Skip to main content

Three Essays on Exponential Hedging with Variable Exit Times

  • Chapter
Inspired by Finance

Abstract

This paper addresses three main problems that are intimately related to exponential hedging with variable exit times. The first problem consists of explicitly parameterizing the exponential forward performances and describing the optimal solution for the corresponding utility maximization problem. The second problem deals with the horizon-unbiased exponential hedging. Precisely, we are interested in describing the dynamic payoffs for which there exists an admissible strategy that minimizes the risk—in the exponential utility framework—whenever the investor exits the market at stopping times. Furthermore, we explicitly describe this optimal strategy when it exists. Our last contribution is concerned with the optimal sale problem, where the investor is looking simultaneously for the optimal portfolio and the optimal time to liquidate her assets.

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover 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

References

  1. Choulli, T., Schweizer, M.: The mathematical structure of horizon-dependence in optimal portfolio choice. Preprint (2009)

    Google Scholar 

  2. Choulli, T., Schweizer, M.: Stability of sigma-martingale densities in LlogL under an equivalent change of measure. Preprint (2010)

    Google Scholar 

  3. Choulli, T., Stricker, C.: Minimal entropy—Hellinger martingale measure in incomplete markets. Math. Finance 15(3), 465–490 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  4. Choulli, T., Stricker, C.: More on minimal entropy—Hellinger martingale measures. Math. Finance 16(1), 1–19 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  5. Choulli, T., Stricker, C., Li, J.: Minimal Hellinger martingale measures of order q. Finance Stoch. 11(3), 399–427 (2007)

    Article  MathSciNet  Google Scholar 

  6. Delbaen, F., Grandits, P., Rheinländer, T., Samperi, D., Schweizer, M., Stricker, C.: Exponential hedging and entropic penalties. Math. Finance 12(2), 99–123 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  7. Delbaen, F., Schachermayer, W.: The Mathematics of Arbitrage. Springer, Heidelberg (2006)

    MATH  Google Scholar 

  8. Dellacherie, C., Meyer, P.-A.: Théorie des Martingales. Hermann, Paris (1980). Chapter V to VIII

    MATH  Google Scholar 

  9. Evans, J., Henderson, V., Hobson, D.: Optimal timing for an indivisible asset sale. Math. Finance 18, 545–567 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  10. Fisher, I.: The impatience theory of interest. Am. Econ. Rev. 3, 610–618 (1931)

    Google Scholar 

  11. Hakansson, N.H.: Optimal investment and consumption strategies under risk, an uncertain lifetime and insurance. Int. Econ. Rev. 10, 443–466 (1969)

    Article  MATH  Google Scholar 

  12. Henderson, V.: Valuing the option to invest in an incomplete market. Math. Finance Econ. 1(2), 103–128 (2007)

    Article  Google Scholar 

  13. Henderson, V., Hobson, D.: Horizon-unbiased utility functions. Stoch. Process. Appl. 117(11), 1621–1641 (2006)

    Article  MathSciNet  Google Scholar 

  14. Jacod, J.: Calcul Stochastique et Problèmes de Martingales. Lecture Notes in Mathematics, vol. 714. Springer, Berlin (1979)

    MATH  Google Scholar 

  15. Jacod, J., Shiryaev, A.: Limit Theorems for Stochastic Processes, 2nd edn. Springer, Berlin (2002)

    Google Scholar 

  16. Kabanov, Y.: On the FTAP of Kreps–Delbaen–Schachermayer. Statistics and control of random processes. The Liptser Festschrift. In: Proceedings of Steklov Mathematical Institute Seminar, pp. 191–203. World Scientific, Singapore (1997)

    Google Scholar 

  17. Kabanov, Y., Stricker, C.: On the optimal portfolio for the exponential utility maximization: remarks to the six-author paper. Math. Finance Econ. 12(2), 125–134 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  18. Karatzas, I., Zamfirescu, M.: Martingale approach to stochastic control with discretionary stopping. Appl. Math. Optim. 53, 163–184 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Larsen, K., Hang, Y.: Horizon dependence of utility optimizers in incomplete models. Preprint (2010)

