Skip to main content
Log in

Hedging under generalized good-deal bounds and model uncertainty

  • Original Article
  • Published:
Mathematical Methods of Operations Research Aims and scope Submit manuscript

Abstract

We study a notion of good-deal hedging, that corresponds to good-deal valuation and is described by a uniform supermartingale property for the tracking errors of hedging strategies. For generalized good-deal constraints, defined in terms of correspondences for the Girsanov kernels of pricing measures, constructive results on good-deal hedges and valuations are derived from backward stochastic differential equations, including new examples with explicit formulas. Under model uncertainty about the market prices of risk of hedging assets, a robust approach leads to a reduction or even elimination of a speculative component in good-deal hedging, which is shown to be equivalent to a global risk minimization in the sense of Föllmer and Sondermann (1986) if uncertainty is sufficiently large.

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

Similar content being viewed by others

References

  • Aliprantis CD, Border KC (2006) Infinite dimensional analysis: a Hitchhiker’s guide, 3rd edn. Springer, Berlin

    MATH  Google Scholar 

  • Artzner P, Delbaen F, Eber JM, Heath D, Ku H (2007) Coherent multiperiod risk adjusted values and Bellman’s principle. Ann Oper Res 152(1):5–22

    Article  MathSciNet  MATH  Google Scholar 

  • Barrieu P, El Karoui N (2009) Pricing, hedging and optimally designing derivatives via minimization of risk measures. In: Carmona R (ed) Indifference pricing: theory and applications. Princeton University Press, Princeton, pp 77–146

    Google Scholar 

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

    Article  MATH  Google Scholar 

  • Becherer D (2009) From bounds on optimal growth towards a theory of good-deal hedging. Radon Ser Comput Appl Math 8:27–52

    MathSciNet  MATH  Google Scholar 

  • Becherer D, Turkedjiev P (2014) Multilevel approximation of backward stochastic differential equations. Preprint, arXiv:14123140

  • Biagini S, Pınar MÇ (2017) The robust Merton problem of an ambiguity averse investor. Math Financ Econ 11(1):1–24. doi:10.1007/s11579-016-0168-6

  • Bion-Nadal J, Di Nunno G (2013) Dynamic no-good-deal pricing measures and extension theorems for linear operators on \(L^\infty \). Financ Stoch 17(3):587–613

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  • Bondarenko O, Longarela I (2009) A general framework for the derivation of asset price bounds: an application to stochastic volatility option models. Rev Deriv Res 12(2):81–107

    Article  MATH  Google Scholar 

  • Boyarchenko N, Cerrato M, Crosby J, Hodges SD (2014) No good deals—no bad models. SSRN-id2467066, NY Federal Reserve Staff Report No. 589. doi:10.2139/ssrn.2192559

  • Carassus L, Temam E (2014) Pricing and hedging basis risk under no good deal assumption. Ann Financ 10:127–170

    Article  MathSciNet  MATH  Google Scholar 

  • Cerný A (2003) Generalized Sharpe ratios and asset pricing in incomplete markets. Eur Financ Rev 7:191–233

    Article  MATH  Google Scholar 

  • Cerný A, Hodges SD (2002) The theory of good-deal pricing in financial markets. In: Geman H, Madan D, Plinska S, Vorst T (eds) Mathematical finance—Bachelier Congress 2000. Springer, Berlin, pp 175–202

  • Chen Z, Epstein LG (2002) Ambiguity, risk and asset returns in continuous time. Econometrica 70(4):1403–1443

    Article  MathSciNet  MATH  Google Scholar 

  • Cheridito P, Filipović D, Yor M (2005) Equivalent and absolutely continuous measure changes for jump-diffusion processes. Ann Appl Probab 15(3):1713–1732

    Article  MathSciNet  MATH  Google Scholar 

  • Cochrane J, Requejo JS (2000) Beyond arbitrage: good deal asset price bounds in incomplete markets. J Polit Econ 108:79–119

