$$\displaystyle{ X_{t} = x +\int _{(0,t]}\kappa (s,X_{s_{-}})dZ_{s},\;t\! \in [0,T],\;X_{0} = x, }$$
(33)

where κ is a Borel function on \([0,T] \times \mathbb{R}\). Its genuine Euler scheme is defined by

$$\displaystyle{ \bar{X}_{t_{k+1}}^{n} =\bar{ X}_{ t_{k}}^{n} +\ \kappa (t_{ k},\bar{X}_{t_{k}})(Z_{t_{k+1}} - Z_{t_{k}}),\;k = 1,\ldots,n,\;\bar{X}_{0} = X_{0} = x }$$
(34)

at discrete times t k n and extended into a continuous time càdlàg process by setting

$$\displaystyle{ \bar{X}_{t}^{n} = x +\int _{ (0,t]}\kappa (\underline{s}_{n-},\bar{X}_{\underline{s}_{n}-}^{n})dZ_{ s},\;t\! \in [0,T]. }$$
(35)

2.1 Convergence of the Euler Scheme Toward a Solution to the Lévy Driven SDE

Proposition 11

  1. (a)

    Let p​ ∈ (1,2]. Assume that ν(|z| p ) < +∞ and that Z has no Brownian component and κ(t,ξ) has linear growth in ξ, uniformly in t​ ∈ [0,T]. Then

    $$\displaystyle{\sup _{n\geq 1}\big\|\sup _{t\in [0,T]}\vert \bar{X}_{t}^{n}\vert \big\|_{ p} +\big\|\sup _{t\in [0,T]}\vert X_{t}\vert \big\|_{p} < +\infty.}$$

    If moreover κ is continuous, then SDE  (33) has at least one weak solution. Finally, under a weak uniqueness assumption, one has

    $$\displaystyle{\bar{X}^{n}\stackrel{\mathcal{L}(Sk)}{\longrightarrow }X.}$$
  2. (b)

    If ν(z 2 ) < +∞, the same result remains true mutatis mutandis if Z has a non-zero Brownian component.

Remark 11

In fact, if (33) has a strong solution, one shows using arguments similar to those developed below, the stronger result

$$\displaystyle{\sup _{t\in [0,T]}\vert \bar{X}_{t}^{n} - X_{ t}\vert \stackrel{\mathbb{P}}{\longrightarrow }0\;\mbox{ as }\;n \rightarrow +\infty.}$$

We refer to [15] (devoted to error bounds) for a simpler proof when κ is homogeneous and \(\mathcal{C}^{\ni }\) on the real line.

Proof

  1. (a)

    We consider the Lévy-Khintchine decomposition of the Lévy process Z = (Z t ) t ∈ [0, T], namely

    $$\displaystyle{Z_{t} =\widetilde{ Z}_{t} + Z^{1},\;t\! \in [0,T],}$$

    where \(\widetilde{Z}\) is a pure jump, square integrable martingale with jumps of size at most 1 and Lévy measure ν( . ∩{ | z | ≤ 1}) and Z 1 is a compensated (hence martingale) Poisson process with (finite) Lévy measure ν( . ∩{ | z |  > 1}).

It is clear from (34) that \(\bar{X}_{t_{k}^{n}}^{n}\! \in L^{p}\) for every k = 0, , n. Then, as ν( | z | p) < +, it follows classically that \(\sup _{u\in [t_{k}^{n},t_{k+1}^{n}]}\vert Z_{u}^{1} - Z_{t_{k}^{n}}^{1}\vert \stackrel{d}{\sim }\sup _{[0,\frac{T} {n} ]}\vert Z_{u}^{1}\vert \!\in L^{p}\) (see e.g. [32]). Combining these two results implies that \(\varphi _{p,n}(t):=\big\|\sup _{s\in [0,t]}\vert \bar{X}_{s}^{n}\vert \big\|_{p}\) is finite for every t​ ∈ [0, T].

