Skip to main content
Log in

Pricing and hedging basis risk under no good deal assumption

  • Research Article
  • Published:
Annals of Finance Aims and scope Submit manuscript

Abstract

We consider the problem of explicitly pricing and hedging an option written on a non-exchangeable asset when trading in a correlated asset is possible. This is a typical case of incomplete market where it is well known that the super-replication concept provides generally too high prices. We study several prices and in particular the instantaneous no-good-deal price (see Cochrane and Saa-Requejo in J Polit Econ 108(1):79–119, 2001) and the global one. We show numerically that the global no-good-deal price can be significantly higher that the instantaneous one. We then propose several hedging strategies and show numerically that the mean-variance hedging strategy can be efficient.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. As the sulfur rate of the Straight Run is lower than that the one of the FO 1 %, the SR 0.5 % is less liquid in the commodity market.

  2. More precisely, their result involves local (instead of true) martingales in the definition of \(\mathcal{M}^{e}({\mathbb {P}})\).

  3. by \({\mathbb {Q}}\in L^2({\mathbb {P}})\), we mean that the density of \({\mathbb {Q}}\) w.r.t. \({\mathbb {P}}\) is in \(L^2({\mathbb {P}})\).

  4. Since \(\mathcal{M}^{2,l}({\mathbb {P}})\) is non-empty, Theorem 2.2 of Cherny (2008) ensures that \(u^l\) is a so-called coherent utility function.

  5. Note a similar result that has already been obtained by Klöppel and Schweizer (2007) using utilities instead of Sharpe ratio.

  6. In fact, if we start with a finite wealth \(X_{0}\), since \(H\) depends on \(W^*\) through the non-traded asset \(V\), we have that for any strategy \((\varPhi ^0,\varPhi ),\,{\mathbb {P}}[ \varPhi ^0_{T}S^0_{T}+ \varPhi _{T}S_{T}<H]\ne 0\). Now, in the case of a call option, if the investor is endowed with one unit of \(V\) : \({\mathbb {P}}(V_T \ge (V_T-K)_+)=1\).

  7. The equality between problems (50) and (51) comes from the following observations: let \(Y^{opt}\) be the solution of problem (51), and \(Y_t^{opt}={\mathbb {E}}^{{\mathbb {Q}}^U}{\small \big (Y^{opt}\mid \mathcal{F}_t^{W^*}\big )}\). As \((Y_t^{opt})_t\) is a \(L^{2}~({\mathbb {Q}}^U,\mathcal{F}^{W^*})\)-martingale, from Theorem of martingale representation (see for example Theorem B.1.3 Musiela and Rutkowski (2007)) there exists \(k_t \in L^{2}_{loc}(W^{*})\) such that \(dY_t^{opt}=k_t d W_t^*\). Let \(\lambda ^{opt}_t=k_t/Y_t^{opt}\) (note that \(Y_t^{opt}>0\)), \(\lambda ^{opt}_t \in \mathcal{F}_t^{W^*}\). By Itô’s formula, \(Y_T^{opt}=Y_0^{opt} + \int _0^T\lambda ^{opt}_t Y_t^{opt}dW^*_t=1+ \int _0^T\lambda ^{opt}_t Y_t^{opt}dW^*_t=Y_T^{\lambda ^{opt}}\). Thus \(\lambda ^{opt}\) satisfies condition of (50) and problem (51) is lower than problem (50). Let \(\lambda \) satisfy condition of (50) then \(Y_T^{\lambda }\) satisfies condition of (51) and thus the values of the two problems are equal.

  8. \(\mathcal{N}\) is the normal distribution function, i.e. \(\mathcal{N}(d) = \int _{-\infty }^d \frac{e^{-x^2/2}}{\sqrt{2\pi }}dx\).

