Skip to main content
Log in

Model uncertainty, recalibration, and the emergence of delta–vega hedging

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

We study option pricing and hedging with uncertainty about a Black–Scholes reference model which is dynamically recalibrated to the market price of a liquidly traded vanilla option. For dynamic trading in the underlying asset and this vanilla option, delta–vega hedging is asymptotically optimal in the limit for small uncertainty aversion. The corresponding indifference price corrections are determined by the disparity between the vegas, gammas, vannas and volgas of the non-traded and the liquidly traded options.

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.

Similar content being viewed by others

Notes

  1. Vega is the sensitivity of the Black–Scholes price with respect to changes in the volatility parameter.

  2. Davis [19, Sect. 2. (b)], Musiela and Rutkowski [44, Sect. 7.1.8] and Wilmott [58, Sect. 7.10.5] raise the same concern.

  3. For simplicity, we restrict ourselves to vanilla options in this introduction. Our main result, Theorem 4.5, is also applicable to a wide range of exotic options like barrier options, lookback options, Asian options, forward-start options, and options on the realised variance of the stock.

  4. As is customary in asymptotic analysis, the powers of the processes \(\sigma^{P}\), \(\nu^{P}\), \(\eta^{P}\) and \(\xi^{P}\) in the dynamics of \((S,\Sigma )\) are chosen so that all of them have a nontrivial effect on the leading-order term in the asymptotic expansions below. Using the uncorrelated volatility \(\sqrt{\xi^{P}}\) of implied volatility instead of the uncorrelated squared volatility \(\xi^{P}\) would only generate a higher-order effect in these expansions; cf. Remark 3.2 for the details.

  5. Here, the partial derivatives \(\mathcal{C}_{\Sigma}\), \(\mathcal{C}_{SS}\), \(\mathcal{C}_{S\Sigma}\) and \(\mathcal {C}_{\Sigma \Sigma}\) of \(\mathcal{C}\) are evaluated in \((t,S_{t},\Sigma_{t})\).

  6. The local martingale property of the liquidly traded assets is sufficient to exclude arbitrage opportunities. It also ensures that the agent has no incentive to invest in the market but only uses it as a hedging instrument for the non-traded option; cf. Remark 2.1.

  7. In contrast, most of the literature on hedging under model uncertainty studies variants of the uncertain volatility model introduced by Avellaneda et al. [5] and Lyons [40]. These and many more recent studies (e.g. [24, 21, 46, 49, 9, 47]) look for hedging strategies that dominate the payoff of the non-traded option almost surely for every model of a prespecified class. This worst-case approach corresponds to preferences with infinite risk and uncertainty aversion.

  8. We refer to [28, Sect. 1] for more details on these preferences and their relation to the standard expected utility framework as well as the worst-case approach.

  9. Our analysis also applies to somewhat more general penalty terms; cf. (2.13)–(2.15). The inclusion of the term \(U'(Y^{\boldsymbol{\varpi}}_{t})\) is not crucial but has some appealing properties. For instance, it renders the preferences invariant under affine transformations of the utility function; cf. Remark 2.6 for more details.

  10. Asymptotic analyses of the uncertain volatility model have been carried out by [40, 2, 3, 23].

  11. A second-order expansion and a next-to-leading order optimal strategy are obtained in [28, Theorem 3.4], where only the stock but no additional vanilla option is used for dynamic hedging.

  12. Delta is the sensitivity of a Black–Scholes option value with respect to changes in the price of the underlying.

  13. The vega of the underlying is obviously zero.

  14. Gamma, vanna and volga are the second-order partial derivatives \(\partial^{2}/\partial S^{2}\), \(\partial^{2}/(\partial S\partial\Sigma)\) and \(\partial^{2}/\partial\Sigma^{2}\) of the Black–Scholes value of an option.

  15. In contrast, if there is no liquidly traded call available as a hedging instrument, then the option’s cash gamma is the only greek that appears in the probabilistic representation of the cash equivalent [28].

  16. According to formula (1.6), a short net volga position is only exposed to the part of the volatility of implied volatility that is uncorrelated with the underlying. However, it can be seen from the proof that the correlated volatility of implied volatility has the same effect, albeit only at the order \(O(\psi^{2})\).

  17. See e.g. [52, 19, 20] for precise conditions.

  18. Semi-static hedging problems have also been analysed numerically in the context of the Lagrangian uncertain volatility model [6, 4].