    Google Scholar 

  20. Lépingle, D., Mémin, J.: Sur l’intégrabilité uniforme des martingales exponentielles. Z. Wahrscheinlichkeitstheor. Verw Geb. 42, 175–203 (1978)

    Article  MATH  Google Scholar 

  21. Maingueneau, M.A.: Temps d’arrêt optimaux et théorie générale. Sémin. Probab. 12, 457–467 (1978)

    MathSciNet  Google Scholar 

  22. Mertens, J.F.: Processus stochastiques généraux et surmartingales. Z. Wahrscheinlichkeitstheor. Verw Geb. 22, 45–68 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  23. Musiela, M., Zariphopoulou, T.: Portfolio choice under dynamic investment performance criteria. Quant. Finance 9(2), 161–170 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  24. Musiela, M., Zariphopoulou, T.: Backward and forward utilities and the associated indifference pricing systems: the case study of the binomial model. In: Carmona, R. (ed.) Indifference Pricing, pp. 3–43. Princeton University Press, Princeton (2009)

    Google Scholar 

  25. Rockafellar, R.T.: Convex Analysis. Princeton University Press, Princeton (1970)

    MATH  Google Scholar 

  26. Yaari, M.E.: Uncertain lifetime, life insurance and the theory of the consumer. Rev. Econ. Stud. 32, 137–158 (1965)

    Article  Google Scholar 

  27. Zariphopoulou, T., Zitkovic, G.: Maturity-independent risk measures. SIAM J. Financ. Math. 1, 266–288 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  28. Zitkovic, G.: A dual characterization of self-generation and log-affine forward performances. Ann. Appl. Probab. 19(6), 2176–2270 (2009)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This research was supported financially by the Natural Sciences and Engineering Research Council of Canada via Choulli’s Grant G121210818.

The first and the second authors would like to thank anonymous referees for the careful reading and valuable inputs and suggestions. Both, the first and the second authors are grateful to Christoph Frei for his fruitful comments and advices. Any remaining error is of our responsibility.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tahir Choulli .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Some Auxiliary Lemmas

This section is contains six lemmas, which were used in previous sections. We note that some of these lemmas are interesting on their own right.

Lemma 5

Let Q be a σ-martingale measure for S, and θL(S) be such that

(40)

Then the process θS is a Q-local martingale and the process exp[θS] is a positive Q-submartingale.

Proof

Since Q is a σ-martingale measure for S, there exists a positive, bounded and predictable process ϕ such that ϕS is a Q-local martingale. As a result, θS is σ-martingale under Q. On the other hand, it is clear that

$$X_{t}:=\exp\left(\frac{1}{2}(\theta\cdot S)_{t}\right) $$

is a positive special semimartingale under Q with the Doob–Meyer decomposition

$$ X=X_{0}+{\overline{N}}+{\overline{B}} $$

where \({\overline{N}}\) is a Q-local martingale and \({\overline{B}}\) is a predictable process with finite variation such that \({\overline{N}}_{0}={\overline{B}}_{0}=0\). Let (T n ) n≥1 be a sequence of stopping times that increases stationarily to T and

$$ E^Q\left([{\overline{N}},{\overline{N}}]^{1/2}_{T_{n}}+{\rm Var}_{T_{n}}{\overline{B}}\right)<\infty. $$

For any predictable process φ such that |φ|≤1, we have:

$$ E^Q|\varphi\cdot X_{T_{n}}|\leq cE^Q\left([{\overline{N}},{\overline{N}}]^{1/2}_{T_{n}}+{\rm Var}_{T_{n}}({\overline{B}})\right), $$
(41)

where c is a constant that does not depend on φ.

Using Ito’s formula, we obtain

$$X=1+\frac{1}{2}X_{-}\cdot(\theta\cdot S)+X_{-}\cdot V(\theta), $$

where V(θ) is a non-decreasing process given by

$$ V(\theta):=\frac{1}{8}\theta^{T}c\theta\cdot A+\left(\exp\left(\frac{1}{2}\theta^{T}x\right)-1-\frac{1}{2}\theta ^{T}x\right)\star\mu. $$