    Article  Google Scholar 

  • Cont R (2006) Model uncertainty and its impact on the pricing of derivative instruments. Math Financ 16(3):519–547

    Article  MathSciNet  MATH  Google Scholar 

  • Coquet F, Hu Y, Mémin J, Peng S (2002) Filtration-consistent nonlinear expectations and related g-expectations. Probab Theory Relat Fields 123(1):1–27

    Article  MathSciNet  MATH  Google Scholar 

  • Delbaen F (2006) The structure of m-stable sets and in particular of the set of risk neutral measures. In: Séminaire de Probabilités XXXIX, Lecture Notes in Mathematics 1874. Springer, Berlin, pp 215–258

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

    Article  MathSciNet  MATH  Google Scholar 

  • Delong Ł (2012) No-good-deal, local mean–variance and ambiguity risk pricing and hedging for an insurance payment process. ASTIN Bull 42(01):203–232

    MathSciNet  MATH  Google Scholar 

  • Donnelly C (2011) Good-deal bounds in a regime-switching diffusion market. Appl Math Financ 18(6):491–515

    Article  MathSciNet  MATH  Google Scholar 

  • Dow J, Werlang S (1992) Uncertainty aversion, risk aversion, and the optimal choice of portfolio. Econometrica 60(1):197–204

  • Drapeau S, Kupper M (2013) Risk preferences and their robust representation. Math Oper Res 38(1):28–62

    Article  MathSciNet  MATH  Google Scholar 

  • Drapeau S, Heyne G, Kupper M (2013) Minimal supersolutions of convex BSDEs. Ann Probab 41(6):3973–4001

    Article  MathSciNet  MATH  Google Scholar 

  • Ekeland I, Temam R (1999) Convex analysis and variational problems. SIAM, Philadelphia

    Book  MATH  Google Scholar 

  • El Karoui N, Quenez M (1995) Dynamic programming and pricing of contingent claims in an incomplete market. SIAM J Control Optim 33(1):29–66

    Article  MathSciNet  MATH  Google Scholar 

  • El Karoui N, Peng S, Quenez MC (1997) Backward stochastic differential equations in finance. Math Financ 1:1–71

    Article  MathSciNet  MATH  Google Scholar 

  • Epstein LG, Schneider M (2003) Recursive multiple-priors. J Econ Theory 113(1):1–31

    Article  MathSciNet  MATH  Google Scholar 

  • Föllmer H, Sondermann D (1986) Hedging of non-redundant contingent claims. In: Hildenbrand W, Mas-Colell A (eds) Contributions to mathematical economics. North-Holland, Amsterdam, pp 205–223

    Google Scholar 

  • Garlappi L, Uppal R, Wang T (2007) Portfolio selection with parameter and model uncertainty: a multi-prior approach. Rev Financ Stud 20(1):41–81

    Article  Google Scholar 

  • Gilboa I, Schmeidler D (1989) Maxmin expected utility with non-unique prior. J Math Econ 18(2):141–153

    Article  MATH  Google Scholar 

  • Gobet E, Turkedjiev P (2016) Linear regression MDP scheme for discrete backward stochastic differential equations under general conditions. Math Comput 85(299):1359–1391

    Article  MathSciNet  MATH  Google Scholar 

  • Hansen LP, Sargent TJ (2001) Robust control and model uncertainty. Am Econ Rev 91:60–66

    Article  Google Scholar 

  • Heath D, Platen E, Schweizer M (2001) A comparison of two quadratic approaches to hedging in incomplete markets. Math Financ 11(4):385–413

    Article  MathSciNet  MATH  Google Scholar 

  • Heston SL (1993) A closed-form solution for options with stochastic volatility, with application to bond and currency options. Rev Financ Stud 6:327–343

    Article  Google Scholar 

  • Jaschke S, Küchler U (2001) Coherent risk measures and good-deal bounds. Finance Stoch 5(2):181–200