It follows from Eq. (35) satisfied by \(\bar{X}\) that

$$\displaystyle{\varphi _{p,n}(t) \leq \vert x\vert +\Big\|\sup _{s\in [0,t]}\Big\vert \int _{(0,s]} \kappa (\underline{u}_{n-},\bar{X}_{\underline{u}_{n-}}^{n})dZ_{ u}\Big\vert \Big\|_{p}}$$

The L p-B.D.G. Inequality implies (since p > 1)

$$\displaystyle{\Big\|\sup _{s\in [0,t]}\Big\vert \int _{(0,s]}\kappa (\underline{u}_{n},\bar{X}_{\underline{u}_{n}-})dZ_{u}\Big\vert \Big\|_{p} \leq c_{p}\Big\|\sum _{0<s\leq t}\kappa (\underline{s}_{n},\bar{X}_{\underline{s}_{n-}})^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} }.}$$

Using that \(\frac{p} {2} \leq 1\), we derive

$$\displaystyle\begin{array}{rcl} \Big\|\sum _{0<s\leq t}\kappa (\underline{s}_{n},\bar{X}_{\underline{s}_{n-}})^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} }& \leq & \Big(\mathbb{E}\sum _{0<s\leq t}\vert \kappa (\underline{s}_{n},\bar{X}_{\underline{s}_{ n-}})\vert ^{p}\vert \varDelta Z_{ s}\vert ^{p}\Big)^{\frac{1} {p} } {}\\ & =& \Big(\nu (\vert z\vert ^{p})\,\mathbb{E}\!\int _{ 0}^{t}\vert \kappa (\underline{s}_{ n},\bar{X}_{\underline{s}_{n-}})\vert ^{p}ds\Big)^{\frac{1} {p} } {}\\ & \leq & C_{\kappa,p}^{p}\nu (\vert z\vert ^{p})^{\frac{1} {p} }\Big(\int _{0}^{t}(1 +\varphi _{p,n}(s)^{p})ds\Big)^{\frac{1} {p} } {}\\ \end{array}$$

where C κ, p is a real constant satisfying \(\vert \kappa (s,\xi )\vert \leq C_{\kappa,p}(1 + \vert \xi \vert ^{p})^{\frac{1} {p} }\), \((s,\xi )\! \in [0,T] \times \mathbb{R}\).

Finally, there exists a positive real constant C′ = C κ, p, ν such that the function φ p, n satisfies

$$\displaystyle{\varphi _{p,n}(t)^{p} \leq C'\Big(\vert x\vert ^{p} + t +\int _{ 0}^{t}\varphi _{ p,n}(s)^{p}ds\Big).}$$

One concludes by Gronwall’s Lemma that

$$\displaystyle{\forall \,t\! \in [0,T],\quad \varphi _{p,n}(t)^{p} \leq e^{C't}C'(T + \vert x\vert ^{p})}$$

or, equivalently, there exists a real constant C″ = C T, κ, p, ν such that

$$\displaystyle{\forall \,t\! \in [0,T],\quad \varphi _{p,n}(t) \leq \varphi _{p}(t) = e^{C''t}C''(1 + \vert x\vert ).}$$

To establish the Skorokhod tightness of the sequence \((\bar{X}^{n})_{n\geq 1}\), we rely on the Aldous tightness criterion (see Definition 3(b) or [17, Theorem 4.5, p. 356]). Let ρ​ ∈ (0, 1]. Let σ and τ be two [0, T]-valued \(\mathcal{F}^{Z}\)-stopping times such that σ ≤ τ ≤ (σ +δ) ∧ T.