Since θS is a σ-martingale under Q, then there exists a predictable process ϕ with values in the interval (0,1] such that ϕθS is a Q-local martingale. Consider a sequence of stopping times, (σ n ) n≥1 increasing stationarily to T such that \((\phi\theta\cdot S)^{\sigma_{n}}\) is a true Q-martingale. Then, for any ε>0, the process

$$\left(\frac{\phi}{\phi+\varepsilon+\varepsilon X_{-}}X_{-}\theta\cdot S\right)^{\sigma_{n}} $$

is also a true Q-martingale. As a result,

$$\begin{aligned} E^Q\int_{0}^{\sigma_{n}\wedge T_{n}}X_{s-}\mathrm{d}V_{s}(\theta) =&\lim_{\varepsilon\downarrow0} E^Q\int_{0}^{\sigma_{n}\wedge T_{n}}\frac{\phi_{s}}{\phi_{s}+\varepsilon+\varepsilon X_{s-}}X_{s-}\mathrm{d}V_{s}(\theta) \\ =&\lim_{\varepsilon\downarrow 0}E^Q\left(\frac{\phi}{\phi+\varepsilon+\varepsilon X_{-}}\cdot X_{\sigma_{n}\wedge T_{n}}\right)<\infty. \end{aligned}$$

The first equality follows from the monotone convergence theorem, while the finiteness of the last quantity is due to (41).

Hence, V(θ) is Q-locally integrable and thus (θS) is Q-locally integrable. This proves that (θS) is really a Q-local martingale. Furthermore, \(\exp\bigl(\frac{1}{2}\theta \cdot S\bigr)\) is a positive Q-local submartingale. Then, the condition (40) and de la Vallée Poussin’s argument imply that \(\exp\bigl(\frac{1}{2}\theta\cdot S\bigr)\) is a positive Q-submartingale which is square integrable. Now the lemma follows from Jensen’s inequality. □

For a random variable H, we denote by \({\widetilde{Q}}^{(H)}\) the minimal entropy martingale measure for S with respect to P (H):=eH(E(eH))−1P. Also, Θ 1 denotes the set of strategies considered in [6]:

Lemma 6

Suppose that S is locally bounded and . Let H be a random variable bounded from below with

$$ E\left[\mathrm{e}^{pH}\right]<\infty, $$

for some p∈(1,∞), and let \(\widetilde{\theta}\in\varTheta_{1}\). Then the assertions (i) and (ii) are equivalent:

(i)

$$1-u_0:=\inf_{\theta\in\varTheta_1}E\exp\bigl(H-(\theta\cdot S)_T\bigr)=E\exp\bigl(H-(\widetilde{\theta}\cdot S)_T\bigr). $$

(ii) For any stopping time σT, we have

Proof

Using the results in [6], we change the probability and work under Q instead of P, where

$$Q:=\frac{\exp(H)}{E[\exp(H)]} P. $$

Suppose that the assertion (i) holds. Putting

(42)

where

we obtain the existence of \(\underline{\xi}\) that belongs to the set

$$\varXi:=\Big\{ \xi>0:\ E(\xi)=1,\ \ E(\xi\eta)=0,\ \ \mbox{for any}\ \ \eta:=(\theta\cdot S)_T,\ \theta\in\varTheta_1\Big\} $$

and satisfies

$$J_0=\min_{\xi\in\varXi}E^{Q}(\xi\log\xi)=E^{Q}(\underline{\xi}\log\underline{\xi}). $$

Thus, Theorem 3.5 of [17] implies that

$$ \underline{\xi}=\exp\bigl(-\log\bigl(E^Q\mathrm{e}^{-(\widetilde{\theta}\cdot S)_T}\bigr)-\widetilde{\theta}\cdot S_{T}\bigr)\quad \mbox{and} \quad u_0=1-\mathrm{e}^{-J_0}. $$
(43)

It is clear that the set is stable under concatenation (for more detail about this see [17]), and due to Proposition 4.1 in [17] we conclude that the optimizer of J t is given by . Denoting \(P^{*}:=Z^{*}_{T} Q\) and using the first equation in (43), we derive that

Equivalently, we have

$$ {{Z^{*}_{T}}\over{Z^{*}_{\sigma}}}=\exp\left[-\int_{\sigma}^T \widetilde{\theta}_u \mathrm{d}S_u+J_{\sigma}\right]. $$
(44)

