Skip to main content

On Small-Noise Equations with Degenerate Limiting System Arising from Volatility Models

  • Conference paper
  • First Online:
Book cover Large Deviations and Asymptotic Methods in Finance

Abstract

The one-dimensional SDE with non Lipschitz diffusion coefficient

$$\begin{aligned} \textit{dX}_{t} = b(X_{t})\textit{dt} + \sigma X_{t}^{\gamma } \textit{dB}_{t}, \quad X_{0}=x, \quad \gamma <1 \end{aligned}$$
(1)

  is widely studied in mathematical finance. Several works have proposed asymptotic analysis of densities and implied volatilities in models involving instances of (1), based on a careful implementation of saddle-point methods and (essentially) the explicit knowledge of Fourier transforms. Recent research on tail asymptotics for heat kernels (Deuschel et al. Comm. in Pure and Applied Math., 67(1):40–82, 2014, [11]) suggests to work with the rescaled variable \( X^{\varepsilon }:=\varepsilon ^{1/(1-\gamma )} X\): while allowing to turn a space asymptotic problem into a small-\(\varepsilon \) problem, the process \(X^{\varepsilon }\) satisfies a SDE in Wentzell–Freidlin form (i.e. with driving noise \(\varepsilon \textit{dB}\)). We prove a pathwise large deviation principle for the process \(X^{\varepsilon }\) as \(\varepsilon \rightarrow 0\). As it will be seen, the limiting ODE governing the large deviations admits infinitely many solutions, a non-standard situation in the Wentzell–Freidlin theory. As for applications, the \(\varepsilon \)-scaling allows to derive leading order asymptotics for path functionals: while on the one hand the resulting formulae are confirmed by the CIR-CEV benchmarks, on the other hand the large deviation approach (i) applies to equations with a more general drift term and (ii) potentially opens the way to heat kernel analysis for higher-dimensional diffusions involving (1) as a component.

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 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
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

Notes

  1. 1.

    The precise statement here is .

  2. 2.

    When \(\beta =0\), \(\gamma =1/2\) and \(\dot{h}\equiv 1\), one retrieves the textbook example of ODE for which uniqueness fails, \(\dot{\varphi }_{t} = \sigma \sqrt{|\varphi _t|}\), whose solutions from \(\varphi _0=0\) are given by the one-parameter family \(\varphi ^{(\theta )}_t = \frac{\sigma ^2}{4} (t-\theta )^2 1_{\{t \ge \theta \}}\).

  3. 3.

    When \(\gamma \in [1/2,1)\), the law of \(X_T\) also possesses an atom at zero, \(\mathbb {P}(X_T=0)=m_T>0\), and an explicit formula for the mass \(m_T\) is available (see again [16, Chap. 6]). From our point of view, this only means that the density \(f_{X_T}\) does not integrate to 1 on \((0,\infty )\), without affecting our analysis of the tail asymptotics at \(\infty \).

  4. 4.

    The second derivative reads \(e^{a \varepsilon ^{-2} (1+y)^{2(1-\gamma )}} \times 2a\varepsilon ^{-2}(1-\gamma )(1+y)^{-2\gamma } \times [1-2\gamma +\frac{2a}{\varepsilon ^2}(1-\gamma )\) \((1+y)^{2(1-\gamma )}]\).

  5. 5.

    By perturbing the initial condition and the drift in (1.2), one can retrieve the trajectory \(\varphi ^*\) in (3.29) as the limit as \(\rho \rightarrow 0\) of the solution of the equation \(d\varphi _{t} = \rho \,+\,\beta \varphi _t \textit{dt}\,+\,\sigma \varphi _t^{\gamma }dh, \varphi _{0} = \rho \), for which existence and uniqueness hold.