    Article  MathSciNet  MATH  Google Scholar 

  • Karatzas I, Shreve SE (2006) Brownian motion and stochastic calculus. Springer, Berlin

    MATH  Google Scholar 

  • Kentia K (2015) Robust aspects of hedging and valuation in incomplete markets and related backward SDE theory. PhD thesis, Humboldt-Universität zu Berlin. urn:nbn:de:kobv:11-100237580

  • Kiseliov YN (1994) Algorithms of projection of a point onto an ellipsoid. Lith Math J 34(2):141–159

    Article  MathSciNet  Google Scholar 

  • Klöppel S, Schweizer M (2007a) Dynamic indifference valuation via convex risk measures. Math Financ 17(4):599–627

    Article  MathSciNet  MATH  Google Scholar 

  • Klöppel S, Schweizer M (2007b) Dynamic utility-based good-deal bounds. Stat Decis 25:311–332

    MathSciNet  MATH  Google Scholar 

  • Leitner J (2007) Pricing and hedging with globally and instantaneously vanishing risk. Stat Decis 25(4):311–332

    MathSciNet  MATH  Google Scholar 

  • Lioui A, Poncet P (2000) The minimum variance hedge ratio under stochastic interest rates. Manag Sci 46(5):658–668

    Article  MATH  Google Scholar 

  • Mahalanobis PC (1936) On the generalized distance in statistics. Proc Natl Inst Sci (Calcutta) 2:49–55

    MATH  Google Scholar 

  • Marroquín-Martínez N, Moreno M (2013) Optimizing bounds on security prices in incomplete markets. Does stochastic volatility specification matter? Eur J Oper Res 225(3):429–442

    Article  MathSciNet  MATH  Google Scholar 

  • Ould-Aly S (2011) Modélisation de la courbe de variance et modèles à volatilité stochastique. PhD thesis, Université Paris-Est. http://tel.archives-ouvertes.fr/tel-00604530/en

  • Pardoux E, Peng S (1990) Adapted solution of a backward stochastic differential equation. Syst Control Lett 14:55–61

    Article  MathSciNet  MATH  Google Scholar 

  • Poulsen R, Schenk-Hopp KR, Ewald CO (2009) Risk minimization in stochastic volatility models: model risk and empirical performance. Quant Financ 9(6):693–704

    Article  MathSciNet  MATH  Google Scholar 

  • Quenez MC (2004) Optimal portfolio in a multiple-priors model. In: Seminar on stochastic analysis, random fields and applications IV, Progress in probability, vol 58. Birkhäuser, Basel, pp 291–321

  • Rockafellar RT (1970) Convex analysis. Princeton University Press, Princeton

    Book  MATH  Google Scholar 

  • Rockafellar RT (1976) Integral functionals, normal integrands and measurable selections. In: Waelbroeck L (ed) Nonlinear operators and calculus of variations, Lecture notes in mathematics 543. Springer, Berlin, pp 157–207

  • Schied A (2008) Robust optimal control for a consumption–investment problem. Math Methods Oper Res 67(1):1–20

    Article  MathSciNet  MATH  Google Scholar 

  • Schweizer M (2001) A guided tour through quadratic hedging approaches. In: Jouini E, Cvitanić J, Musiela M (eds) Option pricing, interest rates and risk management. Cambridge University Press, Cambridge, pp 538–574

    Chapter  Google Scholar 

  • Seifried FT (2010) Optimal investment for worst-case crash scenarios: a martingale approach. Math Oper Res 35(3):559–579

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Support from the German Science Foundation DFG via the Berlin Mathematical School and the Research Training Group 1845 Sto-A is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dirk Becherer.

Appendix

Appendix

This section includes lemmas and derivations omitted from the paper’s main body.