Due to Young’s inequality (xy≤ex+ylogyy), we obtain that

$$\left(-\int_{\sigma}^{T}\theta_{u}dS_{u}\right)\left(\frac {Z^{*}_{T}}{Z^{*}_{\sigma}} \mathrm{e}^{-J_{\sigma}}\right) \leq \mathrm{e}^{-\int _{\sigma}^{T}\theta_{u}\mathrm{d}S_{u}}+\frac{Z^{*}_{T}}{Z^{*}_{\sigma}} \mathrm{e}^{-J_{\sigma}}\log\left(\frac{Z^{*}_{T}}{Z^{*}_{\sigma}} \mathrm{e}^{-J_{\sigma }}\right) -\frac{Z^{*}_{T}}{Z^{*}_{\sigma}} \mathrm{e}^{-J_{\sigma}}. $$

Therefore, by taking conditional expectation on both sides, and using the equalities

and (44), we derive that

Since there is equality for \(\theta=\widetilde{\theta}\), due to

$$\frac{Z^{*}_{T}}{Z^{*}_{\sigma}} \mathrm{e}^{-J_{\sigma}}=\exp\left(-\int_{\sigma }^{T}\widetilde{\theta}_{u}\mathrm{d}S_{u}\right), $$

assertion (ii) follows. The converse is immediate by putting σ=0. This ends the proof of the lemma. □

Lemma 7

Let Z be a given positive local martingale such that ZlogZ is locally integrable. There is a RCLL semimartingale X such that ZX is a local martingale and

$$\log Z=X+h^{E}(Z,P). $$

Proof

Since Z is a positive local martingale, there exists a local martingale N such that N 0=0 and . Due to Ito’s formula, we deduce that

$$ \begin{array}{l}\log Z=\displaystyle{\left(N-\frac{1}{1+\Delta N}\cdot[N,N]\right) +\frac{1}{2}\langle N^{c}\rangle+\sum\frac{(1+\Delta N)\log (1+\Delta N)-\Delta N}{1+\Delta N}}. \end{array} $$

Next, we note that h E(Z,P) is the compensator of the process (1+ΔN)⋅V, where \(Y:=N-\frac{1}{1+\Delta N}\cdot[N,N]\) and V:=log(Z)−Y. Again Ito’s formula implies that

and

where \({\overline{V}}:=V- h^{E}(Z,P)\). Thus, the conclusion follows immediately. □

Lemma 8

Consider a positive σ-martingale density with

$$ {N}={\lambda}\cdot S^{c}+{W}\star(\mu-\nu),\,\,{W}_{t}(x)= \bigl(\mathrm{e}^{{\lambda}_t^{T}x}-1\bigr)\Bigl(1-a_{t}+\int \mathrm{e}^{{\lambda}_t^{T}x}\nu(\{t\},\mathrm{d}x)\Bigr)^{-1}. $$
(45)

Then

$$\begin{aligned} h^{E}({Z},P)&=\frac{1}{2}{\lambda}^{T}c{\lambda}\cdot A+\phi({\gamma }^{-1}\mathrm{e}^{\lambda^T x}-1)\star\nu+\sum(1-a)\phi(\gamma^{-1}-1) \end{aligned}$$
(46)
$$\begin{aligned} &=\frac{1}{2}{\lambda}^{T}c{\lambda}\cdot A+{1\over{\gamma}}\phi\left(\mathrm{e}^{{\lambda}^{T}x}-1\right)\star\nu- \sum{1\over{\gamma}}\left[{\gamma}\log\left({\gamma}\right)-{\gamma }+1\right], \end{aligned}$$
(47)
$$\begin{aligned} \Delta h^{E}({Z},P)&=-\log\gamma, \end{aligned}$$
(48)

where \({\gamma}_{t}:=1-a_{t}+\int \mathrm{e}^{{\lambda}_{t}^{T}x}\nu(\{t\},\mathrm{d}x)\) and ϕ(z):=(1+z)log(1+z)−z.