References

  1. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 10th edn. Dover, New York (1972)

    MATH  Google Scholar 

  2. Andersen, L., Piterbarg, V.: Moment explosions in stochastic volatility models. Financ. Stoch. 11, 29–50 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  3. Azencott, R.: Grandes déviations et applications. Ecole d’été de Probabilités de Saint-Flour VIII-1978. Lecture Notes in Mathematics, vol. 774, pp. 1–176. Springer, Berlin (1980)

    Chapter  Google Scholar 

  4. Baldi, P., Caramellino, L.: General Freidlin-Wentzell large deviations and positive diffusions. Stat. Probab. Lett. 81, 1218–1229 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  5. Ben Arous, G.: Développement asymptotique du noyau de la chaleur hypoelliptique hors du cut-locus. Annales scientifiques de l’Ecole Normale Supérieure 4(21), 307–331 (1988)

    MathSciNet  Google Scholar 

  6. Bingham, N.H., Goldie, C.M., Teugels, J.L.: Regular Variation. Cambridge University Press, Cambridge (1987)

    Book  MATH  Google Scholar 

  7. Bismut, J.-M.: Large Deviations and the Malliavin Calculus. Birkhäuser, Boston (1984)

    MATH  Google Scholar 

  8. Chiarini, A., Fischer, M.: On large deviations for small noise Itô processes. Adv. Appl. Probab. 46(4), 1126–1147 (2014)

    Google Scholar 

  9. De Marco, S.: Smoothness and asymptotic estimates of densities for SDEs with locally smooth coefficients and applications to square root-type diffusions. Ann. Appl. Probab. 4(21), 1282–1321 (2011)

    Article  Google Scholar 

  10. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications, 2nd edn. Applications of Mathematics, Springer, New York (1998)

    Google Scholar 

  11. Deuschel, J.-D., Friz, P., Jacquier, A., Violante, S.: Marginal density expansions for diffusions and stochastic volatility, II: Applications. Comm. in Pure and Applied Math. 67(1), 40–82 (2014)

    Google Scholar 

  12. Deuschel, J-D., Stroock., W.: Large Deviations. Pure and Applied Mathematics. American Mathematical Society, New York, London. Revised edition of: An introduction to the theory of large deviations/D.W. Stroock. cop.1984 (2000)

    Google Scholar 

  13. Donati-Martin, C., Rouault, A., Yor, M., Zani, M.: Large deviations for squares of Bessel and Ornstein-Uhlenbeck processes. Probab. Theory Relat. Fields 129, 261–289 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  14. Dufresne, D.: The integrated square-root process. Research Paper no. 90, Centre for Actuarial Studies, University of Melbourne (2001)

    Google Scholar 

  15. Gulisashvili, A.: Asymptotic formulas with error estimates for call pricing functions and the implied volatility at extreme strikes. SIAM J. Financ. Math. 1(1), 609–641 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  16. Jeanblanc, M., Yor, M., Chesney, M.: Mathematical Methods for Financial Markets. Springer finance. Springer, Dordrecht, Heidelberg, London (2009)

    Google Scholar 

  17. Karatzas, I., Shreve, S.: Brownian Motion and Stochastic Calculus, 2nd edn. Springer, New York (1991)

    MATH  Google Scholar 

  18. Keller-Ressel, M.: Moment explosions and long-term behavior of affine stochastic volatility models. Math. Financ. 21, 73–98 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  19. Klebaner, F., Liptser, R.: Asymptotic analysis of ruin in the constant elasticity of variance model. Theory Probab. Appl. 55(2), 291–297 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  20. Leoni, G.: A First Course in Sobolev Spaces. Graduate studies in mathematics. American Mathematical Society, Cambridge (2009)

    Book  MATH  Google Scholar 

  21. Lions, P.-L., Musiela, M.: Correlations and bounds for stochastic volatility models. Annales de l’Institut H. Poincaré 24, 1–16 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  22. Revuz, D., Yor, M.: Continuous Martingales and Brownian Motion, 3rd edn. Springer, New York (1999)

    Book  MATH  Google Scholar 

  23. Robertson, S.: Sample path large deviations and optimal importance sampling for stochastic volatility models. Stoch. Process. Appl. 120(1), 66–83 (2010)

    Article  MATH  Google Scholar 

  24. Schöbel, R., Zhu, J.: Stochastic volatility with an Ornstein-Uhlenbeck process: an extension. Eur. Financ. Rev. 3(1), 23–46 (1999)

    Article  MATH  Google Scholar 

  25. Stein, E.M., Stein, J.C.: Stock price distribution with stochastic volatility: an analytic approach. Rev. Financ. Stud. 4, 727–752 (1991)