$$\displaystyle\begin{array}{rcl} \mathbb{E}\vert \bar{X}_{\tau }^{n} -\bar{ X}_{\sigma }^{n}\vert ^{\rho } = \mathbb{E}\,\Big\vert \sum _{\sigma <u\leq \tau }\kappa (\underline{u}_{n},\bar{X}_{\underline{u}_{n-}}^{n})\varDelta Z_{ u}\Big\vert ^{\rho }& \leq & \mathbb{E}\Big(\sum _{\sigma <u\leq \tau }\vert \kappa (\underline{u}_{n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{\rho }\vert \varDelta Z_{ u}\vert ^{\rho }\Big) {}\\ & =& \nu (\vert z\vert ^{\rho })\mathbb{E}\int _{\sigma }^{(\sigma +\delta )\wedge T}\vert \kappa (\underline{u}_{ n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{\rho }du {}\\ & \leq & \delta \,\nu (\vert z\vert ^{\rho })\,\mathbb{E}\big[\sup _{t\in [0,T]}\vert \kappa (t,\bar{X}_{t}^{n})\vert ^{\rho }\big] {}\\ & \leq & \delta \,\nu (\vert z\vert ^{\rho })\,C_{\kappa }(1 +\varphi _{p}(T))^{ \frac{\rho }{ p} } {}\\ \end{array}$$

where we used that ρ ≤ 1 ≤ p and ν( | z | ρ) ≤ ν( | z | 2 ∧ 1) +ν( | z | p) < +. Then

$$\displaystyle\begin{array}{rcl} & & \sup \big\{\mathbb{E}\vert \bar{X}_{\tau }^{n} -\bar{ X}_{\sigma }^{n}\vert ^{\rho } + \mathbb{E}\vert Z_{\tau } - Z_{\sigma }\vert ^{\rho },\;\sigma \leq \tau \leq (\sigma +\delta ) \wedge T,\mathcal{F}^{Z}\mbox{ -stopping times}\big\} {}\\ & & \quad \leq \nu (\vert z\vert ^{\rho })(1 + C_{\kappa }(1 +\varphi _{p}(T))^{ \frac{\rho }{ p} }\big)\delta {}\\ \end{array}$$

which goes to 0 as δ → 0. This implies that the sequence \(M_{n} = (\bar{X}^{n},Z)\), n ≥ 1, is Sk-tight. Moreover, following Proposition 3.2 from [18], the sequence (M n ) n ≥ 1 satisfies the U. T. condition since it is Sk-tight and

$$\displaystyle\begin{array}{rcl} \mathbb{E}\sup _{t\in [0,T]}\big(\vert \varDelta \bar{X}_{t}^{n}\vert \vee \vert \varDelta Z_{ t}\vert \big)& \leq & \Big[\mathbb{E}\Big(\sum _{0<t\leq T}\vert \varDelta \bar{X}_{t}^{n}\vert ^{p} + \vert \varDelta Z_{ t}\vert ^{p}\Big)\Big]^{\frac{1} {p} } {}\\ & \leq & \Big[\nu (\vert z\vert ^{p})\mathbb{E}\int _{ 0}^{T}\big(1 + \vert \kappa (\underline{t}_{ n},\bar{X}_{\underline{t}_{n}}^{n})\vert ^{p}\big)dt\Big]^{\frac{1} {p} } {}\\ & \leq & \big(\nu (\vert z\vert ^{p})\big)^{\frac{1} {p} }\big(T + C_{\kappa,p}^{p}(1 +\varphi _{p}(T))\big)^{\frac{1} {p} } < +\infty. {}\\ \end{array}$$

On the other hand, the sequence \(\left ((\kappa \big(\underline{t}_{n},\bar{X}_{\underline{t}_{n}}^{n}))_{t\in [0,T]},M_{n}\right )_{n\geq 1}\) is Sk-tight, owing to the following lemma.