Proof

Notice that h E(Z,P) is the compensator of V E(N), where

$$ V^{E}\left({N}\right)=\frac{1}{2}\langle {N}^{c}\rangle+\sum\left[\left(1+\Delta{N} \right)\log\left(1+\Delta{N}\right)-\Delta{N}\right]. $$
(49)

From (45) we derive that

$$1+\Delta{N}_{t}=\frac{\mathrm{e}^{{\lambda}_t^{T}\Delta S_{t}}}{{\gamma}_{t}}I_{\{\Delta S_{t}\neq 0\}}+\frac{1}{{\gamma}_{t}}I_{\{\Delta S_{t}=0\}}. $$

After simplification, this leads to the identity

$$\begin{aligned} & \sum\left[\left(1+\Delta{N} \right)\log\left(1+\Delta{N}\right)-\Delta N\right]\\ &\quad =\phi\left ({\gamma}^{-1}\mathrm{e}^{{\lambda}^{T}x}-1\right)\star\mu+ \sum\phi\left({\gamma}^{-1}-1\right)I_{\{\Delta S=0\}}. \end{aligned}$$

By plugging this representation into (49) and compensating, we obtain (46).

Inserting the expression

$$\phi\left({\gamma}^{-1}\mathrm{e}^{{\lambda}^{T}x}-1\right)\star\nu=\gamma ^{-1}\phi\left(\mathrm{e}^{{\lambda}^{T}x}-1\right)\star\nu+\sum{{(1-a-\gamma )\log(\gamma)+(\gamma-1)a}\over{\gamma}} $$

into (46) and simplifying, we get (47).

Calculating the jumps in both sides of (47), we have

$$\begin{aligned} \Delta h^{E}\left({Z},P\right)&=\frac{1}{ {\gamma}}\left(a-\int \mathrm{e}^{{\lambda}^{T}x}F(\mathrm{d}x)\Delta A\right)-\frac{{\gamma}\log\left ({\gamma}\right)-{\gamma}+1}{{\gamma}}\\ &=\frac{-{\gamma}+1}{{\gamma}}-\frac{{\gamma}\log\left ({\gamma}\right)-{\gamma}+1}{{\gamma}} =-\log\left({\gamma}\right). \end{aligned}$$

Note that the first equality follows because

$$\displaystyle\int x\mathrm{e}^{{\lambda}^{T}x}F(\mathrm{d}x)\Delta A=\displaystyle \int x\mathrm{e}^{{\lambda}^{T}x}\nu(\{.\},\mathrm{d}x)=0 $$

in virtue of the fact that Z is a σ-martingale density for S. This ends the proof. □

Lemma 9

Suppose that (3) holds. Then the function

$${\overline{K}}(\lambda):=b^{T}\lambda+\frac{1}{2}\lambda^{T}c\lambda+\int \left(\mathrm{e}^{\lambda^{T}x}-1-\lambda^{T}h(x)\right)F(\mathrm{d}x),\quad\lambda\in \mathbb{R}^d, $$

is convex, proper, closed, and continuously differentiable with

$$\nabla{\overline{K}}(\lambda)=b+c\lambda+\int\left(x\mathrm{e}^{\lambda ^{T}x}-h(x)\right)F(\mathrm{d}x),\quad\lambda\in\mathbb{R}^d. $$

Proof

The proof of this lemma is obvious. For the definitions of proper and closed convex functions, we refer the reader to [25]. □

Lemma 10

The following assertions are equivalent:

(i) For any \(\lambda\in\mathbb{R}^{d}\),

$$ \int_{\{\vert x\vert>1\}}\mathrm{e}^{\lambda^T x} F(\mathrm{d}x)<\infty. $$

(ii) For any \(\lambda\in\mathbb{R}^{d}\),

$$ \int_{\{\vert x\vert>1\}}\vert x\vert \mathrm{e}^{\lambda^T x} F(\mathrm{d}x)<\infty. $$

As a result, if (i) holds, then for any \(\lambda\in\mathbb{R}^{d}\) and q∈(0,∞),

$$ \int_{\{\vert x\vert>1\}}\vert x\vert^q \mathrm{e}^{\lambda^T x} F(\mathrm{d}x)<\infty. $$