  19. General superhedging duality results in the semi-static context have been obtained, among others, by [1, 7, 22, 25, 10]; see also the references therein.

  20. Other early articles on risk-neutral dynamics for stochastic implied volatility models include [41, 11, 39]. For more recent developments on arbitrage-free market models for (parts of or the whole) option price surface, we refer the reader to [55, 54, 13, 35, 14, 15, 36] and the references therein.

  21. The parametrisation in terms of the squared volatility of implied volatility is explained in Remark 3.2.

  22. For example, [33] find in a Lévy model that the (drift-dependent) variance-optimal hedge is virtually identical to the (drift-independent) Black–Scholes delta hedge.

  23. See Assumption 4.2 for the precise details.

  24. Recall that \(M\) is the running maximum of \(S\) and \(A\) is a general state variable with dynamics of the form (2.8) which can track exotic features of the option like the average stock price or the stock price at an intermediate time; cf. [28, Sect. 4.2] for examples.

  25. For locally bounded, progressively measurable integrands, the stochastic integrals in (2.11) are well defined under each measure in \(\mathfrak{P}^{0}\). The delta–vega hedge considered in our main result, Theorem 4.5, is even continuous.

  26. In [28], only the underlying but no liquid call is available for dynamic hedging, and the spot volatility is the only control variable of the fictitious adversary.

  27. Note that the penalty is imposed on the fictitious adversary who chooses the model \(P\) after the agent has chosen her trading strategy \({\boldsymbol{\varpi}}\). Alternatively, it can be interpreted as a fictitious bonus for the agent.

  28. More general functions \(f\) are considered in [28], where it becomes apparent that only the locally quadratic structure at the minimum matters for the leading-order asymptotics.

  29. Formally, this corresponds to directly imposing the penalty in monetary terms, i.e., inside the utility function in (2.12).

  30. Using \(U'(Y_{0})\) instead of \(U'(Y^{{\boldsymbol{\varpi}},P}_{t})\) would yield the same expansion for \(v(\psi)\) as in Theorem 4.5. Formally, the delta–vega hedge and the candidate optimal controls for the fictitious adversary would still be leading-order optimal. This is because the P&L process converges to a constant in the limit of small uncertainty aversion. Consequently, one could also remove \(U'(Y^{{\boldsymbol{\varpi}},P}_{t})\) from the penalty term by replacing the matrix \(\Psi\) by \(\Psi/U'(Y_{0})\). Then \(U'(Y_{0})\) would reappear in the candidate feedback control for the fictitious adversary and hence also in the cash equivalent \(\widetilde{w}_{0}\). Keeping \(U'(Y^{{\boldsymbol{\varpi}},P}_{t})\) in the penalty term avoids that the candidate optimal controls depend on the current P&L of the agent. This avoids some mathematical subtleties in the formulation of the hedging problem; cf. [28], where the P&L process \(Y\) lives on the canonical space so that (progressively measurable) controls may depend on \(Y\).

  31. In the context of robust portfolio choice, Maenhout [43] also observes that some modification of the standard (non-wealth-dependent) entropic penalty is reasonable to avoid that the agent’s uncertainty aversion wears off as her wealth rises, and tackles this effect by directly modifying the HJBI equation.

  32. In view of [28], it is expected that \(\psi\) (and not e.g. \(\psi^{1/2}\) or \(\psi^{2}\)) is the correct power for the expansion of the value function. Alternatively, one could write \(\psi ^{\alpha}\) instead of \(\psi\) in (3.5) and then find \(\alpha= 1\) by matching the powers of the penalty term and the drift term of the P&L process in the expansion of the HJBI equation in such a way that the optimisation over \(\widetilde{\boldsymbol{\zeta}}\) becomes nontrivial.

  33. Here and in the following, we assume that all relevant partial derivatives of \(\mathcal{C}\) and \(\mathcal{V}\) exist; precise conditions are given in Assumption 4.2 below.

  34. This holds e.g. for a “smooth put”, whose payoff is the Black–Scholes put value with some arbitrarily short maturity.

  35. See also [28, Remark 3.2] for a discussion of such regularity assumptions in a similar setting.

  36. Recall from Remark 4.3 (a) that the delta–vega hedge \({\boldsymbol{\varpi}}^{\star}\) can always be included into the set of trading strategies \(\mathfrak {Y}\) by making the constant \(K_{\mathfrak{Y}}\) from Assumption 4.2 (a) larger if necessary. The existence of a candidate asymptotic model family is discussed in Sect. 4.3.