Lemma 6

Let \(\mathcal{V}_{[0,T]}^{+}\) be the set of functions μ: [0,T] → [0,T] such that μ(0) = 0 and μ(T) = T endowed with the sup norm. Assume \(\kappa: [0,T] \times \mathbb{R} \rightarrow \mathbb{R}\) is continuous. Then the mapping \(\varPsi: \mathcal{V}_{[0,T]}^{+} \times I\!\!D([0,T], \mathbb{R}^{d}) \rightarrow I\!\!D([0,T], \mathbb{R}^{1+d})\) defined by \(\varPsi (\mu,\alpha ) =\big (\kappa (\mu (.),\alpha ^{1}(.)),\alpha \big)\) is continuous (α = (α 1 ,…,α d )) for the product topology.

Proof (Proof of the Lemma)

Let (λ n ) n ≥ 1 be a sequence of increasing homeomorphisms of [0, T] such that λ n  → Id [0, T] and α n λ n  → α uniformly and let μ n  → μ in \(\mathcal{V}_{[0,T]}^{+}\) where Id [0, T] denotes the identity on [0, T]. Then the closure of (α n λ n (t)) n ≥ 1, t ∈ [0, T] is a compact set K of \(\mathbb{R}^{d}\) so that the function κ is uniformly continuous on [0, T] × K. On the other hand

$$\displaystyle{\|\mu _{n} \circ \lambda _{n} - Id_{[0,T]}\|_{\sup } \leq \|\mu _{n} - Id_{[0,T]}\|_{\sup } +\|\lambda _{n} - Id_{[0,T]}\|_{\sup }\quad \mbox{ as }\quad n \rightarrow +\infty }$$

and \(\|\alpha _{n} \circ \lambda _{n} -\alpha \|_{\sup }\rightarrow 0\) as n → +. The conclusion follows.

Up to an extraction, we may assume that the triplet \(\big(\big(\kappa (\underline{t}_{n'},\bar{X}_{\underline{t}_{n'}}^{n'}\big)_{t\in [0,T]},M_{n'}\big)_{n\geq 1}\) weakly converges for the Skorokhod topology toward a probability \(\mathbb{P}\) on the canonical Skorokhod space \((I\!\!D([0,T], \mathbb{R}^{3}),(\mathcal{D}_{t})_{t\in [0,T]})\).

By Theorem 2.6 from [18] for the functional convergence of stochastic integrals, we know that

$$\displaystyle{\Big(\kappa (\underline{t}_{n'},\bar{X}_{\underline{t}_{n'}}^{n'}),(\bar{X}_{ t}^{n'},Z_{ t}),\int _{0}^{t}\kappa (\underline{s}_{ n'-},\bar{X}_{\underline{s}_{n'-}}^{n'})dZ_{ s}\Big)_{t\in [0,T]}\stackrel{\mathcal{L}(Sk)}{\longrightarrow }\mathbb{Q}}$$

where \(\mathbb{Q}\) is a probability on \(\mathbb{D}([0,T], \mathbb{R}^{4})\) such that the canonical process Y = (Y i) i = 1: 4 satisfies \(Y \stackrel{d}{\sim }\big(Y ^{1},(Y ^{2},Y ^{3}),\int _{ 0}^{.}Y _{ s}^{2}dY _{ s}^{3})\) where Y 3 is a Lévy process with respect to the \(\mathbb{Q}\) and the \(\mathbb{Q}\)-completed right continuous canonical filtration \((\mathcal{D}_{t}^{\mathbb{Q}})_{t\in [0,T]}\) on \(\mathbb{D}([0,T], \mathbb{R}^{4})\) having the distribution of Z (i.e. \(\mathbb{Q}_{Y ^{3}} = \mathcal{L}(Z)\)). Furthermore, we know that Y 1 = κ(. , Y 2. ) \(\mathbb{Q}\)-a. s. since the mapping (μ, (α i) i = 1: 4) ↦ α 1κ(μ, α 2) is continuous from \(\mathcal{V}_{[0,T]}^{+} \times \mathbb{D}([0,T], \mathbb{R}^{4})\) to \(\mathbb{D}([0,T], \mathbb{R})\) (and t n converges uniformly to Id [0, T]).