Proof

The implication (ii)⟹(i) is obvious. We focus on proving the reverse. Let e i be the element of \(\mathbb{R}^{d}\) that has the i th component equal to one and the other components null. Due to the equivalence of norms in \(\mathbb{R}^{d}\), we may work with the norm \(\vert x\vert=\sum_{i=1}^{d} \vert x_{i}\vert\). We get that

$$\begin{aligned} \int_{\{\vert x\vert>1\}}\vert x\vert \mathrm{e}^{\lambda^T x} F(\mathrm{d}x)&=\sum_{i=1}^d \int_{\{\vert x\vert>1\}} \left((e_i^T x)^+ +(-e_i^T x)^+\right) \mathrm{e}^{\lambda^T x} F(\mathrm{d}x)\\ & \leq \sum_{i=1}^d \int_{\{\vert x\vert >1\}}\mathrm{e}^{(e_i+\lambda)^T x} F(\mathrm{d}x)+\sum_{i=1}^d\int_{\{\vert x\vert >1\}}\mathrm{e}^{(-e_i+\lambda)^T x} F(\mathrm{d}x). \end{aligned}$$

Due to (i) the last term in the rhs of the above string is finite for any \(\lambda\in\mathbb{R}^{d}\). The proof of the remaining part of the lemma follows by the same arguments. □

Appendix 2: MEH σ-Martingale Density Under Change of Probability

In this section, we focus on describing the MEH σ-martingale density when we change probability. This case can be derived easily from a more general case where one works with respect to a positive local martingale density, Z, that may not be uniformly integrable. First, we generalize the characterization of the MEH σ-martingale density for the case when S may not be bounded nor quasi-left continuous. For the case of bounded and quasi-left continuous S, a more elaborate result is given in [3].

Theorem 9

Suppose that and (3) holds. If is the MEH σ-martingale density then there exists \(\widetilde{H}\in L(S)\) such that

$$ \log(\widetilde{Z})=\widetilde{H}\cdot S+h^E(\widetilde{Z},P). $$
(50)

Furthermore, \(\widetilde{H}\) can be described as root of the equation

$$ 0=\left\{ \begin{array}{l@{\quad}l} b+c\theta+\displaystyle \int\bigl[\mathrm{e}^{\theta^T x}x-h(x)\bigr] F(\mathrm{d}x),& \mathit{on}\ \{\Delta A=0\}\\ \displaystyle\int \mathrm{e}^{\theta^T x}x F(\mathrm{d}x),& \mathit{on}\ \{\Delta A\not=0\}. \end{array} \right. $$
(51)

Proof

Notice that the assumptions of Theorem 3.3 in [4] are fulfilled. Hence, a direct application of this theorem implies that \(\widetilde{Z}\) is given by

where \(\widetilde{\beta}\) is a root of (51). Therefore, in the remaining part of this proof we will focus on showing (50). Thus,

$$\begin{aligned} \log(\widetilde{Z})&=\widetilde{N}-\frac{1}{2}\langle \widetilde{N}\rangle+\sum[\log(1+\Delta \widetilde{N})-\Delta\widetilde{N}]\\ & =\widetilde{\beta}\cdot S^{c}+\widetilde{W}\star(\mu-\nu)-\frac{1}{2}\widetilde{\beta }^{T}c\widetilde{\beta}\cdot A+\sum\bigg[\log\frac{\mathrm{e}^{\widetilde{\beta}^{T}x}}{\widetilde{\gamma }}-\frac{\mathrm{e}^{\widetilde{\beta}^{T}x}}{\widetilde{\gamma}}+1\bigg]I_{\{ \Delta S\neq0\}}\\ &\quad {}+\sum\bigg[\log\frac{1}{\widetilde{\gamma}}-\frac {1}{\widetilde{\gamma}}+1\bigg]I_{\{\Delta S=0\}}\\ &=\widetilde{\beta}\cdot S^{c}+\widetilde{W}\star(\mu-\nu)-\frac{1}{2}\widetilde{\beta }^{T}c\widetilde{\beta}\cdot A+\sum\bigg[\frac{-\widetilde{\gamma}\log\widetilde{\gamma}+\widetilde {\gamma}-1}{\widetilde{\gamma}}\bigg]\\ &\quad {}+ \frac{\widetilde{\gamma}\widetilde{\beta}^{T}x-\mathrm{e}^{\widetilde{\beta }^{T}x}+1}{\widetilde{\gamma}}\star\mu. \end{aligned}$$