  37. With a slight abuse of notation, \({\boldsymbol {\varpi}}^{\star}_{t}\) always denotes the time-\(t\) value of the process \({\boldsymbol{\varpi}}^{\star}\) and not the partial derivative of the function \({\boldsymbol{\varpi}}^{\star}\) with respect to the first variable.

  38. Note that the definition of \(H^{\psi}\) already contains the candidate first-order expansion \(w^{\psi}\) of the value function and thus does not feature a general solution function and its derivatives as arguments.

References

  1. Acciaio, B., Beiglböck, M., Penkner, F., Schachermayer, W.: A model-free version of the fundamental theorem of asset pricing and the super-replication theorem. Math. Finance 26, 233–251 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  2. Ahn, H., Muni, A., Swindle, G.: Misspecified asset price models and robust hedging strategies. Appl. Math. Finance 4, 21–36 (1997)

    Article  MATH  Google Scholar 

  3. Ahn, H., Muni, A., Swindle, G.: Optimal hedging strategies for misspecified asset price models. Appl. Math. Finance 6, 197–208 (1999)

    Article  MATH  Google Scholar 

  4. Avellaneda, M., Buff, R.: Combinatorial implications of nonlinear uncertain volatility models: the case of barrier options. Appl. Math. Finance 6, 1–18 (1999)

    Article  MATH  Google Scholar 

  5. Avellaneda, M., Levy, A., Parás, A.: Pricing and hedging derivative securities in markets with uncertain volatilities. Appl. Math. Finance 2, 73–88 (1995)

    Article  Google Scholar 

  6. Avellaneda, M., Parás, A.: Managing the volatility risk of portfolios of derivative securities: the Lagrangian uncertain volatility model. Appl. Math. Finance 3, 21–52 (1996)

    Article  MATH  Google Scholar 

  7. Beiglböck, M., Henry-Labordère, P., Penkner, F.: Model-independent bounds for option prices—a mass transport approach. Finance Stoch. 17, 477–501 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bertsekas, D.: Nonlinear Programming, 2nd edn. Athena Scientific, Belmont (1999)

    MATH  Google Scholar 

  9. Biagini, S., Bouchard, B., Kardaras, C., Nutz, M.: Robust fundamental theorem for continuous processes. Math. Finance (2017), doi:10.1111/mafi.12110. Available online at http://onlinelibrary.wiley.com/doi/10.1111/mafi.12110/full

    Google Scholar 

  10. Bouchard, B., Nutz, M.: Arbitrage and duality in nondominated discrete-time models. Ann. Appl. Probab. 25, 823–859 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Brace, A., Goldys, B., Klebaner, F., Womersley, R.: Market model of stochastic implied volatility with application to the BGM model. Preprint (2001). Available online at http://users.monash.edu.au/~fimak/stochimplvol.pdf

  12. Brown, H., Hobson, D., Rogers, L.C.G.: Robust hedging of barrier options. Math. Finance 11, 285–314 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  13. Carmona, R., Nadtochiy, S.: Local volatility dynamic models. Finance Stoch. 13, 1–48 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  14. Carmona, R., Nadtochiy, S.: Tangent models as a mathematical framework for dynamic calibration. Int. J. Theor. Appl. Finance 14, 107–135 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  15. Carmona, R., Nadtochiy, S.: Tangent Lévy market models. Finance Stoch. 16, 63–104 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  16. Carr, P., Lee, R.: Hedging variance options on continuous semimartingales. Finance Stoch. 14, 179–207 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  17. Cox, A., Obłój, J.: Robust hedging of double touch barrier options. SIAM J. Financ. Math. 2, 141–182 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  18. Cox, A., Obłój, J.: Robust pricing and hedging of double no-touch options. Finance Stoch. 15, 573–605 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  19. Davis, M.: Complete-market models of stochastic volatility. Proc. R. Soc. Lond. A 460, 11–26 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  20. Davis, M., Obłój, J.: Market completion using options. In: Stettner, Ł. (ed.) Advances in Mathematics of Finance, pp. 49–60. Polish Academy of Sciences, Institute of Mathematics, Warsaw (2008)