  9. The term \({\mathbb {E}}^{2}[A]\) denotes \(({\mathbb {E}}[A])^{2}\).

References

  • Bayraktar, E., Young, V.: Pricing options in incomplete equity markets via the instantaneous Sharpe ratio. Ann. Financ. 4(4), 399–429 (2008)

    Article  Google Scholar 

  • Becherer, D.: From bounds on optimal growth towards a theory of good-deal hedging. In: Albrecher, H. et al., (eds.) Advanced Financial Modelling, vol. 8, pp. 27–51. Berlin: Walter de Gruyter. Radon Series on, Computational and Applied Mathematics (2009)

  • Björk, T., Slinko, I.: Towards a general theory of good-deal bounds. Rev. Financ. 10(2), 221–260 (2006)

    Article  Google Scholar 

  • Cherny, A.S.: Pricing with coherent risk. Theory Probab. Appl. 52(3), 389–415 (2008)

    Article  Google Scholar 

  • Cochrane, J., Saa-Requejo, J.: Beyond arbitrage: good-deal asset price bounds in incomplete markets. J. Polit. Econ. 108(1), 79–119 (2001)

    Article  Google Scholar 

  • Delbaen, F., Schachermayer, W.: A general version of the fundamental theorem of asset pricing. Math. Ann. 300(3), 463–520 (1994)

    Article  Google Scholar 

  • Duffie, D., Richardson, H.R.: Mean variance hedging in continuous time. Ann. Appl. Probab. 1(1), 1–15 (1991)

    Article  Google Scholar 

  • Föllmer, H., Sondermann, D.: Hedging of nonredundant contingent claims. In: Contributions to mathematical economics, pp. 205–223. North-Holland, Amsterdam (1986)

  • Gourieroux, C., Laurent, J., Pham, H.: Mean-variance hedging and numéraire. Math. Financ. 8(3), 179–200 (1998)

    Article  Google Scholar 

  • Karatzas, I., Shreve, S.: Brownian motion and stochastic calculus, vol. 113 of Graduate Texts in Mathematics, 2nd edn. Springer, New York (1991)

  • Klöppel, S.: Dynamic valuations in incomplete markets, Ph.D. thesis (2006)

  • Klöppel, S., Schweizer, M.: Dynamic utility-based good deal bounds. Stat. Decis. 25(4), 285–309 (2007)

    Google Scholar 

  • Musiela, M., Rutkowski, M.: Martingale Methods in Financial Modeling, vol. 36 of Stochastic Modelling and Applied Probability. Springer, Berlin (2007)

  • Protter, P.: Stochastic integration and differential equations, vol. 21 of Applications of Mathematics (New York). A new approach. Springer, Berlin (1990).

  • Revuz, D., Yor, M.: Continuous martingales and Brownian motion, vol. 293 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 2nd edn. Springer, Berlin (1994)

  • Schweizer, M.: Mean-variance hedging for general claims. Ann. Appl. Probab. 2(1), 171–179 (1992)

    Article  Google Scholar 

  • Schweizer, M.: A guided tour through quadratic hedging approaches. In Jouini, E. et al. (eds.) Option pricing, interest rates and risk management, pp. 538–574. Handbooks in Mathematical Finance, Cambridge University Press, Cambridge (2001)

Download references

Acknowledgments

We thank an anonymous referee for his/her careful reading and for valuable comments which benefited this version.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. Carassus.

Appendices

Appendices

1.1 Technical results

1.1.1 On Black-Scholes formula

We recall the following formula which is analogous to the Black-Scholes formula. All proofs are omitted since they are completely similar to the one of Black-Scholes model which can be found for example in Musiela and Rutkowski (2007) (starting from page 94).

Let \(Y\) be a geometric Brownian motion, with drift \(\eta \) and volatility \(\varphi \), i.e.

$$\begin{aligned} Y_t = Y_0 \exp \left( \left( \eta - \frac{\varphi ^2}{2} \right) t + \varphi W_t \right) . \end{aligned}$$