Note that

$$\begin{aligned} & {1\over{\widetilde{\gamma}}}\bigl(\widetilde{\gamma}\widetilde {\beta}^{T}x-\mathrm{e}^{\widetilde{\beta}^{T}x}+1\bigr)\star\mu\\ &\quad = \widetilde{\beta}^{T}\bigl(x-h(x)\bigr)\star\mu+{1\over{\widetilde{\gamma }}}\bigl(\widetilde{\gamma}\widetilde{\beta}^{T}h(x)-\mathrm{e}^{\widetilde{\beta }^{T}x}+1\bigr)\star(\mu-\nu)\\ &\qquad {}+ {\widetilde{\gamma}}^{-1}\bigl(\widetilde{\gamma}\widetilde{\beta}^{T}h(x)-\mathrm{e}^{\widetilde{\beta}^{T}x}+1\bigr)\star\nu, \end{aligned}$$

since the function \({\widetilde{\gamma}}^{-1}(\widetilde{\gamma}\widetilde{\beta }^{T}h(x)-\mathrm{e}^{\widetilde{\beta}^{T}x}+1)\) is (μν)-integrable which is due to the (μν)−integrability of \({\widetilde{\gamma}}^{-1}(\mathrm{e}^{\widetilde{\beta}^{T}x}-1)=W(x)\) and the boundedness of h(x). Therefore, we get that

$$\begin{aligned} \log(\widetilde{Z})&=\widetilde{\beta}\cdot S^{c}+\widetilde{\beta}^{T}h(x)\star(\mu-\nu)+\widetilde{\beta }^{T}(x-h(x))\star\mu\\ &\quad {}+{\widetilde{\gamma}}^{-1}\bigl(\widetilde{\gamma}\widetilde{\beta }^{T}h(x)-\mathrm{e}^{\widetilde{\beta}^{T}x}+1\bigr)\star\nu-\frac{1}{2}\widetilde {\beta}^{T}c\widetilde{\beta}\cdot A\\ &\quad {}+\sum{\widetilde{\gamma}}^{-1}(-\widetilde{\gamma}\log(\widetilde {\gamma})+\widetilde{\gamma}-1). \end{aligned}$$

Equivalently, we deduce that

$$ \log\widetilde{Z}=\widetilde{\beta}\cdot S+\frac{1}{2}\widetilde{\beta}^{T}c\widetilde{\beta}\cdot A+ \frac{\widetilde{\beta}^{T}x\mathrm{e}^{\widetilde{\beta}^{T}x}-\mathrm{e}^{\widetilde {\beta}^{T}x}+1}{\widetilde{\gamma}}\star\nu+ \sum{\widetilde{\gamma}}^{-1}(-\widetilde{\gamma}\log(\widetilde{\gamma })+\widetilde{\gamma}-1), $$

since

$$\widetilde{\beta}\cdot S=\widetilde{\beta}\cdot S^{c}+\widetilde{\beta}^{T}b\cdot A+\widetilde{\beta}^{T}h(x)\star(\mu-\nu)+\widetilde{\beta }^{T}(x-h(x))\star\mu. $$

Therefore, by a direct application of Lemma 8 for \(\lambda=\widetilde{\beta}\), (50) follows immediately from putting \(\widetilde{H}=\widetilde{\beta}\). This ends the proof of the theorem. □

In what follows, we denote by a positive local martingale with

$$ N:=\beta\cdot S^{c}+W\star(\mu-\nu)+g\star\mu+\overline{N},\qquad W_{t}(x):=f_{t}(x)+\frac{\widehat{f_{t}}}{1-a_{t}}I_{\{a_{t}<1\}}, $$
(52)

where \(\left(\beta,f,g,\overline{N}\right)\) are Jacod’s components of N. Here, we define:

where is given by (2).