    Chapter  Google Scholar 

  21. Denis, L., Martini, C.: A theoretical framework for the pricing of contingent claims in the presence of model uncertainty. Ann. Appl. Probab. 16, 827–852 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  22. Dolinsky, Y., Soner, H.M.: Martingale optimal transport and robust hedging in continuous time. Probab. Theory Relat. Fields 160, 391–427 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  23. Fouque, J.-P., Ren, B.: Approximation for option prices under uncertain volatility. SIAM J. Financ. Math. 5, 260–383 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  24. Frey, R.: Superreplication in stochastic volatility models and optimal stopping. Finance Stoch. 4, 161–187 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  25. Galichon, A., Henry-Labordère, P., Touzi, N.: A stochastic control approach to no-arbitrage bounds given marginals, with an application to lookback options. Ann. Appl. Probab. 24, 312–336 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  26. Goodman, J., Ostrov, D.: An option to reduce transaction costs. SIAM J. Financ. Math. 2, 512–537 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  27. Hansen, L., Sargent, T.: Robust control and model uncertainty. Am. Econ. Rev. 91(2), 60–66 (2001)

    Article  Google Scholar 

  28. Herrmann, S., Muhle-Karbe, J., Seifried, F.: Hedging with small uncertainty aversion. Finance Stoch. 21, 1–64 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  29. Hobson, D.: Robust hedging of the lookback option. Finance Stoch. 2, 329–347 (1998)

    Article  MATH  Google Scholar 

  30. Hobson, D., Klimmek, M.: Model-independent hedging strategies for variance swaps. Finance Stoch. 16, 611–649 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  31. Hobson, D., Klimmek, M.: Robust price bounds for the forward starting straddle. Finance Stoch. 19, 189–214 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  32. Hobson, D., Neuberger, A.: Robust bounds for forward start options. Math. Finance 22, 31–56 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  33. Hubalek, F., Kallsen, J., Krawczyk, L.: Variance-optimal hedging for processes with stationary independent increments. Ann. Appl. Probab. 16, 853–885 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  34. Hull, J., White, A.: Hedging the risks from writing foreign currency options. J. Int. Money Financ. 6, 131–152 (1987)

    Article  Google Scholar 

  35. Jacod, J., Protter, P.: Risk-neutral compatibility with option prices. Finance Stoch. 14, 285–315 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  36. Kallsen, J., Krühner, P.: On a Heath–Jarrow–Morton approach for stock options. Finance Stoch. 19, 583–615 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  37. Kallsen, J., Muhle-Karbe, J.: Option pricing and hedging with small transaction costs. Math. Finance 25, 702–723 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  38. Karatzas, I., Shreve, S.: Brownian Motion and Stochastic Calculus, 2nd edn. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  39. Ledoit, O., Santa-Clara, P., Yan, S.: Relative pricing of options with stochastic volatility. Preprint (2002). Available online at http://www.ledoit.net/9-98.pdf

  40. Lyons, T.: Uncertain volatility and the risk-free synthesis of derivatives. Appl. Math. Finance 2, 117–133 (1995)

    Article  Google Scholar 

  41. Lyons, T.: Derivatives as tradable assets. In: Seminario de Matemática Financiera MEFF-UAM, vol. 2, pp. 213–232 (1997)

    Google Scholar 

  42. Maccheroni, F., Marinacci, M., Rustichini, A.: Ambiguity aversion, robustness and the variational representation of preferences. Econometrica 74, 1447–1498 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  43. Maenhout, P.: Robust portfolio rules and asset pricing. Rev. Financ. Stud. 17, 951–983 (2004)

    Article  Google Scholar 

  44. Musiela, M., Rutkowski, M.: Martingale Methods in Financial Modelling, 2nd edn. Springer, Berlin (2005)

    MATH  Google Scholar 

  45. Neuberger, A.: The log contract. J. Portf. Manag. 20, 74–80 (1994)

    Article  Google Scholar 

  46. Neufeld, A., Nutz, M.: Superreplication under volatility uncertainty for measurable claims. Electron. J. Probab. 18(48), 1–14 (2013)

    MathSciNet  MATH  Google Scholar 

  47. Nutz, M.: Superreplication under model uncertainty in discrete time. Finance Stoch. 18, 791–803 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  48. Obłój, J., Ulmer, F.: Performance of robust hedges for digital double barrier options. Int. J. Theor. Appl. Finance 15, 1250003 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  49. Possamaï, D., Royer, G., Touzi, N.: On the robust superhedging of measurable claims. Electron. Commun. Probab. 18(95), 1–13 (2013)

    MathSciNet  MATH  Google Scholar 

  50. Rebonato, R.: Volatility and Correlation, 2nd edn. Wiley, Hoboken (2004)

    Book  MATH  Google Scholar 

  51. Rogers, L.C.G., Williams, D.: Diffusions, Markov Processes and Martingales, vol. 2, 2nd edn. Cambridge University Press, Cambridge (2000)