Then, the function \(BS(Y_t,T-t,K,\eta ,\varphi )\) defined by

$$\begin{aligned} BS\left( Y_t,T-t,K,\eta ,\varphi \right) = {\mathbb {E}}\left[ (Y_T-K)_+ \mid \mathcal{F}_{t}\right] , \end{aligned}$$
(80)

can be explicitly expressed as

$$\begin{aligned} BS(Y_t,T-t,K,\eta ,\varphi )&= Y_t e^{\eta (T-t)} \mathcal{N}\left( d_1(Y_t,T-t,K,\eta ,\varphi )\right) \nonumber \\&- K \mathcal{N}\left( d_0(Y_t,T-t,K,\eta ,\varphi )\right) \end{aligned}$$
(81)

whereFootnote 8

$$\begin{aligned} d_1(Y_t,T-t,K,\eta ,\varphi )&= \frac{ \ln \left( \frac{Y_t}{K} \right) + \left( \eta + \frac{\varphi ^2}{2} \right) (T-t) }{\varphi \sqrt{T-t}} ;\\ d_0(Y_t,T-t,K,\eta ,\varphi )&= d_1 -\varphi \sqrt{T-t}. \end{aligned}$$

By Ito’s Formula, one can show for any \(x>0\) that:

$$\begin{aligned} d BS(x,T-t,K,\eta ,\varphi )&= \varphi e^{\eta (T-t)} \mathcal{N}(d_1(x,T-t,K,\eta ,\varphi ))xdW_t. \end{aligned}$$
(82)
$$\begin{aligned} \frac{\partial BS}{\partial \eta } (x,T-t,K,\eta ,\varphi )&= (T-t)K \mathcal{N}(d_{0}(x,T-t,K,\eta ,\varphi ))>0 \end{aligned}$$
(83)
$$\begin{aligned} \frac{\partial BS}{\partial \varphi } (x,T-t,K,\eta ,\varphi )&= \sqrt{T-t} x \frac{e^{\eta (T-t)}}{\sqrt{2\pi }}e^{-\frac{d_{1}^{2}(x,T-t,K,\eta ,\varphi )}{2}} > 0, \end{aligned}$$
(84)

(83) and (84) implies that the Black-Scholes function \(BS\) is increasing with respect to \(\eta \) and \(\varphi \). Moreover, the function BS satisfies the following linearity property:

$$\begin{aligned} \forall \alpha \in {\mathbb {R}}_{+}, \qquad BS(\alpha Y_t,T-t,\alpha K,\eta ,\varphi ) = \alpha BS(Y_t,T-t,K,\eta ,\varphi ). \end{aligned}$$
(85)

1.1.2 Proof of Lemma 1

Proof

Let \(\lambda \in \varLambda \). Since \(Z_{T}^\lambda \) is a martingale, one has \({\mathbb {E}}Z_T^\lambda =Z_{0}^{\lambda }=1\). We define the following process

$$\begin{aligned} \bar{Z}_{t}^\lambda = \exp \left( -2h_S W_t - 2h_S^2 t +2\int \limits _0^t \lambda _s dW^*_s - 2 \int \limits _0^t \lambda _s^2 ds\right) . \end{aligned}$$

It is a Doléans-Dade process and thus a continuous local martingale (see Karatzas and Shreve 1991, page 191). It is easy to see from the maximal inequality that \(\bar{Z}^\lambda \) is in fact a martingale (since it is of class (DL)).

Then, we can define the following probability measure

$$\begin{aligned} d\tilde{{\mathbb {Q}}}/d{\mathbb {P}}= \bar{Z}_{T}^\lambda . \end{aligned}$$

Using Bayes Formula

$$\begin{aligned} {\mathbb {E}}\left( \left( \frac{Z_T^\lambda }{Z_t^\lambda }\right) ^2 \mid \mathcal{F}_t\right) = e^{h_S^2(T-t)}{\mathbb {E}}^{\tilde{{\mathbb {Q}}}}\left( e^{\int _t^T \lambda _s^2ds}\mid \mathcal{F}_t \right) . \end{aligned}$$

\(\square \)

1.1.3 Proof of Lemma 5

We recall the following lemma for ease of exposition. Let \(X \in L^2,\,X \ge 0\) such that \(X {\mathbf 1}_{X>0}\) has a density with respect to Lebesgue measure and \(\gamma \) be a positive number. One has

$$\begin{aligned} p_0^{opt} = {\mathop {\mathop {\sup }\limits _{Y \ge 0, \; {\mathbb {E}}Y =1 }}\limits _{{\mathbb {E}}Y^2 \le 1 +\gamma ^2}} {\mathbb {E}}\left[ Y X \right] \end{aligned}$$
(86)

Lemma 6

The optimal solution of problem (86) is given as follows:

  1. 1.