Theorem 10

Consider Z defined in (52) and suppose that

Then the minimization problem

(53)

admits a solution given by

$$\widetilde{N}=\widetilde{\beta}\cdot S^{c,Z}+\widetilde{W}\star(\mu-\nu ^{Z}),\qquad \widetilde{W}_{t}(x)=\frac{\mathrm{e}^{\widetilde{\beta }_t^{T}x}-1}{1-a_{t}^{Z}+\int \mathrm{e}^{\widetilde{\beta}_t^{T}y}\nu^{Z}(\{t\},\mathrm{d}y)}, $$

where \(\widetilde{\beta}\) is the root of the equation

$$ 0=b^{Z}+c\lambda+\int(\mathrm{e}^{\lambda^{T}x}x-h(x))F^{Z}(\mathrm{d}x). $$
(54)

Here S c,Z, b Z, a Z, ν Z and F Z are given by

$$\begin{aligned} S^{c,Z}&:=S^{c}-c\beta\cdot A,\qquad b^{Z}:=b+c\beta-\int f(x)h(x)F(\mathrm{d}x),\\ a^{Z}_{t}&:=\nu^{Z}(\{t\},\mathbb{R}^{d}\setminus\{0\}) \end{aligned}$$

and

$$\nu^{Z}(\mathrm{d}t,\mathrm{d}x):=F^{Z}_t(\mathrm{d}x)\mathrm{d}A_{t},\qquad F^{Z}_{t}(\mathrm{d}x):=(1+f_{t}(x))F_{t}(\mathrm{d}x). $$

Proof

Consider a sequence of stopping times, (T n ) n≥1, stationarily increasing to T and such that \(Z^{T_{n}}\) is a true martingale. For a fixed but arbitrary n we denote \(Q:=Z_{T_{n}}\cdot P\). Note that all the equations in the theorem are robust under stopping. Due to Lemma 1, it is sufficient to prove that the theorem holds on 〚0,T n 〛. Then, we obtain that \(\nu^{Q}(\mathrm{d}t,\mathrm{d}x)=(1+f_{t}(x))I_{\{t\leq T_{n}\}}\nu(\mathrm{d}t,\mathrm{d}x)\), , and \(\int_{\{\vert x\vert>1\}}\mathrm{e}^{\lambda^{T}x}F^{Q}(\mathrm{d}x)<\infty\) for \(\lambda\in\mathbb{R}^{d}\).

Therefore, the assumptions of Theorem 3.3 in [4] are fulfilled. By direct application of this theorem for \(S^{T_{n}}\) and under the measure \(Q=Z_{T_{n}}\cdot P\), we deduce that the problem defined in (53) admits a solution , where \(\widetilde{N}^{Q}\) is given, on 〚0,T n 〛, by

$$\widetilde{N}^{Q}=\widetilde{\beta}\cdot S^{c,Q}+\widetilde{W}\star(\mu -\nu^{Q}),\qquad \widetilde{W}_{t}(x)=\frac{\mathrm{e}^{\widetilde{\beta }_t^{T}x}-1}{1-a_{t}^{Q}+\int \mathrm{e}^{\widetilde{\beta}_t^{T}y}\nu^{Q}(\{t\},\mathrm{d}y)}. $$

Herein S c,Q is the continuous local martingale part of S under Q and ν Q is the Q-compensator measure of μ, and \(a^{Q}_{t}=\nu^{Q}(\{t\},\mathbb{R}^{d}\setminus\{ 0\})\). Moreover, \(\widetilde{\beta}\) is given by the equation

(55)

It is then clear that \(\widetilde{N}^{Q}\) coincides with \(\widetilde{N}^{T_{n}}\) of the theorem and that the equation (55) is exactly the equation (54) on 〚0,T n 〛. This ends the proof of theorem. □

Theorem 11

Let Z be a positive local martingale and let . If the assumptions of Theorem 10 are fulfilled and \(\widetilde{Z}\) is the MEH local martingale density with respect to Z, then

$$ \log\widetilde{Z}=\widetilde{\beta}\cdot S+h^E(\widetilde{Z}, Z) $$

and \(\widetilde{\beta}\) is a root of (54).

Proof

The proof of this theorem follows from the same arguments as in the proofs of Theorems 9 and 10. □

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Choulli, T., Ma, J., Morlais, MA. (2014). Three Essays on Exponential Hedging with Variable Exit Times. In: Kabanov, Y., Rutkowski, M., Zariphopoulou, T. (eds) Inspired by Finance. Springer, Cham. https://doi.org/10.1007/978-3-319-02069-3_7

Download citation

Publish with us

Policies and ethics