    Article  Google Scholar 

  26. Stroock, D., Varadhan, S.R.S.: Multidimensional Diffusion Processes. Grundlehren der mathematischen Wissenschaften. Fundamental Principles of Mathematical Sciences, vol. 233. Springer, Berlin (1979). Reprinted in 2006

    MATH  Google Scholar 

  27. Trevisan, D.: Zero noise limits using local times. Electron. Commun. Probab. 18(31), 1–7 (2013)

    MathSciNet  Google Scholar 

Download references

Acknowledgments

We would like to thank an anonymous referee for the careful reading of the paper and for several valuable comments which helped to improve the presentation. We thank Peter Friz for stimulating discussions and Antoine Jacquier for useful references on integrated CIR processes. SDM (affiliated with TU-Berlin when this work was started) acknowledges partial financial support from Matheon. GC acknowledges financial support from Berlin Mathematical School. SDM and GC acknowledge financial support for travel expenses from the research program ‘Chaire Risques Financiers’ of the Fondation du Risque.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stefano De Marco .

Editor information

Editors and Affiliations

Appendix A

Appendix A

We complete the proof of Proposition 3.2 here.

Proof of Proposition 3.2 Let us define an auxiliary process \(\overline{X}\) by

$$\begin{aligned} d\overline{X}_{t} = |\alpha |_{\infty }\textit{dt} +\sigma \exp (-(1-\gamma ) |\beta | t ) \overline{X}_{t}^{\gamma }\textit{dB}_{t}, \quad \overline{X}_{0}=x; \end{aligned}$$

after a simple application of the product rule, one has that the process \(Z_{t} := \exp (|\beta | t)\overline{X}_{t} \) is a solution to

$$\begin{aligned} dZ_{t} = \bigl (|\alpha |_{\infty } \exp (|\beta | t) + |\beta | Z_{t}\bigr ) \textit{dt} + \sigma Z_{t}^{\gamma }\textit{dB}_{t}, \quad Z_{0} = x. \end{aligned}$$

Since \(|\alpha |_{\infty } \exp (|\beta | t) \ge |\alpha |_{\infty }\), an application of the comparison principle for SDE’s [17, Proposition 5.2.18] yields \(Z_{t} \ge \tilde{X}_{t}\), for all \(t \ge 0\). Therefore, if \(\overline{X}^{2(1-\gamma )}\) admits (some) exponential moments, so does \(Z^{2(1-\gamma )}_{t}\) and by comparison \(\tilde{X}^{2(1-\gamma )}_{t}\). In this sense, the process \(\overline{X}\) is not covered by Proposition 3.3 in [9], since the latter deals with the case of a diffusion coefficient that does not depend on time (see [9, Eq. (3.1)]); nonetheless, the essential condition that [9, Proposition 3.3] relies on is the presence of a non-strictly positive slope coefficient, say b in the drift term \(a+bX\) (cf. [9, Eq. (3.3)]). Since this is the case for the process \(\overline{X}\) (which has zero slope coefficient b), it is straightforward to extend the proof to the present setting: in particular, in the spirit of Lamperti’s change-of-variable argument, one still defines the function \(\varphi (x)=\int _0^{x}\frac{1}{\sigma x^{\gamma }}=\frac{1}{\sigma (1-\gamma )}x^{1-\gamma }\) and studies the process \(\tilde{\varphi }(X_t)\), where the function \(\tilde{\varphi }\) is a modification of \(\varphi \) identically null around zero. Itô’s formula shows that \(\tilde{\varphi }(X_t)\) is an Itô process with bounded quadratic variation and a bounded drift term; the existence of quadratic exponential moments for \(\tilde{\varphi }(X_t)\), then, is a consequence of Dubins–Schwarz time-change argument and Fernique’s theorem. As a consequence, there exist \(c',C > 0\) such that \(\sup _{t \le T} \mathbb {E}[\exp (c' \overline{X}^{2(1-\gamma )}_{t})] \le C\); it follows \(\sup _{t \le T} \mathbb {E}[\exp (c \tilde{X}^{2(1-\gamma )}_{t})] \le \sup _{t \le T} \mathbb {E}[\exp (c Z^{2(1-\gamma )}_{t})] \le C\) with \(c:=c' \exp (-2|\beta |(1-\gamma )T)\), and the claim is proved.\(\blacksquare \)

We report the statement given in [26, Chap. 2, Theorem 2.13].