    Book  MATH  Google Scholar 

  52. Romano, M., Touzi, N.: Contingent claims and market completeness in a stochastic volatility model. Math. Finance 7, 399–412 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  53. Schönbucher, P.: A market model for stochastic implied volatility. Philos. Trans. R. Soc. Lond. A 357, 2071–2092 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  54. Schweizer, M., Wissel, J.: Arbitrage-free market models for option prices: the multi-strike case. Finance Stoch. 12, 469–505 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  55. Schweizer, M., Wissel, J.: Term structures of implied volatilities: absence of arbitrage and existence results. Math. Finance 18, 77–104 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  56. Scott, L.: Random variance option pricing: empirical tests of the model and delta-sigma hedging. Adv. Futures Options Res. 5, 113–135 (1991)

    Google Scholar 

  57. Whalley, A., Wilmott, P.: An asymptotic analysis of an optimal hedging model for option pricing with transaction costs. Math. Finance 7, 307–324 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  58. Wilmott, P.: Paul Wilmott on Quantitative Finance, 2nd edn. Wiley, Hoboken (2006)

    MATH  Google Scholar 

Download references

Acknowledgements

The authors thank Martin Herdegen, David Hobson, Jan Kallsen and Frank Seifried for fruitful discussions, and in particular Martin Schweizer for pertinent remarks on the first draft. Detailed and helpful comments from two anonymous referees are also gratefully acknowledged.

The first author gratefully acknowledges financial support by the Swiss Finance Institute.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sebastian Herrmann.

Appendix: Linearly constrained quadratic programming

Appendix: Linearly constrained quadratic programming

Lagrangian duality. We first recall some basic Lagrange duality results from [8, Sect. 5.1.5]. Fix \(n \in\mathbb{N}\) and functions \(f,g,h:\mathbb{R}^{n} \to\mathbb{R}\). We refer to the problem

$$\begin{aligned} \text{minimise } f(\mathbf{z}) \qquad\text{subject to } \mathbf{z}\in\mathbb{R}^{n}, h(\mathbf{z}) = 0, g(\mathbf{z}) \leq0 \end{aligned}$$
(A.1)

as the primal problem and denote by

$$\begin{aligned} f^{*} &= \inf\lbrace f(\mathbf{z}) : \mathbf{z} \in\mathbb{R}^{n}, h(\mathbf {z}) = 0, g(\mathbf{z}) \leq0 \rbrace \end{aligned}$$

its optimal value. The corresponding Lagrangian is

$$\begin{aligned} L(\mathbf{z}, \lambda, \mu) &= f(\mathbf{z}) + \lambda h(\mathbf{z}) + \mu g(\mathbf{z}), \quad \mathbf{z} \in\mathbb{R}^{n}, \lambda,\mu\in\mathbb{R}, \end{aligned}$$

and a pair \((\mu^{*}, \lambda^{*})\) is called a Lagrange multiplier if

$$\begin{aligned} f^{*} &= \inf_{\mathbf{z} \in\mathbb{R}^{n}} L(\mathbf{z},\lambda^{*},\mu^{*}) \quad\text{and}\quad \mu^{*} \geq0. \end{aligned}$$

The dual problem for (A.1) is

$$\begin{aligned} \text{maximise } q(\lambda,\mu) \qquad\text{subject to } \lambda\in\mathbb{R}, \mu\geq0, \end{aligned}$$

where the dual function \(q\) is

$$\begin{aligned} q(\lambda,\mu) &= \inf_{\mathbf{z} \in\mathbb{R}^{n}} L(\mathbf{z},\lambda,\mu), \quad \lambda,\mu\in\mathbb{R}. \end{aligned}$$

Finally, \(q^{*} = \sup\lbrace q(\lambda,\mu):\lambda\in\mathbb{R}, \mu \geq0 \rbrace\) denotes the optimal value of the dual problem.

A quadratic programming problem with linear equality and inequality constraints. The following lemma provides the solution to a primal problem with a strictly convex quadratic cost function and specific linear equality and inequality constraints.