On the other hand we know that \(\bar{X}_{t}^{n'} = x +\int _{ 0}^{t}\kappa (\underline{s}_{ n'-},\bar{X}_{\underline{s}_{n'-}}^{n'})dZ_{ s},\,t\! \in [0,T]\) which in turn implies that \((Y _{t}^{2} = x +\int _{ 0}^{t}\kappa (s,Y _{ s_{-}}^{2})dZ_{ s},\,t\! \in [0,T])\) \(\mathbb{Q}\)-a. s. . This shows the existence of a weak solution to the SDE \(X_{t} = x +\int _{ 0}^{t}\kappa (s,X_{s_{-}})dZ_{s},\,t\! \in [0,T]\).

Under the weak uniqueness assumption, the distribution \(\mathbb{Q}_{Y ^{2}}\) of Y 2 is unique equal, say, to \(\mathbb{P}_{X}\).

  1. (b)

    We assume that the Lévy measure has a finite second moment ν(z 2) < + on the whole real line. Then one can decompose Z as

    $$\displaystyle{Z_{t} = a\,W_{t} +\widetilde{ Z}_{t},\;t\! \in [0,T],\quad (a \geq 0)}$$

    where a ≥ 0 and \(\widetilde{Z}\) is a pure jump martingale Lévy process with Lévy measure ν. Then one shows like in the Brownian case that \(\varphi (t) = \mathbb{E}\big(\sup _{s\in [0,t]}\vert \bar{X}_{s}^{n}\vert ^{2}\big)\) is finite over [0, T] using that all \(\bar{X}_{t_{k}}\) are square integrable and \(\mathbb{E}\big(\sup _{s\in [t_{k},t_{k+1})}\vert Z_{s} - Z_{t_{k}}\vert ^{2}\big) = \mathbb{E}\big(\sup _{s\in [0,\frac{T} {n} ]}\vert Z_{s}\vert ^{2}\big) < +\infty \). Then, using Doob’s Inequality, we show that

    $$\displaystyle{\varphi (t) \leq 4C_{\kappa }^{2}(a^{2} +\nu (z^{2})\big)\Big(t +\int _{ 0}^{t}\varphi (s)ds\Big)}$$

    where C κ is a real constant satisfying \(\kappa (t,\xi ) \leq C_{\kappa }(1 + \vert \xi \vert ^{2})^{\frac{1} {2} }\), \(\xi \!\in \mathbb{R}\).

To establish the Skorokhod tightness of the sequence, we rely again on Aldous’ tightness criterion (see Definition 3(b) or[17, Theorem 4.5, p. 356]). Let σ, τ be two [0, T]-valued \(\mathcal{F}^{Z}\)-stopping times such that σ ≤ τ ≤ (σ +δ) ∧ T. Applying Doob’s Inequality, to the martingale \(\Big(\int _{\sigma }^{\sigma +s}\kappa (\underline{u}_{n},\bar{X}_{\underline{u}_{n-}})dZ_{u}\Big)_{s\geq 0}\) yields

$$\displaystyle\begin{array}{rcl} \mathbb{E}\vert \bar{X}_{\tau }^{n} -\bar{ X}_{\sigma }^{n}\vert ^{2}& \leq & 4a^{2}\mathbb{E}\left (\int _{\sigma }^{\tau }\vert \kappa (\underline{u}_{ n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{2}du\right ) + 4\,\mathbb{E}\left (\sum _{\sigma <u\leq \tau }\vert \kappa (\underline{u}_{n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{2}\vert \varDelta Z_{ u}\vert ^{2}\right ) {}\\ & =& 4\big(a^{2} +\nu (z^{2})\big)\mathbb{E}\left (\int _{\sigma }^{\tau }\vert \kappa (\underline{u}_{ n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{2}du\right ) {}\\ & \leq & 4\big(a^{2} +\nu (z^{2})\big)\mathbb{E}\left (\int _{\sigma }^{(\sigma +\delta )\wedge T}\vert \kappa (\underline{u}_{ n},\bar{X}_{\underline{u}_{n-}}^{n})\vert ^{2}du\right ) {}\\ & \leq & 4(a^{2} +\nu (z^{2})\big)\delta \,C_{\kappa }^{2}(1 +\varphi (T)). {}\\ \end{array}$$