Lemma A.1

(Garsia-Rodemich-Rumsey’s Lemma) Let p and \(\Psi \) be continuous, strictly increasing functions on \([0, +\infty )\) such that \(p(0) = \Psi (0)=0\) and \(\lim _{t \rightarrow + \infty } \Psi (t)= + \infty \). If \(\omega \in \Omega \) is such that:

$$\begin{aligned} \int _{0}^{T} \int _{0}^{T} \Psi \left( \frac{|\omega _{t} - \omega _{s} |}{p(|t-s|)}\right) \textit{ds} \textit{dt} \le K, \end{aligned}$$
(A.1)

then

$$\begin{aligned} |\omega _{t} - \omega _{s} | \le 8 \int _{0}^{|t-s|} \Psi ^{-1} \left( \frac{4K}{u^2}\right) dp(u). \end{aligned}$$
(A.2)

Lemma A.1 allows us to prove Proposition 3.3:

Proof of Proposition 3.3 Assume that (3.3) holds true with the left hand side replaced by \(K>0\). Applying Lemma A.1 with the choice of functions \(\Psi (y)=\exp (\varepsilon ^{-2} y)-1, p(y) = \sqrt{y}\), one has for all st

$$\begin{aligned} \left| \omega _{t}- \omega _{s} \right| \le 8\int _{0}^{|t-s|} \Psi ^{-1} \left( \frac{4K}{u^2} \right) dp(u)&= 8\varepsilon ^{2}\int _{0}^{|t-s|} \log \left( \frac{4K}{u^2} +1 \right) dp(u) \\&\le 8 \varepsilon ^{2} \left[ \int _{0}^{|t-s|} \log \left( 4K + T^2 \right) dp(u)\right. \\&\qquad \qquad \left. + \int _{0}^{|t-s|} \log \left( u^{-2} \right) dp(u) \right] \\&\le 8 \varepsilon ^{2}\left[ \sqrt{|t-s|} \log \left( 4K+T^{2} \right) \right. \\&\qquad \qquad \left. +\,\sqrt{|t-s|} \left( 4-2\log \left( |t-s|\right) \right) \right] . \end{aligned}$$

Dividing on both sides by \((t-s)^{\eta }\) and taking suprema we obtain

$$\begin{aligned} \left\| \omega \right\| _{\eta } \le 8 \varepsilon ^{2} \left( \log \left( 4K+T^{2} \right) T^{1/2-\eta } +4T^{1/2-\eta } + K_{\eta } \right) . \end{aligned}$$

Since the right hand side in the last estimate is \(K^{-1}_{\varepsilon , \eta } \left( K \right) \), (3.3) yields (3.4).\(\blacksquare \)

Finally, we prove Lemma 3.10.

Proof of Lemma 3.10 Denote \(T^{\varepsilon }\) the stopping time

$$\begin{aligned} T^{\varepsilon } (\omega ) = \inf \left\{ t \ge 0 : \omega _{t} \le \frac{1}{2} \varepsilon x^{1-\gamma } \right\} . \end{aligned}$$
(A.3)

We can apply Itô formula to the function \(f(x) = x^{1-\gamma }\) up to time \(T^{\varepsilon }(Y^{\varepsilon ,h})\), and obtain

$$\begin{aligned} Y^{\varepsilon ,h}_{t}\,-\,\varepsilon x^{1-\gamma }= \int _{0}^{t} \tilde{b}^{\varepsilon } (Y^{\varepsilon ,h}_{s})\textit{ds}\,+\,\sigma (1\,-\,\gamma ) h_{t}\,+\,\varepsilon \sigma (1\,-\,\gamma ) B_t, \quad \forall \, t \le T^{\varepsilon }(Y^{\varepsilon ,h}), \quad a.s. \end{aligned}$$
(A.4)

where \(\tilde{b}^{\varepsilon }\) is given by

$$\begin{aligned} \tilde{b}_{\varepsilon }(y) := (1-\gamma )\varepsilon ^{\frac{1}{1-\gamma }}\alpha (\varepsilon ^{-\frac{1}{(1-\gamma )}}y^{\frac{1}{(1-\gamma )}})\frac{1}{y^{\frac{\gamma }{1-\gamma }}} - \frac{\sigma ^{2} \gamma (1-\gamma )}{2}\varepsilon ^{2} \frac{1}{y} + \beta (1-\gamma ) y\quad \end{aligned}$$
(A.5)