    If \(1-\gamma \frac{{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} \ge 0\) then one has \(Y^{opt} = 1+\gamma \frac{X-{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}}\) and \(p_0^{opt} = {\mathbb {E}}\left( X\right) + \gamma \sqrt{\mathop {\mathrm{Var}}\nolimits {X}}\).

  2. 2.

    If \(1-\gamma \frac{{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}}< 0\) then there exists a positive number \(\alpha \), such thatFootnote 9

    $$\begin{aligned} \frac{{\mathbb {E}}\left( X-\alpha \right) ^2_+}{{\mathbb {E}}^2\left( X-\alpha \right) _+}&= 1+\gamma ^2, \end{aligned}$$
    (87)

    and one has \(Y^{opt} = \frac{\left( X-\alpha \right) _+}{{\mathbb {E}}\left( X-\alpha \right) _+}\) and \(p_0^{opt} = \alpha + (1+\gamma )^2{\mathbb {E}}\left( X-\alpha \right) _+\).

Proof

1. If \(1-\gamma \frac{{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} \ge 0\), then it is straightforward that \({\mathbb {E}}Y^{opt} X={\mathbb {E}}\left( X\right) + \gamma \sqrt{\mathop {\mathrm{Var}}\nolimits {X}}\). Let \(Y\) such that \(Y \ge 0,\,{\mathbb {E}}Y =1\) and \({\mathbb {E}}Y^2 \le 1 +\gamma ^2\) then \(\mathop {\mathrm{Var}}\nolimits Y \le \gamma ^2\) and by the Cauchy-Schwarz inequality

$$\begin{aligned} {\mathbb {E}}(Y X)&\le {\mathbb {E}}\left( (Y-{\mathbb {E}}Y) (X-{\mathbb {E}}X) \right) + {\mathbb {E}}X \le \sqrt{\mathop {\mathrm{Var}}\nolimits X} \sqrt{\mathop {\mathrm{Var}}\nolimits Y} + {\mathbb {E}}X \\&\le \gamma \sqrt{\mathop {\mathrm{Var}}\nolimits X} + {\mathbb {E}}X ={\mathbb {E}}[ Y^{opt} X]. \end{aligned}$$

To prove that \(Y^{opt}\) is the optimal solution of (86), it remains to verify that it satisfies the constraints of (86). One has \(Y^{opt} = 1-\gamma \frac{ {\mathbb {E}}X}{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} +\gamma \frac{X}{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} \ge 0\) by assumption (recall that \(X \ge 0\)). The two others constraints are straightforward.

2. If \(1-\gamma \frac{{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} < 0\), assume that there exists \(\alpha \) such that condition (87) is satisfied. Then it is straightforward that

$$\begin{aligned} {\mathbb {E}}\left[ Y^{opt} X \right]&= \frac{{\mathbb {E}}\left[ \left( X-\alpha \right) _+(X-\alpha +\alpha ) \right] }{{\mathbb {E}}\left( X-\alpha \right) _+} \\&= \alpha + \frac{{\mathbb {E}}\left( X-\alpha \right) ^2_+}{{\mathbb {E}}\left( X-\alpha \right) _+} = \alpha + (1+\gamma ^2){\mathbb {E}}\left( X-\alpha \right) _+ \end{aligned}$$

using condition (87). Let \(Y\) such that \(Y \ge 0,\,{\mathbb {E}}Y =1\) and \({\mathbb {E}}Y^2 \le 1 +\gamma ^2\). Then by the Cauchy-Schwarz inequality and condition (87)

$$\begin{aligned} {\mathbb {E}}[Y X]&= {\mathbb {E}}\left( Y\left( X-\alpha \right) _+ \right) + \alpha + {\mathbb {E}}\left( Y\left( X-\alpha \right) {\mathbf 1}_{X < \alpha } \right) \\&\le \sqrt{{\mathbb {E}}{\left( X-\alpha \right) ^2_+}} \sqrt{{\mathbb {E}}Y^2} + \alpha \\&\le \sqrt{1 +\gamma ^2}\sqrt{1 +\gamma ^2} {\mathbb {E}}{\left( X-\alpha \right) _+} + \alpha ={\mathbb {E}}\left[ Y^{opt} X\right] . \end{aligned}$$

The solution \(Y^{opt}\) is thus optimal for the optimization problem (86) since it satisfies the constraints (see condition (87)).