It follows that \(\mathbb{E}\vert \bar{X}_{\tau } -\bar{ X}_{\sigma }\vert ^{2} + \mathbb{E}\vert Z_{\tau } - Z_{\sigma }\vert ^{2} \leq 4\,C_{\kappa }^{2}(a^{2} +\nu (z^{2})\big)\nu (z^{2})(1 +\varphi (T)\big)\delta\) which clearly implies the Sk-tightness of the sequence \(M_{n} = (\bar{X}^{n},Z)\), n ≥ 1.

The sequence satisfies the U. T. condition from [18] since (M n ) n ≥ 1 is Sk-tight and (see Proposition 3.2 from [18])

$$\displaystyle\begin{array}{rcl} \mathbb{E}\Big[\sup _{t\in [0,T]}\big(\vert \varDelta \bar{X}_{t}^{n}\vert \vee \vert \varDelta Z_{ t}\vert \big)\Big]& \leq & \Big(\mathbb{E}\Big[\sum _{0<t\leq T}\vert \varDelta \bar{X}_{t}^{n}\vert ^{2} + \vert \varDelta Z_{ t}\vert ^{2}\Big]\Big)^{\frac{1} {2} } {}\\ & \leq & \Big(\nu (z^{2})\mathbb{E}\int _{ 0}^{T}\big(1 + \vert \kappa (\underline{t}_{ n},\bar{X}_{\underline{t}_{n}}^{n})\vert ^{2}\big)dt\Big)^{\frac{1} {2} } {}\\ & \leq & \Big(\nu (z^{2})(T + C_{\kappa }\big(1 +\varphi (T))\big)\Big)^{\frac{1} {2} } < +\infty. {}\\ \end{array}$$

From this point, the proof is similar to that of claim (a).

2.2 Higher Moments

Let \(Z_{t} = aW_{t} +\widetilde{ Z}_{t}\), t​ ∈ [0, T], be the decomposition of the Lévy process Z where W is a standard B.M. and \(\widetilde{Z}\) is a pure jump Lévy process independent of W.

Proposition 12

Let p​ ∈ [2,+∞). If ν(|z| p ) < +∞ then

$$\displaystyle{\sup _{n\geq 1}\Big\|\sup _{t\in [0,T]}\vert \bar{X}_{t}^{n}\vert \Big\|_{ p} < +\infty.}$$

Proof

If p​ ∈ (1, 2], the claim follows from the above Proposition 11. Assume from now on p​ ∈ [2, +). Let \(\varphi _{p,n}(t) = \mathbb{E}\big(\sup _{t\in [0,T]}\vert \bar{X}_{t}^{n}\vert ^{p}\big)\). Let p be the unique integer defined by the inequality \(2^{\ell_{p}} < p \leq 2^{\ell_{p}+1}\). It is straightforward, using the same arguments as above, that φ p, n (T) < + since sup t ∈ [0, T] | Z t  | p​ ∈ L 1 (see [32, Theorem 25.18, p. 166]) and \(X_{t_{k}}\! \in L^{p}\) by induction using (34). For convenience, we set \(\kappa _{s_{-}} =\kappa (\underline{s}_{n},\bar{X}_{\underline{s}_{n}-}^{n})\).