We need to prove

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0 } W \left( \sup _{t \in [0,T]} | Y^{\varepsilon ,h}_{t}- \mathcal {S}_{0}(h)_{t}| \le R \right) = 1 \qquad \forall R>0. \end{aligned}$$
(A.6)

In order to simplify the notation, there is no ambiguity in writing Y instead of \(Y^{\varepsilon ,h}\) inside this proof.

Step 1. We first prove (A.6) under the assumption

$$\begin{aligned} k:=\inf _{t \in [0,T]}\dot{h}_{t} >0 \end{aligned}$$
(A.7)

Let us fist show that

$$\begin{aligned} \lim _{ \varepsilon \rightarrow 0} W \left( T^{\varepsilon } \left( Y^{\varepsilon ,h} \right) \le T \right) =0 \end{aligned}$$
(A.8)

A direct computation shows that there exist a constant \(c>0\) depending on \(x,\sigma ,\alpha (\cdot )\) such that:

$$\begin{aligned} \inf _{y \ge \frac{1}{2} \varepsilon x^{1-\gamma }} \Bigl \{ \tilde{b}^{\varepsilon } (y)-\beta (1-\gamma ) y \Bigr \} \ge - c \varepsilon . \end{aligned}$$
(A.9)

Define \((Z_{t})_{t \in [0,T]}\) by

$$\begin{aligned} Z_{t} = \varepsilon x^{1-\gamma } + \left( -c\varepsilon +\sigma (1-\gamma ) k \right) t + \beta (1-\gamma ) \int _{0}^{t} Z _{s}\textit{ds} + \varepsilon \sigma (1-\gamma ) B_{t} \end{aligned}$$
(A.10)

Using (A.9), it follows from the comparison principle for SDEs that

$$\begin{aligned} \quad Y_{t} \ge Z_{t} \quad \forall \, t \le T^{\varepsilon }(Y), \quad a.s. \end{aligned}$$
(A.11)

We claim that

$$\begin{aligned} W \left( T^{\varepsilon } \left( Z\right) \le T \right) \rightarrow 0 \end{aligned}$$
(A.12)

holds true. Since \(W \left( T^{\varepsilon } \left( Y \right) \le T \right) \le W \left( T^{\varepsilon } \left( Z\right) \le T \right) \) by (A.11), then (A.8) holds. We prove (A.12) later on. Now, it follows from the definition of \(S_0(h)_t\) and an application of Gronwall’s Lemma that

$$\begin{aligned} |Y_{t} -\mathcal {S}_{0}(h)_{t}| \le \varepsilon \biggl ( c+ \sigma (1-\gamma ) \sup _{t \in [0,T]}| B_{t}| \biggr ) e^{|\beta |(1-\gamma )T} =:\Theta _T \qquad \forall t\le T^{\varepsilon } \left( Y \right) , \end{aligned}$$

therefore, for any \(R >0\) and \(\varepsilon \) small enough

$$\begin{aligned} W \biggl ( \sup _{t \in [0,T^{\varepsilon }] } |Y_{t} -\mathcal {S}_{0}(h)_{t}| \le R \biggr )&\ge W \biggl (\biggl \{ \sup _{t \in [0,T^{\varepsilon }(Y)] }|Y_{t} -\mathcal {S}_{0} (h)_{t}| \le \Theta _T \biggr \} \cap \left\{ \Theta ^{\varepsilon }_T \le R \right\} \biggr ) \\&\ge W \left( \left\{ T^{\varepsilon } (Y) \ge T \right\} \cap \left\{ \Theta ^{\varepsilon }_T \le R \right\} \right) . \end{aligned}$$

Since both the events in the right hand side of the last inequality have probability converging to 1, (A.6) follows, and Lemma 3.10 is proved under condition (A.7).