It remains to prove that there exists some \(\alpha \) such that condition (87) is satisfied. Let \(f(x)= \frac{{\mathbb {E}}\left( X-x\right) ^2_+}{{\mathbb {E}}^2\left( X-x\right) _+}\) then \(f(0)=\frac{\mathop {\mathrm{Var}}\nolimits X}{{\mathbb {E}}^2 X} + 1 < 1+ \gamma ^2\) by assumption. Below we show that there exists \(\alpha _0>0\) such that \(f(\alpha _0) \ge 1+ \gamma ^2\), thus by continuity of \(f\) there will exist some \(\alpha >0\) such that \(f(\alpha ) = 1+ \gamma ^2\). We prove first that there exist \(\alpha _0\) such that \({\mathbb {P}}\left( X>\alpha _0\right) =\frac{1}{1+ \gamma ^2}\). Such an \(\alpha _0\) exists since

$$\begin{aligned} 1-\gamma \frac{{\mathbb {E}}\left( X\right) }{\sqrt{\mathop {\mathrm{Var}}\nolimits {X}}} < 0&\Leftrightarrow \mathop {\mathrm{Var}}\nolimits \left[ X {\mathbf 1}_{X>0}\right] <\gamma ^2 {\mathbb {E}}^2\left[ X {\mathbf 1}_{X>0}\right] \\&\Leftrightarrow {\mathbb {E}}\left[ X^2 {\mathbf 1}_{X>0}\right] <(\gamma ^2+1) {\mathbb {E}}^2\left[ X \left( \sqrt{{\mathbf 1}_{X>0}}\right) ^2\right] \\&\Rightarrow {\mathbb {E}}\left[ X^2 {\mathbf 1}_{X>0}\right] <(\gamma ^2+1) {\mathbb {E}}\left[ X^2 {\mathbf 1}_{X>0}\right] {\mathbb {P}}(X>0), \end{aligned}$$

by the Cauchy-Schwarz inequality. Thus one has \({\mathbb {P}}(X=0)\le \frac{\gamma ^2}{1+ \gamma ^2}\) and by continuity of \(x\rightarrow {\mathbb {P}}(X\le x)\), for \(x>0\) there exists \(\alpha _0\) such that \({\mathbb {P}}(X\le \alpha _0)=\frac{\gamma ^2}{1+ \gamma ^2}\). Then, by the Cauchy-Schwarz inequality, one has

$$\begin{aligned} {\mathbb {E}}^2\left( X\!-\!\alpha _0\right) _+&= {\mathbb {E}}^2( X {\mathbf 1}_{X >\alpha _0}) -2 \alpha _0 {\mathbb {P}}(X>\alpha _0) {\mathbb {E}}(X {\mathbf 1}_{X >\alpha _0})+\alpha _0^2 {\mathbb {P}}^2(X>\alpha _0) \\&\le {\mathbb {P}}(X\!>\!\alpha _0) {\mathbb {E}}(X^2 {\mathbf 1}_{X >\alpha _0}) \!-\!2 \alpha _0 {\mathbb {P}}(X\!>\!\alpha _0){\mathbb {E}}(X {\mathbf 1}_{X >\alpha _0})\!+\!\alpha _0^2 {\mathbb {P}}^2(X\!>\!\alpha _0) \\&\le \frac{1}{1+ \gamma ^2} \left( {\mathbb {E}}X^2 {\mathbf 1}_{X >\alpha _0} - 2 \alpha _0 {\mathbb {E}}X {\mathbf 1}_{X >\alpha _0}+\alpha _0^2 {\mathbb {P}}(X>\alpha _0) \right) \\&\le \frac{1}{1+ \gamma ^2} {\mathbb {E}}\left( X-\alpha _0\right) ^2_+. \end{aligned}$$