Lemma 6

For \(d< n\), let \(\sigma \in {\mathbb {R}}^{d\times n}\) be of full-rank, \(A\in {\mathbb {R}}^{n\times n}\) be symmetric and positive definite, and \(h>0, Z\in {\mathbb {R}}^n, \xi \in \mathrm {Im}\ \sigma ^{\text {tr}}\). Let \(\alpha '>0\) be a constant of ellipticity of \(A^{-1}\) and assume that \(\left| {\xi } \right| <h\sqrt{\alpha '}\) and \(A^{-1}(\mathrm {Ker}\ \sigma ) = \mathrm {Ker}\ \sigma \). Then \({\bar{\phi }}:= {\varPi }(Z)+\left( {\varPi }^\bot (Z)^{\text {tr}}A^{-1}{\varPi }^\bot (Z)\right) ^{1/2}\left( h^2-\xi ^{\text {tr}}A\xi \right) ^{-1/2}A\xi \) is the unique minimizer of the function \(\phi \mapsto F(\phi ):=-\xi ^*\phi +h\left( \left( Z-\phi \right) ^{\text {tr}}A^{-1}\left( Z-\phi \right) \right) ^{1/2}\) on \(\mathrm {Im}\ \sigma ^{\text {tr}}\).

Proof

The proof is an application of the classical Kuhn–Tucker Theorem (cf. Rockafellar 1970, Sect. 28). For details see Kentia (2015, Lemma 3.35) and proof. \(\square \)

Lemma 7

Let \(d< n, h>0\) be constant, \(Z\in {\mathbb {R}}^n, A\in {\mathbb {R}}^{n\times n}\) a symmetric positive definite matrix, \(\sigma \in {\mathbb {R}}^{d\times n}\) a full (d)-rank matrix, and \(\xi ^0\in {\varPhi } := \text {Im}\ \sigma ^{\text {tr}}\). Let \({\varTheta }\subset {\mathbb {R}}^n\) be a convex-compact set, and \(F: {\mathbb {R}}^n\times {\mathbb {R}}^n\ni (\phi ,\theta )\mapsto \theta ^\text {tr}(Z-\phi )-{\xi ^0}^\text {tr}\phi + h\left( (Z-\phi )^\text {tr}A^{-1}(Z-\phi )\right) ^{1/2}\). Then the minmax identity \( \inf _{\phi \in {\varPhi }}\ \sup _{\theta \in {\varTheta }}F(\phi ,\theta ) = \sup _{\theta \in {\varTheta }}\ \inf _{\phi \in {\varPhi }}F(\phi ,\theta ). \) holds.

Proof

For all \(\phi \in {\mathbb {R}}^n\), the function \(\theta \mapsto F(\phi ,\theta )\) is concave, continuous. For all \(\theta \in {\mathbb {R}}^n\) the function \(\phi \mapsto F(\phi ,\theta )\) is convex and continuous. As \({\varTheta }\subset {\mathbb {R}}^n\) is convex and compact, and \({\varPhi } = \text {Im}\ \sigma ^{\text {tr}}\) is convex and closed, a minimax theorem (Ekeland and Temam 1999, Ch. VI, Prop. 2.3) applies and the minmax identity holds. \(\square \)

Proof (of Lemma 1)