Step 2. We assume that (A.7) holds only on the time interval \([0,\rho ]\), that is \(\dot{h}_{t} \ge k\) for every \(t \le \rho \), for some \(k, \rho >0\). Repeating the argument of Step 1 with \(T= \rho \), we have

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0} W\biggl (\sup _{t \in [0,\rho ]} | Y_{t}-S_{0}(h)_{t} | \le R^{'} \biggr ) =1, \quad \forall R^{'}>0 \end{aligned}$$
(A.13)

We apply estimate (A.13) together with a localization argument. Define a time-shift operator \(\tau _{\rho } \omega \), for every \(\omega \in \Omega \), by \( (\tau _{ \rho } \omega )_{t} = \omega _{\rho +t}\) for all \(t \in [0, T-\rho ]\). For any fixed \(y>0\), denote \(X^{y,\rho }\) the strong solution of the SDE:

$$\begin{aligned} X^{y,\rho }_{t} = y^{\frac{1}{(1-\gamma )}} + \int _{0}^{t} b^{\varepsilon }(X^{y,\rho }_{s}) + \sigma |X^{y,\rho }_{s}|^{\gamma } \dot{h}_{\rho +s} \textit{ds} + \varepsilon \sigma \int _{0}^{t} |X^{y,\rho }_{s}|^{\gamma } \textit{dB}_{s} \end{aligned}$$

and set

$$\begin{aligned} Y^{y,\rho }:= (X^{y,\rho })^{1-\gamma }. \end{aligned}$$

Note that \(Y^{y,\rho }\) is well defined since \(X^{y,\rho } \ge 0\) for all \(t \in [0,T]\), W-almost surely. If \(h=0\) the non negativity of the trajectories of \(X^{y,\rho }\) follows from an application Proposition 3.1 in [9] and extends to \(h \in H \) by an application of the Girsanov theorem. By definition of Y and \(Y^{y,\rho }\), the Markov property yields

$$\begin{aligned} \mathbb {E} ( f(\tau _{\rho }Y) | \mathcal {F}_{\rho }) = \mathbb {E} ( f(Y^{Y_{\rho },\rho }) ) \end{aligned}$$

By the continuity of the map \((h,y) \mapsto \mathcal {S}_{y}(h)\) we can choose \(R'>0\) such that

$$\begin{aligned} \sup _{y \in B(S_{0}(h)_{\rho },R')} \sup _{t \in [0,T-\rho ]} | \mathcal {S}_{y}(\tau _{\rho } h)_{t} - \mathcal {S}_{\mathcal {S}_{0}(h)_{\rho }}(\tau _{\rho }h)_{t} | \le \frac{R}{2} \end{aligned}$$
(A.14)

Therefore, using (A.14) the following inclusion of events holds (assume w.lo.g \(R' \le \frac{R}{2}\)):

$$\begin{aligned} \biggl \{ \sup _{t \in [0,T]} |Y_{t} - \mathcal {S}_{0} (h)_{t} | \le R \biggr \}&\supseteq \biggl \{ \sup _{[0,\rho ]}|Y_{t} - \mathcal {S}_{0}(h)_{t} | \le R' \biggr \}\\&\cap \biggl \{ \sup _{t\in [0,T-\rho ]} |\tau _{\rho }(Y)_{t} - \mathcal {S}_{Y_{\rho }}(\tau _{\rho }h)_{t} | \le \frac{R}{2} \biggr \} \end{aligned}$$