Lemma A.1

Fix \(n \in\mathbb{N}\), a diagonal matrix \(D = \operatorname {diag}(d_{1},\ldots,d_{n}) \in\mathbb{R}^{n\times n}\) with positive diagonal entries, and vectors \(\mathbf{v} = (v_{1},\ldots,v_{n})^{\top}\) and \(\mathbf{c} = (c_{1},\ldots, c_{n})^{\top}\) in \(\mathbb{R}^{n}\) such that \(c_{i} \neq0\) for some \(i \in\lbrace1,\ldots,n-1\rbrace\). Moreover, set \(\mathbf{1}_{A} = 1\) if \(v_{n} - \frac{\mathbf{c}^{\top}D^{-1}\mathbf {v}}{\mathbf{c}^{\top}D^{-1}\mathbf{c}} c_{n} < 0\) and \(\mathbf{1}_{A} = 0\) otherwise, and define

$$\begin{aligned} \lambda^{*} &= \frac{\mathbf{c}^{\top}D^{-1}\mathbf{v} - c_{n} v_{n} d_{n}^{-1} \mathbf {1}_{A}}{\mathbf{c}^{\top}D^{-1}\mathbf{c} - c_{n}^{2} d_{n}^{-1}\mathbf {1}_{A}}, \end{aligned}$$
(A.2)
$$\begin{aligned} \mu^{*} &= (v_{n} - \lambda^{*} c_{n})^{-}, \end{aligned}$$
(A.3)
$$\begin{aligned} \mathbf{z}^{*} &= D^{-1}(\mathbf{v} - \lambda^{*} \mathbf{c} + \mu^{*} \mathbf{e}_{n}). \end{aligned}$$
(A.4)
  1. (a)

    \(\mathbf{z}^{*}\) is the unique optimiser of the primal problem

    $$\begin{aligned} \textit{minimise }\frac{1}{2}\mathbf{z}^{\top}D \mathbf{z} - \mathbf {v}^{\top}\mathbf{z} \qquad\textit{ subject to } \mathbf{z}\in\mathbb{R}^{n}, \mathbf{c}^{\top}\mathbf{z} = 0, z_{n} \geq0 \end{aligned}$$
    (A.5)

    and satisfies the bound \(\left\vert\mathbf{z}^{*}\right\vert \leq d_{\min}^{-1} \left\vert\mathbf{v}\right\vert\), where \(d_{\min} = \min(d_{1},\ldots,d_{n})\).

  2. (b)

    \((\lambda^{*},\mu^{*})\) is the unique optimiser of the dual problem for (A.5), which can be written as

    $$\begin{aligned} &\textit{maximise } -\frac{1}{2}(\mathbf{v} - \lambda\mathbf{c} + \mu \mathbf{e}_{n})^{\top}D^{-1}(\mathbf{v} - \lambda\mathbf{c} + \mu \mathbf{e}_{n}) \\ & \textit{subject to } \lambda\in\mathbb{R}, \mu\geq0. \end{aligned}$$
    (A.6)

    The optimiser satisfies the bound

    $$\begin{aligned} \vert\lambda^{*} \mathbf{c} - \mu^{*} \mathbf{e}_{n}\vert &\leq\left(1 + \frac{d_{\max}}{d_{\min}}\right)\left\vert \mathbf{v}\right\vert, \end{aligned}$$
    (A.7)

    where \(d_{\max} = \max(d_{1},\ldots,d_{n})\).

  3. (c)

    The optimal values of the primal and dual problems coincide (i.e., there is no duality gap) and equal

    $$\begin{aligned} -\frac{1}{2}\mathbf{v}^{\top}\mathbf{z}^{*}. \end{aligned}$$
  4. (d)

    The triplet \((\mathbf{z}^{*},\mu^{*},\lambda^{*})\) satisfies the optimality conditions

    $$\begin{aligned} \mathbf{z}^{*} = \arg\min_{\mathbf{z}\in\mathbb{R}^{n}} L(\mathbf{z},\lambda ^{*},\mu ^{*}),\quad \mathbf{c}^{\top}\mathbf{z}^{*} = 0,\quad z^{*}_{n} \geq0,\quad \mu^{*} \geq0,\quad \mu^{*}z_{n}^{*} = 0, \end{aligned}$$
    (A.8)

    where \(L\) is the Lagrangian corresponding to the primal problem. Moreover, the pair \((\lambda^{*},\mu^{*})\) is a Lagrange multiplier for the primal problem.