Thus one has \(f(\alpha _0) \ge 1+ \gamma ^2\), which concludes the proof. \(\square \)

1.1.4 Proof of Theorem 6

Proof

In order to apply the results of Schweizer (1992), we need to compute the decomposition of \(e^{-rt}{p}^{0}_{t}\) w.r.t. \(W^{0}_{t}\) and \(W^{0,*}_{t}\). The process \(V\) under the probability \({\mathbb {Q}}^{0}\) is a geometric Brownian motion and we can achieve these decomposition using Itô’s formula. Under \({\mathbb {Q}}^0,\,V\) satisfies the following SDE (see (8))

$$\begin{aligned} \frac{dV_t}{V_t} = \eta ^0 dt + \sigma _V \left( \rho dW^{0}_t + \sqrt{1- \rho ^2} dW^{*,0}_{t}\right) , \end{aligned}$$

see (9) for the definition of \(\eta ^0\) and (6) for the definition of the \({\mathbb {Q}}^{0}\)-Brownian motions \(W^{0}\) and \(W^{0,*}\). From (72) and (80), we obtain that

$$\begin{aligned} e^{-rt}{p}^{0}_{t} = e^{-rT} {\mathbb {E}}^{{\mathbb {Q}}^0}\left( (V_T-K)_{+}\right) =e^{-rT}BS \left( V_t,T-t,K,\eta ^0,\sigma _{V}\right) . \end{aligned}$$

Equations (82) and (64) imply

$$\begin{aligned} d BS\left( V_t,T-t,K,\eta ^0,\sigma _{V}\right) =e^{r(T-t)} L_{t}^{0} \left( \rho dW^{0}_t + \sqrt{1- \rho ^2} dW^{*,0}_{t}\right) , \end{aligned}$$

where we used the shorter notation \(d_1\) for \(d_1(V_t,T-t,K,\eta ^0 ,\sigma _V)\). Thus,

$$\begin{aligned} d (e^{-rt}{p}^{0}_{t})&= e^{-rt} L^0_{t}\left( \rho dW^{0}_t + \sqrt{1- \rho ^2} dW^{*,0}_t\right) \\&= e^{-rt} L^0_{t}\left( \frac{\rho e^{rt}}{\sigma _{S}S_{t}} d(e^{-rt}S_{t}) + \sqrt{1- \rho ^2} dW^{*,0}_t\right) . \end{aligned}$$

We now apply the theorem stated on page 175 of Schweizer (1992) and his Eq. (2.4). Let \(G^{*}\) be the solution of the SDE

$$\begin{aligned} dG^{*}_{t} = \varPhi (G^{*}_{t}) \sigma _{S} e^{-rt}S_{t}dW^{0}_{t}, \quad G^{*}_{0}=0, \end{aligned}$$
(88)

where

$$\begin{aligned} \varPhi (x) = \frac{\rho L^0_{t}}{\sigma _{S}S_{t}} + \frac{h_{S} e^{rt}}{\sigma _{S} S_{t}} \left( e^{-rt}{p}^{0}_{t} -X_{0} - x \right) . \end{aligned}$$
(89)

Then the hedging strategy \(\varPhi (G^{*}_{t})\) solves the optimal solution for problem (68). It remains to compute \(G^{*}_{t}\), i.e. solve the SDE (88). If we set

$$\begin{aligned} \bar{G}^{*}_{t} = e^{-rt}{p}^{0}_{t} -X_{0} -G^{*}_{t}, \end{aligned}$$
(90)

the process \(\bar{G}^{*}_{t}\) is solution of the following SDE:

$$\begin{aligned} d\bar{G}^{*}_{t}&= d\left( e^{-rt} {p}^{0}_{t}\right) - dG^{*}_{t} = e^{-rt} L^0_{t}\left( \rho dW^{0}_t + \sqrt{1- \rho ^2} dW^{*,0}_t\right) \\&- \varPhi (G^{*}_{t}) \sigma _{S}S_{t} e^{-rt} dW^{0}_{t} \\&= e^{-rt} L^0_{t}\left( \rho dW^{0}_t + \sqrt{1- \rho ^2} dW^{*,0}_t\right) \\&- \rho e^{-rt} L^0_{t} dW^{0}_{t} - h_{S} \left( e^{-rt}{p}^{0}_{t} -X_{0} - G^{*}_{t} \right) d W^{0}_{t}\\&= e^{-rt} L^0_{t} \sqrt{1- \rho ^2} dW^{*,0}_t - h_{S} \bar{G}^{*}_{t}d W^{0}_{t}, \end{aligned}$$