Applying the Markov property

$$\begin{aligned} W \biggl ( \sup _{t \in [0,T]}&|Y_{t} - \mathcal {S}_{0}(h)_{t} | \le R \biggr )\nonumber \\&\ge \mathbb {E} \biggl ( \mathbb {1}_{ \{ \sup _{t \in [0,\rho ]}|Y_{t} - \mathcal {S}_{0}(h)_{t} | \le R' \}} W \biggl ( \sup _{t \in [0,T-\rho ]} |Y^{ Y_{\rho }, \rho }_{t} - S_{Y_{\rho }}(\tau _{\rho }h)_{t} | \le \frac{R}{2} \biggr ) \biggr ) \nonumber \\&\ge W \biggl ( \sup _{t \in [0,\rho ]}|Y_{t} - \mathcal {S}_{0}(h)_{t} | \le R' \biggr )\\&\quad \times \inf _{y \in B(\mathcal {S}_{0} (h)_{\rho } ,R') } W \biggl ( \sup _{t \in [0,T-\rho ]} | Y^{y,\rho }_{t} - \mathcal {S}_{y}(\tau _{ \rho }h)_{t}| \le \frac{R}{2} \biggr ) \end{aligned}$$
(A.15)

We want to show that

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0 }\inf _{ y \in B \left( S_{0} \left( h \right) _{\rho },R' \right) } W \left( \sup _{t \in [0,T-\rho ]} | Y^{y,\rho }_{t} - \mathcal {S}_{y} (\tau _{ \rho } h)_{t} | \le \frac{R}{2} \right) = 1 \end{aligned}$$
(A.16)

It follows from the hypothesis \(\mathcal {S}_{0}(h)_{t} >0 \ \forall t>0\) and the continuity of the map \((y,h) \mapsto \mathcal {S}_{y}(h)\) that, if \(R',R\) are small enough

$$\begin{aligned} y^{*}:=\inf _{y \in B\left( \mathcal {S}_{0} \left( h \right) _{\rho },R' \right) } \inf _{t \in [0,T-\rho ]} \mathcal {S}_{y} ( \tau _{\rho } h )_{t} -\frac{R}{2} >0. \end{aligned}$$
(A.17)

Define \(U^{y,\rho }\) as the unique strong solution of the SDE:

$$\begin{aligned} U^{y,\rho }_{t} = y + \int _{0}^{t} \bigl ( \tilde{b}^{\varepsilon }_{u} (U^{y,\rho }_{s}) + \sigma (1-\gamma ) \dot{h}_{s+\rho } \bigr )\textit{ds} + \varepsilon \sigma (1-\gamma ) B_{t}, \end{aligned}$$

where

$$\begin{aligned} \tilde{b}^{\varepsilon }_{u}(y) = {\left\{ \begin{array}{ll} \tilde{b}^{\varepsilon }(y) &{} \text {if}\, y \ge y^* \\ \beta (1-\gamma )y + (1-\gamma )\varepsilon ^{\frac{1}{1-\gamma }}\alpha (\varepsilon ^{-\frac{1}{(1-\gamma )}}(y^*)^{\frac{1}{(1-\gamma )}}) \frac{1}{(y^*)^{\frac{\gamma }{1-\gamma }}} - \frac{\sigma ^{2} \gamma (1-\gamma )}{2}\varepsilon ^{2} \frac{1}{y^{*}} &{} \text {if}\, y < y^{*}. \end{array}\right. } \end{aligned}$$

Then one has

$$\begin{aligned} W \biggl ( \sup _{t \in [0,T - \rho ]} | Y^{y,\rho }_{t} - S_{y}(\tau _{ \rho }h)_{t} | \le \frac{R}{2} \biggr ) = W \biggl ( \sup _{t \in [0, T-\rho ]} | U^{y,\rho }_{t} - S_{y}(\tau _{\rho }h )_{t}| \le \frac{R}{2} \biggr ). \end{aligned}$$
(A.18)

Now observing that \(\tilde{b}^{u}_{\varepsilon } \) is globally Lipschitz continuous \(\forall \varepsilon >0\) and \(C^{\varepsilon } := \sup _{y \in \mathbb {R}}|\tilde{b}^{\varepsilon }_{u} (y) -\beta (1-\gamma )y| \rightarrow 0\), an application of Gronwall’s lemma gives

$$\begin{aligned} \mathbb {E} \biggl ( \sup _{t \in [0,T-\rho ]} \left| U^{y,\rho }_{t} - \mathcal {S}_{y}(\tau _{\rho } h)_{t} \right| \biggr ) \le (C^{\varepsilon }T + 2\varepsilon \sigma (1-\gamma )\sqrt{T} )\exp ( |\beta (1-\gamma )| T). \end{aligned}$$
(A.19)