Now, combining the integral and the regular Minkowski Inequalities with the B.D.G. Inequality implies

$$\displaystyle\begin{array}{rcl} \varphi _{p,n}(t)^{\frac{1} {p} }& \leq & \vert x\vert + c_{p}\Big\|a^{2}\int _{0}^{t}\kappa _{s_{ -}}^{2}ds +\sum _{ 0<s\leq t}\kappa _{s_{-}}^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} } \\ & \leq & \vert x\vert + c_{p}\Big(a\Big\|\int _{0}^{t}\kappa _{ s_{-}}^{2}ds\Big\|_{\frac{p} {2} }^{ \frac{1} {2} } +\Big\|\sum _{0<s\leq t}\kappa _{s_{ -}}^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} }\Big){}\end{array}$$
(36)

where we used in the second inequality that \(\sqrt{ u + v} \leq \sqrt{u} + \sqrt{v}\), u, v ≥ 0. First note that by two successive applications of Hölder Inequality to dt and \(d\mathbb{P}\), we obtain

$$\displaystyle{ \Big\|\int _{0}^{t}\kappa _{ s_{-}}^{2}ds\Big\|_{\frac{p} {2} }^{ \frac{1} {2} } \leq T^{\frac{1} {2} -\frac{1} {p} }\Big(\int _{0}^{t}\mathbb{E}\,\vert \kappa _{s_{ -}}\vert ^{p}ds\Big)^{\frac{1} {p} }. }$$
(37)

Using that for every ​ ∈ { 1, ,  p }, \(\Big(\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2^{\ell}}\vert \varDelta Z_{ s}\vert ^{2^{\ell}} -\int _{ 0}^{t}\vert \kappa _{ s_{-}}\vert ^{2^{\ell}}ds\,\nu (\vert z\vert ^{2^{\ell}})\Big)_{ t\!\in [0,T]}\), is a true martingale, we have by combining this time the Minkowski inequality, the B.D.G. Inequality applied with \(\frac{p} {2^{\ell}} > 1\) and the elementary inequality (u + v)r ≤ u r + v r, u, v ≥ 0, r​ ∈ (0, 1] that:

$$\displaystyle\begin{array}{rcl} \Big\|\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2^{\ell}}(\varDelta Z_{ s})^{2^{\ell}}\Big\|_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} }& \leq & \Big\|\sum _{0<s\leq t}\vert \kappa _{s_{ -}}\vert ^{2^{\ell}}(\varDelta Z_{ s})^{2^{\ell}} -\int _{ 0}^{t}\vert \kappa _{ s_{-}}\vert ^{2^{\ell}}ds\,\nu (\vert z\vert ^{2^{\ell}})\Big\|_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} } {}\\ & & +\Big\|\int _{0}^{t}\vert \kappa _{ s_{-}}\vert ^{2^{\ell}}ds\Big\|_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} }\nu (\vert z\vert ^{2^{\ell} })^{\frac{1} {2^{\ell}} } {}\\ & \leq & c_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} }\Big\|\sum _{0<s\leq t}\vert \kappa _{s_{ -}}\vert ^{2^{\ell+1} }(\varDelta Z_{s})^{2^{\ell+1} }\Big\|_{ \frac{p} {2^{\ell+1}} }^{ \frac{1} {2^{\ell+1}} } {}\\ & & +\Big\|\int _{0}^{t}\vert \kappa _{ s_{-}}\vert ^{2^{\ell}}ds\Big\|_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} }\nu (\vert z\vert ^{2^{\ell} })^{\frac{1} {2^{\ell}} }. {}\\ \end{array}$$

Then two applications of Hölder Inequality applied to dt and \(d\mathbb{P}\) successively imply

$$\displaystyle{\Big\|\int _{0}^{t}\vert \kappa _{ s_{-}}\vert ^{2^{\ell}}ds\Big\|_{\frac{p} {2^{\ell}} }^{ \frac{1} {2^{\ell}} } \leq T^{\frac{1} {2^{\ell}} -\frac{1} {p} }\Big(\int _{0}^{t}\mathbb{E}\,\vert \kappa _{s_{ -}}\vert ^{p}ds\Big)^{\frac{1} {p} }.}$$