Part (a) is classical (see Delbaen 2006 and cf. previously given other references). As for Part (b), m-stability and convexity of \({\mathcal {M}}^e\) follow from Delbaen (2006, Prop. 5). Convexity of \({\mathcal {Q}}^{\mathrm {ngd}}\) follows from that of \({\mathcal {M}}^e\) and the values of C. To show m-stability of \({\mathcal {Q}}^{\mathrm {ngd}}\), let \(Z^i={\mathcal {E}}(\lambda ^i\cdot W)\in {\mathcal {Q}}^{\mathrm {ngd}},\ i=1,2, \tau \le T\) be a stopping time and \(Z=I_{[0,\tau ]}Z^1_\cdot + I_{]\tau ,T]}Z^1_\tau Z^2_\cdot /Z^2_\tau \). Since \({\mathcal {M}}^e\) is m-stable, then \(Z\in {\mathcal {M}}^e\) and one has \(Z={\mathcal {E}}(\lambda \cdot W)\) for some predictable process \(\lambda \). It remains to show that \(\lambda \) is bounded and that \(\lambda \in C\). From the expression of Z, writing the densities \(Z,Z^1,Z^2\) as ordinary exponentials by distinguishing \(t\le \tau \) and \(t\ge \tau \), and taking the logarithm yields \((\lambda -I_{[0,\tau ]}\lambda ^1-I_{]\tau ,T]}\lambda ^2)\cdot W= \frac{1}{2}\int _0^\cdot \left( \left| {\lambda _s} \right| ^2-I_{[0,\tau ]}(s)\left| {\lambda ^1_s} \right| ^2-I_{]\tau ,T]}(s)\left| {\lambda ^2_s} \right| ^2\right) ds.\) Since \({\mathbb {F}}\) is the augmented Brownian filtration, then \([0,\tau ]\) and \(]\tau ,T]\) are predictable and so is \(\lambda -I_{[0,\tau ]}\lambda ^1-I_{]\tau ,T]}\lambda ^2\). Hence \(\left( \lambda -I_{[0,\tau ]}\lambda ^1-I_{]\tau ,T]}\lambda ^2\right) \cdot W\) is a continuous local martingale of finite variation and is thus equal to zero. As a consequence \(\lambda =I_{[0,\tau ]}\lambda ^1+I_{]\tau ,T]}\lambda ^2\) is bounded since \(\lambda ^1,\lambda ^2\) are, and satisfies \(\lambda \in C\) since C is convex-valued. \(\square \)

Proof (of Theorem 1)

As shown in Kentia (2015, Thm. 3.7), \(\pi ^u_\cdot (X)\) admits under \({\widehat{Q}}\) the Doob–Meyer decomposition \(\pi ^{u}_\cdot (X) = \pi ^{u}_0(X)+Z\cdot {\widehat{W}} - A=\pi ^{u}_0(X)+Z\cdot W + \int _0^\cdot \xi _t^{\text {tr}}Z_tdt- A,\) where \(Z\in {\mathcal {H}}^2({\widehat{Q}})\) and A is a non-decreasing predictable process with \(A_0=0\). Alternatively one rewrites \(\displaystyle -d\pi ^u_t(X) =g_t(Z_t)dt - Z_t^{\text {tr}}dW_t+dK_t,\) with \( K:= A -\int _0^\cdot \xi _t^{\text {tr}}Z_tdt - \int _0^\cdot \mathop {\text{ ess } \text{ sup }}_{\lambda \in {\varLambda }}\lambda _t^{\text {tr}}Z_t\ dt\) being finite-valued and predictable. For \((\pi ^u_\cdot (X),Z,K)\) to be a supersolution to the BSDE with parameters (gX) it suffices to show that K is non-decreasing. For any \(\lambda =-\xi +\eta \in {\varLambda }\), one can construct the sequence of \(\lambda ^n=-\xi +\eta ^n\in {\varLambda }\) Girsanov kernels of measures \(Q^n\in {\mathcal {Q}}^{\mathrm {ngd}}\) with \(\eta ^n=\eta I_{\{|\eta |\le n\}}\) such that \(\lambda ^n\rightarrow \lambda \ P\otimes dt\text {-a.e.}\) as \(n\rightarrow \infty \). For each \(Q^n\) it holds \( \pi ^{u}_\cdot (X) = \pi ^{u}_0(X)+Z\cdot W^{Q^n} + \int _0^\cdot {Z_t}^{\text {tr}}\eta ^n_tdt- A.\) Since \(\pi ^u_\cdot (X)\) is a bounded \(Q^n\)-supermartingale, then \(dA_t - \xi _t^{\text {tr}}Z_tdt \ge Z_t^{\text {tr}}\lambda ^n_tdt,\ \text {for all } n\in {\mathbb {N}}.\) Taking the limit as \(n\rightarrow \infty \) and using dominated convergence one obtains \(dA_t - \xi _t^{\text {tr}}Z_tdt \ge Z_t^{\text {tr}}\lambda _tdt.\) Now taking the essential supremum over all \(\lambda \in {\varLambda }\) yields \(dK_t\ge 0\).

To show that the supersolution \((\pi ^u_\cdot (X),Z,K)\) is minimal, it suffices by the dynamic principle (cf. Kentia 2015, Lem. 3.6) to show that the Y-component of any other supersolution is a càdlàg Q-supermartingale for any \(Q\in {\mathcal {Q}}^{\mathrm {ngd}}\). Let \(({\bar{Y}},{\bar{Z}},{\bar{K}})\) be a supersolution of the BSDE for parameters (gX), with \({\bar{Y}}\in {\mathcal {S}}^\infty \). By change of measure, under a \(Q\in {\mathcal {Q}}^{\mathrm {ngd}}\) with Girsanov kernel \(\lambda ^Q\in {\varLambda }\) we have

$$\begin{aligned}&-d{\bar{Y}}_t =\left( \mathop {\text{ ess } \text{ sup }}_{\lambda \in {\varLambda }}\lambda _t^{\text {tr}}{\bar{Z}}_t - {\bar{Z}}_t^{\text {tr}}\lambda ^Q_t\right) dt - {\bar{Z}}_t^{\text {tr}}dW^Q_t+d{\bar{K}}_t,\quad t\in [0,T], \end{aligned}$$
(68)
$$\begin{aligned}&\text {and get} \quad d{\bar{K}}_t+ \left( \mathop {\text{ ess } \text{ sup }}_{\lambda \in {\varLambda }}\lambda _t^{\text {tr}}{\bar{Z}}_t - {\bar{Z}}_t^{\text {tr}}\lambda ^Q_t\right) dt\ge 0, \quad t\in [0,T], \end{aligned}$$
(69)

by using that \({\bar{K}}\) is non-decreasing. By (6869) and \({\bar{Y}}\in {\mathcal {S}}^\infty \), the local martingale \({\bar{Z}}\cdot W^Q\) is bounded from below, and thus is a supermartingale. As \({\bar{Y}}\in {\mathcal {S}}^\infty \), the integral of (69) on [0, T] is in \(L^1(Q)\) and so \({\bar{Y}}\) is a Q-supermartingale. \(\square \)

Proof (Derivation of (28), (30))

The stochastic exponential \({\mathcal {E}}\left( (\varepsilon /\sqrt{\nu })\cdot W^{\nu }\right) \) is a uniformly integrable martingale which defines a measure \({\bar{Q}}\in \overline{{\mathcal {Q}}^{\text {ngd}}}\supseteq {\mathcal {Q}}^{\text {ngd}}\) (see (4) for definition of \(\overline{{\mathcal {Q}}^{\text {ngd}}}\subset {\mathcal {M}}^e\)) with Girsanov kernel \({\bar{\lambda }}:=\varepsilon /\sqrt{\nu }\), i.e. \(d{\bar{Q}}/dP = {\mathcal {E}}((\varepsilon /\sqrt{\nu })\cdot W^{\nu })\). Indeed, applying Cheridito et al. (2005, Thm. 2.4 and Sect. 6) one gets that \({\mathcal {E}}\left( (\varepsilon /\sqrt{\nu })\cdot W^{\nu }\right) \) and \(S=S_0{\mathcal {E}}\left( \sqrt{\nu }\cdot W^S\right) \) are uniformly integrable P- respectively Q-martingales. The variance process \(\nu \) under \({\bar{Q}}\) is again a CIR process with parameters \(({\bar{a}},b,\beta ,\rho )\) where \({\bar{a}}:=a+\beta \varepsilon \sqrt{1-\rho ^2}>a\) and the Feller condition \(\beta ^2\le 2{\bar{a}}\) still holds. For a put option \(X=(K-S_T)^+ \in L^\infty , {\bar{Y}}_t:=E^{{\bar{Q}}}_t[X]\) are given by the Heston formula (cf. Heston 1993), applied under \({\bar{Q}}\) (instead of P). Since the Heston price is non-decreasing in the mean reversion level of the variance process (Ould-Aly 2011, Prop. 5.3.1) one expects that \(\pi ^u_t\left( X\right) ={\bar{Y}}_t= E^{{\bar{Q}}}_t[X]\). Let us make this precise. For \(Q\in \overline{{\mathcal {Q}}^{\mathrm {ngd}}}\) with Girsanov kernel \(\lambda \) satisfying \(\left| {\lambda } \right| \le \varepsilon /\sqrt{\nu }\), one has \(Y^Q_T={\bar{Y}}_T=X\) with \(Y^Q_t=E^Q_t[X]\). Using Feynman–Kac, \({\bar{Y}}_t=u(t,S_t,\nu _t)\) for a function \(u\in {\mathcal {C}}^{1,2,2}([0,T]\times {\mathbb {R}}^+\times {\mathbb {R}}^+)\) with \(\frac{\partial u}{\partial \nu }\ge 0\) (see Ould-Aly 2011, Thm. 5.3.1, Cor. 5.3.1). By Itô’s formula and change of measure follows