Proof

First of all, note that the Lagrangian

$$\begin{aligned} L(\mathbf{z},\lambda,\mu) &= \frac{1}{2}\mathbf{z}^{\top}D \mathbf{z} - \mathbf{v}^{\top}\mathbf{z} + \lambda\mathbf{c}^{\top}\mathbf{z} - \mu \mathbf{e}_{n}^{\top}\mathbf{z} \end{aligned}$$

corresponding to the primal problem is strictly convex over \(\mathbf{z} \in\mathbb{R}^{n}\). Hence, the dual function \(q(\lambda,\mu)= \inf _{\mathbf{z}\in\mathbb{R}^{n}} L(\mathbf{z},\lambda,\mu)\) can be computed explicitly by substituting the solution \(\mathbf{z}'\) to the first-order condition \(\mathbf{0} = \operatorname{D}_{\mathbf{z}} L(\mathbf{z}',\lambda,\mu) = D \mathbf{z}' - \mathbf{v} + \lambda \mathbf{c} - \mu \mathbf{e}_{n}\) back into \(L(\mathbf{z}',\mu ,\lambda)\). This yields

$$\begin{aligned} q(\lambda,\mu) &= -\frac{1}{2}(\mathbf{v} - \lambda\mathbf{c} + \mu \mathbf {e}_{n})^{\top}D^{-1}(\mathbf{v} - \lambda\mathbf{c} + \mu \mathbf{e}_{n}), \end{aligned}$$

and thus the dual problem takes the form (A.6).

The crucial part of the proof is to show that the triplet \((\mathbf {z}^{*},\lambda^{*},\mu^{*})\) satisfies the optimality conditions (A.8). As \(L\) is strictly convex over \(\mathbf{z} \in\mathbb{R}^{n}\), the optimality conditions are equivalent to

$$\begin{aligned} D\mathbf{z}^{*} - \mathbf{v} + \lambda^{*} \mathbf{c} - \mu^{*} \mathbf {e}_{n} = 0,\quad \mathbf{c}^{\top}\mathbf{z}^{*} = 0,\quad z^{*}_{n} \geq0,\quad \mu^{*} \geq0,\quad \mu^{*} z^{*}_{n} = 0. \end{aligned}$$
(A.9)

Recall the definitions of \(\lambda^{*}\), \(\mu^{*}\) and \(\mathbf{z}^{*}\) in (A.2)–(A.4) and note that the assumption that \(c_{i} \neq0\) for some \(i\in\lbrace 1,\ldots ,n-1\rbrace\) together with the positive definiteness of \(D^{-1}\) ensures that \(\lambda^{*}\) is well defined. The stationarity condition \(D\mathbf{z}^{*} - \mathbf{v} + \lambda^{*} \mathbf{c} - \mu^{*} \mathbf {e}_{n} = 0\) holds by definition of \(\mathbf{z}^{*}\). For the other conditions, we distinguish two cases. First, suppose that \(\mathbf{1}_{A} = 0\), i.e., \(v_{n} - \frac{\mathbf {c}^{\top}D^{-1}\mathbf{v}}{\mathbf{c}^{\top}D^{-1} \mathbf{c}} c_{n} \geq 0\). Then \(\lambda^{*} = \frac{\mathbf{c}^{\top}D^{-1}\mathbf {v}}{\mathbf {c}^{\top}D^{-1} \mathbf{c}}\), \(\mu^{*} = 0\) and \(z^{*}_{n} = d_{n}^{-1}(v_{n} - \lambda^{*} c_{n}) \geq0\). Moreover, \(\mathbf{c}^{\top}\mathbf{z}^{*} = \mathbf{c}^{\top}D^{-1} \mathbf{v} - \lambda^{*} \mathbf{c}^{\top}D^{-1} \mathbf{c} = 0\). Second, suppose that \(\mathbf{1}_{A} = 1\), i.e., \(v_{n} - \frac{\mathbf {c}^{\top}D^{-1}\mathbf{v}}{\mathbf{c}^{\top}D^{-1} \mathbf{c}} c_{n} < 0\), or, equivalently (multiply by \(\mathbf{c}^{\top}D^{-1}\mathbf{c}\), add and subtract \(c_{n}^{2} v_{n} d_{n}^{-1}\), and then divide by \(\mathbf{c}^{\top}D^{-1} \mathbf{c} - c_{n}^{2} d_{n}^{-1} > 0\)), \(v_{n} -\lambda^{*} c_{n} < 0\). Then \(\mu^{*} > 0\) and \(z^{*}_{n} = 0\) by the definition of \(\mu^{*}\) and \(\mathbf{z}^{*}\). Finally, setting \(c = \mathbf{c}^{\top}D^{-1} \mathbf {c}\) and \(v = \mathbf{c}^{\top}D^{-1} \mathbf{v}\) for brevity, we obtain