Summing up these inequalities in cascade finally yields a positive real constant K p, ν, a, T (0) such that

$$\displaystyle\begin{array}{rcl} \Big\|\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} }& \leq & K_{p,\nu,a,T}^{(0)}\Big(\Big(\int _{0}^{t}\mathbb{E}\,\vert \kappa _{s_{ -}}\vert ^{p}ds\Big)^{\frac{1} {p} } {}\\ & & \quad +\Big\|\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2^{\ell_{p}+1} }(\varDelta Z_{s})^{2^{\ell_{p}+1} }\Big\|_{ \frac{p} {2^{\ell_{p}+1}} }^{ \frac{1} {2^{\ell_{p}+1}} }\Big). {}\\ \end{array}$$

Now, as \(\frac{p} {2^{\ell_{p}+1}} \leq 1\), one gets by the compensation formula

$$\displaystyle\begin{array}{rcl} \Big\|\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2^{\ell_{p}+1} }\vert \varDelta Z_{s}\vert ^{2^{\ell_{p}+1} }\Big\|_{ \frac{p} {2^{\ell_{p}+1}} }^{ \frac{1} {2^{\ell_{p}+1}} }& \leq & \Big(\mathbb{E}\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{p}(\varDelta Z_{s})^{p}\Big)^{\frac{1} {p} } {}\\ & =& \Big(\int _{0}^{t}\mathbb{E}\vert \kappa _{ s_{-}}\vert ^{p}ds\Big)^{\frac{1} {p} }\nu (\vert z\vert ^{p})^{\frac{1} {p} }. {}\\ \end{array}$$

Hence, there exists a real constant K p, ν, a, T (1) > 0

$$\displaystyle{ \Big\|\sum _{0<s\leq t}\vert \kappa _{s_{-}}\vert ^{2}(\varDelta Z_{ s})^{2}\Big\|_{\frac{p} {2} }^{ \frac{1} {2} } \leq K_{p,\nu,a,T}^{(1)}\Big(\int _{0}^{t}\mathbb{E}\vert \kappa _{s_{ -}}\vert ^{p}ds\Big)^{\frac{1} {p} }. }$$
(38)

Finally, plugging (37) and (38) in (36), there exist positive real constants K p, ν, a, T (),  = 2, 3, such that

$$\displaystyle{\varphi _{p,n}(t)^{\frac{1} {p} } \leq K_{p,\nu,a,T}^{(2)}\Big(\vert x\vert +\Big (\int _{0}^{t}\mathbb{E}\vert \kappa _{s_{ -}}\vert ^{p}ds\Big)^{\frac{1} {p} }\Big) \leq K_{p,\nu,a,T}^{(3)}\Big(\vert x\vert + 1 +\Big (\int _{0}^{t}\varphi _{p,n}(s)ds\Big)^{\frac{1} {p} }\Big)}$$

where we used in the second inequality that κ has linear growth. Hence

$$\displaystyle{\varphi _{p,n}(t) \leq 2^{p-1}(K_{ p,\nu,a,T}^{'(3)})^{p}\Big(\big(\vert x\vert + 1\big)^{p} +\int _{ 0}^{t}\varphi _{ p,n}(s)ds\Big).}$$

Gronwall’s lemma completes the proof since it implies that

$$\displaystyle{\varphi _{p,n}(t) \leq e^{2^{p-1}(K_{ p,\nu,a,T}^{'(3)})^{p}\,t }2^{p-1}(K_{ p,\nu,a,T}^{'(3)})^{p}(\vert x\vert + 1)^{p}.}$$

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Pagès, G. (2016). Convex Order for Path-Dependent Derivatives: A Dynamic Programming Approach. In: Donati-Martin, C., Lejay, A., Rouault, A. (eds) Séminaire de Probabilités XLVIII. Lecture Notes in Mathematics(), vol 2168. Springer, Cham. https://doi.org/10.1007/978-3-319-44465-9_3

Download citation

Publish with us

Policies and ethics