$$\begin{aligned} d{\bar{Y}}_t&= \beta \sqrt{1-\rho ^2}\sqrt{\nu _t}\left( \lambda _t-\frac{\varepsilon }{\sqrt{\nu _t}}\right) \frac{\partial u}{\partial \nu }(t,S_t,\nu _t)dt\nonumber \\ {}&\quad +\beta \sqrt{1-\rho ^2}\sqrt{\nu _t}\frac{\partial u}{\partial \nu }(t,S_t,\nu _t)dW^{Q,\nu }_t\nonumber \\&\quad +\left( S_t\sqrt{\nu _t}\frac{\partial u}{\partial S}(t,S_t,\nu _t)+\beta \rho \sqrt{\nu _t}\frac{\partial u}{\partial \nu }(t,S_t,\nu _t)\right) dW^{S}_t,\quad t\in [0,T]. \end{aligned}$$
(70)

Since X is bounded, then \({\bar{Y}}\) is in \({\mathcal {S}}^\infty (Q)\) and a Q-supermartingale by (70) . Hence \(Y^Q_t \le {\bar{Y}}_t\) for all \(Q\in \overline{{\mathcal {Q}}^{\text {ngd}}}\), which by Part 1 of Kentia (2015, Thm. 3.7) implies the claim and thus we obtain the Heston type formula (28).

Since \({\bar{Q}}\in \overline{{\mathcal {Q}}^{\text {ngd}}}\) and \(\pi ^u_0(X)=E^{{\bar{Q}}}[X]\) with \(X\in L^\infty \), Corollary 1 implies that the good-deal bound is the Y-component of the minimal solution \(({\bar{Y}},{\bar{Z}})\in {\mathcal {S}}^\infty \times {\mathcal {H}}^2 \) (note \(P={\widehat{Q}}\)) of the BSDE (29) with generator \(g_t(z)={\bar{\lambda }}_tz^2=\varepsilon z^2/\sqrt{\nu _t},\) for \(z=(z^1,z^2)\), and terminal condition X. Now consider the strategy

$$\begin{aligned} {\bar{\phi }}_t={\bar{Z}}^1_t=S_t\sqrt{\nu _t}\frac{\partial u}{\partial S}(t,S_t,\nu _t)+\beta \rho \sqrt{\nu _t}\frac{\partial u}{\partial \nu }(t,S_t,\nu _t)=S_t\sqrt{\nu _t}\ {\varDelta }_t+\frac{\beta \rho }{2}{\mathcal {V}}_t. \end{aligned}$$