By letting \(\varepsilon \rightarrow 0\) and applying the Markov inequality, observing that the right hand side of (A.19) does not depend on y, we have proven (A.16). By letting \(\varepsilon \rightarrow 0\) in (A.15) and applying (A.13) and (A.16), the proof of Lemma 3.10 is complete.\(\blacksquare \)

Proof of (A.12). Observe that \(\tilde{Z}:=\frac{1}{\varepsilon }Z\) is an Ornstein-Uhlenbeck process,

$$\begin{aligned} \tilde{Z}_t = x^{1-\gamma } + \mu _{\varepsilon } t + \beta (1-\gamma )\int _{0}^{t}\tilde{Z}_{s}\textit{ds} + \sigma (1-\gamma ) B_t \end{aligned}$$
(A.20)

where \(\mu _{\varepsilon }:= \frac{1}{\varepsilon }(-c \varepsilon \,+\,\sigma (1\,-\,\gamma )k) = -c\,+\,\frac{\sigma (1-\gamma )k}{\varepsilon }\). It is immediate by the definition of \(\tilde{Z}\) that \(W \left( T^{\varepsilon }(Z) \le T \right) = W \left( \inf _{t \in [0,T]} \tilde{Z} \le \frac{x^{1-\gamma }}{2} \right) \). The explicit representation of \(\tilde{Z}\) reads

$$\begin{aligned} \tilde{Z}_t := x^{1-\gamma }e^{\beta (1-\gamma ) t} + f_{\varepsilon }(t) + \sigma (1-\gamma )\exp (\beta (1-\gamma ) t) \int _{0}^{t} \exp (- \beta (1-\gamma ) s) \textit{dB}_s \end{aligned}$$
(A.21)

with \(f_{\varepsilon }(t) = -\frac{\mu _{\varepsilon } (1-\exp (\beta (1-\gamma ) t) )}{\beta (1-\gamma )}\). Consider a deterministic time \(\tau _{\varepsilon }\) with \(\tau _{\varepsilon } \rightarrow 0\) as \(\varepsilon \rightarrow 0\), to be chosen precisely later on. Noting that \(f_{\varepsilon }\) is a decreasing function, for \(\tau _{\varepsilon } \le t \le T\) one has

$$\begin{aligned} \tilde{Z}_t \ge f_{\varepsilon }(\tau _{\varepsilon }) - \sigma (1-\gamma ) \Bigl | \int _{0}^{t} \exp (-\beta (1-\gamma ) s ) \textit{dB}_s \Bigr |; \end{aligned}$$
(A.22)

hence, using Markov’s inequality and Doob’s inequality

Now, the choice \(\tau _{\varepsilon } = \sqrt{\varepsilon }\) gives \(f_{\varepsilon }(\tau _{\varepsilon }) \sim \mu _{\varepsilon } \tau _{\varepsilon } \rightarrow \infty \) as \(\varepsilon \rightarrow 0\), so that \(\left( f_{\varepsilon }(\tau _{\varepsilon }) - x^{1-\gamma }/2 \right) ^{-1} \rightarrow 0\). On the other hand, \(\inf _{t \in [0,\tau _{\varepsilon }]} \tilde{Z}_t \rightarrow x^{1-\gamma }\) a.s. as \(\varepsilon \rightarrow 0\), hence \(W \Bigl ( \inf _{t \in [0,\tau _{\varepsilon }]} \tilde{Z}_t \le x/2 \Bigr ) \rightarrow 0\) as \(\varepsilon \rightarrow 0\), and the claim is proven.

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this paper

Cite this paper

Conforti, G., De Marco, S., Deuschel, JD. (2015). On Small-Noise Equations with Degenerate Limiting System Arising from Volatility Models. In: Friz, P., Gatheral, J., Gulisashvili, A., Jacquier, A., Teichmann, J. (eds) Large Deviations and Asymptotic Methods in Finance. Springer Proceedings in Mathematics & Statistics, vol 110. Springer, Cham. https://doi.org/10.1007/978-3-319-11605-1_17

Download citation

Publish with us

Policies and ethics