with the initial condition \(\bar{G}^{*}_{0} = {p}_{0}^{0} -X_{0}\). The solution of this SDE is given by:

$$\begin{aligned} \bar{G}^{*}_{t}&= e^{-h_{S} W^{0}_{t} - \frac{h_{S}^{2}}{2}t} \left( {p}_{0}^{0} -X_{0} + \int \limits _{0}^{t} \sqrt{1-\rho ^{2}} L^0_{s} e^{h_{S} W^{0}_{s} +\frac{h_{S}^{2}}{2}s-rs} dW^{0,*}_{s} \right) \\&= U_te^{-rt} \left( {p}_{0}^{0} -X_{0} + \int \limits _{0}^{t} \sqrt{1-\rho ^{2}} \frac{L^0_{s}}{U_s} dW^{0,*}_{s} \right) , \end{aligned}$$

since by (69) \( U_{t} = e^{-h_SW_t +(r -3/2h_S^2)t}= e^{rt}e^{-h_SW^0_t -1/2h_S^2t}\) (recall that \(W^0_t= W_t +h_St\) and \(W^{0,*}_t= W^*_t \)). Thus using (90) we obtain that:

$$\begin{aligned} G^{*}_{t} = e^{-rt}{p}^{0}_{t} -X_{0} - U_te^{-rt} \left( {p}_{0}^{0} -X_{0} + \int \limits _{0}^{t} \sqrt{1-\rho ^{2}} \frac{L^0_{s}}{U_s} dW^{0,*}_{s}\right) . \end{aligned}$$
(91)

Using (89) we obtain that:

$$\begin{aligned} \varPhi (G^{*}_{t}) = \frac{\rho L^0_{t}}{\sigma _{S}S_{t}} + \frac{h_{S}U_{t}}{\sigma _{S} S_{t}} \left( {p}_{0}^{0} -X_{0} + \sqrt{1-\rho ^{2}} \int \limits _{0}^{t} \frac{L^0_{s} }{U_{s}}dW^{*}_{s} \right) \end{aligned}$$

and (73) is proved. To obtain (74), we rewrite the problem (68). Using (88), we obtain that \(\int _0^T\varPhi (G^{*}_{t})d(e^{-rt}S_t)=\int _0^T\varPhi (G^{*}_{t})S_te^{-rt} \sigma _SdW^0_t =G^{*}_{T}\). Thus using (91), we obtain that

$$\begin{aligned} v^{X_{0}}(H)&= e^{2rT} {\mathbb {E}}\left[ U_Te^{-rT} \left( {p}_{0}^{0} -X_{0} + \int \limits _{0}^{T} \sqrt{1-\rho ^{2}} \frac{L^0_{t}}{U_t} dW^{0,*}_{t}\right) \right] ^{2} \\&= {\mathbb {E}}\left[ U_{T}^{2} \right] {\mathbb {E}}^{{\mathbb {Q}}^{U}} \left[ {p}_{0}^{0} -X_{0} + \sqrt{1-\rho ^{2}}\int \limits _{0}^{T} \frac{ L^0_{t} }{U_{t}} dW^{U,*}_{t} \right] ^{2}, \end{aligned}$$

since \(W^{0,*}=W^{U,*}\) is a Brownian motion under \({\mathbb {Q}}^{U}\) (see (71)) and since \({\mathbb {E}}\left[ U_{T}^{2} \right] =e^{(2r-h_{S}^{2})T}\): (74) follows. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Carassus, L., Temam, E. Pricing and hedging basis risk under no good deal assumption. Ann Finance 10, 127–170 (2014). https://doi.org/10.1007/s10436-013-0246-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10436-013-0246-1

Keywords

JEL Classification

Navigation