Clearly \({\bar{\phi }}\) is in the set \({\varPhi } = {\mathcal {H}}^2({\mathbb {R}})\) of permitted trading strategies since \({\bar{Z}}\in {\mathcal {H}}^2 ({\mathbb {R}}^2)\). Recall that \({\mathcal {P}}^{\text {ngd}}\) consists of \(dQ/dP={\mathcal {E}}\left( (\lambda ^S,\lambda ^\nu )\cdot W\right) \) such that \(\big |(\lambda ^S,\lambda ^\nu )\big |\le \varepsilon /\sqrt{\nu }\) with \((\lambda ^S,\lambda ^\nu )\) being bounded. For \(Q\in {\mathcal {P}}^{\text {ngd}}\), any wealth process \(\phi \cdot W^S, \phi \in {\varPhi }\), is thus in \({\mathcal {S}}^1(Q)\). As \({\mathcal {Q}}^{\text {ngd}}\subseteq {\mathcal {P}}^{\text {ngd}}\) holds, clearly \(\pi ^u_t(X) \le \rho _t(X-\int _t^T\phi _sdW^S_s)\) for any strategy \(\phi \in {\varPhi }\). To prove that \({\bar{\phi }}\) is a good-deal hedging strategy, we show the reverse inequality \(\pi ^u_t(X) \ge E^Q_t\left[ X-\int _t^T{\bar{\phi }}_sdW^S_s\right] \) for all \(Q\in {\mathcal {P}}^{\text {ngd}}\). Let \(Q\in {\mathcal {P}}^{\text {ngd}}\) with Girsanov kernel \((\lambda ^S,\lambda ^\nu )\). Like in (70), we obtain for any stopping time \(\tau \) that

$$\begin{aligned} {\bar{Y}}_{\tau \wedge T}- \int _{\tau \wedge t}^{\tau \wedge T}{\bar{\phi }}_sdW^S_s&= {\bar{Y}}_{\tau \wedge t} + \int _{\tau \wedge t}^{\tau \wedge T}\beta \sqrt{1-\rho ^2}\sqrt{\nu _s}\left( \lambda ^\nu _s-\frac{\varepsilon }{\sqrt{\nu _s}}\right) \frac{\partial u}{\partial \nu }(s,S_s,\nu _s)ds\nonumber \\&\quad + L_{\tau \wedge T}-L_{\tau \wedge t}, \end{aligned}$$
(71)

for the local Q-martingale \(L := \int _0^\cdot \beta \sqrt{1-\rho ^2}\sqrt{\nu _s}\frac{\partial u}{\partial \nu }(s,S_s,\nu _s)dW^{Q,\nu }_s\). By \(\frac{\partial u}{\partial \nu }\ge 0\) and \(\big |\lambda ^\nu \big |\le \varepsilon /\sqrt{\nu }\) follows that \( {\bar{Y}}_{\tau \wedge T}- \int _{\tau \wedge t}^{\tau \wedge T}{\bar{\phi }}_sdW^S_s \) is less than \( {\bar{Y}}_{\tau \wedge t} + L_{\tau \wedge T}-L_{\tau \wedge t}\). Localizing L along a sequence of stopping times \(\tau _n\uparrow \infty \) and taking conditional Q-expectations yields \(E^Q_t[{\bar{Y}}_{\tau _n\wedge T}- \int _{\tau _n\wedge t}^{\tau _n\wedge T}{\bar{\phi }}_sdW^S_s] \le {\bar{Y}}_{\tau _n\wedge t}\). Using \(X\in L^\infty \) and \({\bar{\phi }}\cdot W^S\in S^1(Q)\), the claim then follows by dominated convergence. Hence (30) holds for \({\mathcal {V}}_t:=\frac{\partial u}{\partial \sigma }(t,S_t,\nu _t) = 2\sigma _t\frac{\partial u}{\partial \nu }(t,S_t,\nu _t)\) and volatility \(\sigma _t=\sqrt{\nu _t}\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Becherer, D., Kentia, K. Hedging under generalized good-deal bounds and model uncertainty. Math Meth Oper Res 86, 171–214 (2017). https://doi.org/10.1007/s00186-017-0588-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00186-017-0588-y

Keywords

JEL Classification

Navigation