$$\begin{aligned} \mathbf{c}^{\top}\mathbf{z}^{*} &= \mathbf{c}^{\top}D^{-1}(\mathbf{v} - \lambda^{*}\mathbf{c}+ \mu ^{*}\mathbf{e}_{n}) = v -\lambda^{*} c +\mu^{*} c_{n} d_{n}^{-1}\\ &= v -\lambda^{*} c - (v_{n} - \lambda^{*} c_{n}) c_{n} d_{n}^{-1} = v - \lambda^{*}(c - c_{n}^{2} d_{n}^{-1}) - c_{n} v_{n} d_{n}^{-1}\\ &= v - (v - c_{n} v_{n} d_{n}^{-1}) - c_{n} v_{n} d_{n}^{-1} = 0. \end{aligned}$$

So, (A.9) holds in both cases. By the characterisation of primal optimal solutions [8, Proposition 5.1.5], this implies that \(\mathbf{z}^{*}\) is an optimiser for the primal problem, that \((\lambda^{*},\mu^{*})\) is a Lagrange multiplier, and that there is no duality gap. Moreover, \((\lambda^{*},\mu^{*})\) is an optimiser for the dual problem by a corollary [8, Proposition 5.1.4 (a)] of the weak duality theorem [8, Proposition 5.1.3]. As the primal and dual problems are strictly convex and strictly concave, respectively, the optimisers are unique.

Plugging the optimiser \((\lambda^{*},\mu^{*})\) of the dual problem into the cost function of the dual problem (A.6) and using the definition of \(\mathbf{z}^{*}\), the optimal value \(q^{*}\) (of both the primal and the dual problem) reads

$$\begin{aligned} q^{*}&:=-\frac{1}{2}(\mathbf{v} - \lambda^{*} \mathbf{c} + \mu^{*} \mathbf{e}_{n})^{\top}D^{-1}(\mathbf{v} - \lambda^{*} \mathbf{c} + \mu \mathbf{e}_{n}) = -\frac{1}{2} (\mathbf{v} - \lambda^{*} \mathbf{c} + \mu^{*} \mathbf {e}_{n})^{\top}\mathbf{z}^{*}. \end{aligned}$$

Now, note that \(\mathbf{c}^{\top}\mathbf{z}^{*} = 0\) and \(\mu^{*} \mathbf{e}_{n}^{\top}\mathbf{z}^{*} = \mu^{*} z_{n}^{*} = 0\) by (A.9). Hence, \(q^{*} = -\frac{1}{2}\mathbf {v}^{\top}\mathbf{z}^{*}\).

Next, using that \(\mathbf{z}^{*}\) achieves the optimal value \(-\frac {1}{2}\mathbf{v}^{\top}\mathbf{z}^{*}\) for the primal problem and applying the Cauchy–Schwarz inequality, we obtain

$$\begin{aligned} \frac{1}{2}d_{\min} \vert\mathbf{z}^{*}\vert^{2} &\leq\frac{1}{2}(\mathbf{z}^{*})^{\top}D \mathbf{z}^{*} = \frac{1}{2}\mathbf{v}^{\top}\mathbf{z}^{*} \leq\frac{1}{2} \vert\mathbf{v}\vert \vert\mathbf{z}^{*}\vert. \end{aligned}$$

This yields the last claim of part (a). Finally, using (A.4), the triangle inequality and the bound \(\left\vert\mathbf{z}^{*}\right\vert \leq d_{\min}^{-1}\left\vert \mathbf{v}\right\vert\) which we just proved, we obtain

$$\begin{aligned} \vert\lambda^{*} \mathbf{c} - \mu^{*} \mathbf{e}_{n}\vert &= \vert D\mathbf{z}^{*} - \mathbf{v}\vert \leq\vert D\mathbf{z}^{*}\vert + \vert\mathbf{v}\vert \leq d_{\max} \vert\mathbf{z}^{*}\vert + \vert\mathbf{v}\vert \leq\bigg(1+\frac{d_{\max}}{d_{\min}}\bigg) \vert\mathbf {v}\vert. \end{aligned}$$

This proves the last claim of part (b) and thereby concludes the proof. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Herrmann, S., Muhle-Karbe, J. Model uncertainty, recalibration, and the emergence of delta–vega hedging. Finance Stoch 21, 873–930 (2017). https://doi.org/10.1007/s00780-017-0342-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-017-0342-6

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation