This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Large deviations and stochastic volatility with jumps:
asymptotic implied volatility for affine models

Antoine Jacquier Institut für Mathematik - Technische Universität Berlin, Germany jacquier@math.tu-berlin.de Martin Keller-Ressel Institut für Mathematik - Technische Universität Berlin, Germany mkeller@math.tu-berlin.de  and  Aleksandar Mijatović Department of Statistics, University of Warwick, UK a.mijatovic@warwick.ac.uk
Abstract.

Let σt(x)\sigma_{t}(x) denote the implied volatility at maturity tt for a strike K=S0extK=S_{0}\mathrm{e}^{xt}, where xx\in\mathbb{R} and S0S_{0} is the current value of the underlying. We show that σt(x)\sigma_{t}(x) has a uniform (in xx) limit as maturity tt tends to infinity, given by the formula σ(x)=2(h(x)1/2+(h(x)x)1/2)\sigma_{\infty}(x)=\sqrt{2}\left(h^{*}(x)^{1/2}+\left(h^{*}(x)-x\right)^{1/2}\right), for xx in some compact neighbourhood of zero in the class of affine stochastic volatility models. The function hh^{*} is the convex dual of the limiting cumulant generating function hh of the scaled log-spot process. We express hh in terms of the functional characteristics of the underlying model. The proof of the limiting formula rests on the large deviation behaviour of the scaled log-spot process as time tends to infinity. We apply our results to obtain the limiting smile for several classes of stochastic volatility models with jumps used in applications (e.g. Heston with state-independent jumps, Bates with state-dependent jumps and Barndorff-Nielsen-Shephard model).

Key words and phrases:
Large deviation principle; Stochastic volatility with jumps; Affine processes; Implied volatility in the large maturity limit
2000 Mathematics Subject Classification:
60G44, 60F10, 91G20
AJ would like to thank MATHEON for financial support.

1. Introduction

Let the process S=eXS=\mathrm{e}^{X} model a risky security under an equivalent martingale measure and let σt(x)\sigma_{t}(x) denote the implied volatility at maturity tt for a strike K=S0extK=S_{0}\mathrm{e}^{xt} (see (53) for the precise definition of σt(x)\sigma_{t}(x)). The main result of the present paper (Theorem 14) states that, if the log-spot XX follows an affine stochastic volatility process with jumps, then σt(x)\sigma_{t}(x) converges to σ(x)\sigma_{\infty}(x) as the maturity tt tends to infinity, where σ\sigma_{\infty} is given by the formula

(1) σ(x)\displaystyle\sigma_{\infty}(x) =\displaystyle= 2(h(x)1/2+(h(x)x)1/2).\displaystyle\sqrt{2}\left(h^{*}(x)^{1/2}+\left(h^{*}(x)-x\right)^{1/2}\right).

The function hh is the limiting cumulant generating function of the scaled log-spot (Xt/t)t1(X_{t}/t)_{t\geq 1} and hh^{*} is its convex dual (i.e. the Fenchel-Legendre transform of hh). Locally uniform convergence of the implied volatility to σ\sigma_{\infty} is also established.

In [FJ11, FJM11] the limiting behaviour of the smile at large maturities in the Heston model is investigated. Theorem 14 can be viewed as a generalisation of the main result in [FJ11, FJM11]. Not only does it cover a large class of stochastic volatility models with jumps rather than a single affine model with continuous trajectories, but furthermore provides a better understanding of the limit: Theorem 14 states that the limit holds also at the critical points xx^{*} and x~\widetilde{x}^{*}, which are excluded from the analysis in [FJ11, FJM11], and the convergence on the set {x,x~}\mathbb{R}\setminus\{x^{*},\widetilde{x}^{*}\} is shown to be uniform on compact subsets.

In the class of affine stochastic volatility models, the formula for the limiting implied volatility for a fixed strike proved in Tehranchi [Teh09] (see also [Lew00] in the case of the Heston model) also follows from (1), since in Theorem 14 the convergence is uniform on a compact neighbourhood of the origin. In [GL11], the authors give various representations for the implied volatility, including in the large-maturity regime, based on an assumed asymptotic behaviour of certain European derivatives in the underlying model, which is not specified. This representation is not fully explicit in terms of the model parameters and it is therefore unclear how to apply it directly to the class of affine stochastic volatility models.

Contribution of the paper is twofold. First we study the properties of the limiting cumulant generating function hh of the affine stochastic volatility models. Results in Lemma 9, Theorem 10 and Corollary 11 give new properties of the function hh, which are crucial for the understanding of the large deviation behaviour of the model. Second, the problem of understanding the limiting behaviour of option prices and the corresponding implied volatilities using the large deviation principle is tackled. The uniform limit in xx (on all compact subsets of \mathbb{R}) of vanilla option prices is given in Theorem 13 for non-degenerate affine stochastic volatility models and exponential Lévy models (i.e. degenerate affine stochastic volatility models). As mentioned above Theorem 14 deals with the limiting implied volatility smiles in these classes of models.

Besides giving a formula, which relates model parameters and the limiting implied volatility smile, these theoretical results have the following practical consequences:
(1) in the large-maturity regime studied in this paper, the jumps in the model influence the limiting implied volatility smile as maturity tends to infinity (see examples in 6.1);
(2) for every affine stochastic volatility model there exists an exponential Lévy model such that the smiles of the two models in the limit coincide. In other words the stochasticity of volatility does not (in the affine class) enlarge the family of possible limiting implied volatility smiles (see Section 6 for details).

The starting point of the analysis of the large deviation behaviour of an affine stochastic volatility process (X,V)(X,V) in the present paper is Theorem 8, taken from [KR11, Theorem 3.4]. This result describes certain properties of the limiting cumulant generating function hh, which are however insufficient to understand the essential smoothness of hh required in establishing the large deviation principle of (Xt/t)t1(X_{t}/t)_{t\geq 1}. The main contribution of this paper in the area of affine processes is Theorem 10, which identifies sufficient conditions for the process (X,V)(X,V) that imply essential smoothness of the function hh. The conditions in Theorem 10 are easy to apply to the models of interest (see e.g. Section 2.2). Its proof goes beyond the analysis in [KR11] as one is forced to study the special Lévy-Khintchine form of the characteristics of the process, since their general convexity properties no longer suffice to establish the required behaviour of the limit.

The rest of the paper is organized as follows. In Section 2 we define the class of affine stochastic volatility processes and recall some of their properties. In Section 3 we review briefly basic concepts in the theory of large deviations and state the Gärtner-Ellis theorem. Section 4 establishes the large deviations principle for the scaled log-stock of an affine stochastic volatility model as maturity tends to infinity. Sections 5 and 6 respectively translate this result into option price and implied volatility asymptotics. Numerical examples are given at the end of Section 6.

2. Affine stochastic volatility models with jumps

Consider a stochastic model for a risky security S=(St)t0S=(S_{t})_{t\geq 0} given by

(2) St\displaystyle S_{t} =\displaystyle= exp((rd)t+Xt),t0,\displaystyle\exp((r-d)t+X_{t}),\quad t\geq 0\;,

where the interest rate rr and the dividend yield dd are non-negative and constant and the log-price process X=(Xt)t0X=(X_{t})_{t\geq 0} starts at X0X_{0}\in\mathbb{R}. Since the dynamics of SS is given under a risk-neutral measure, the forward price process is (exp(Xt))t0(\exp(X_{t}))_{t\geq 0}. We assume throughout the paper without loss of generality that SS is a forward price process (i.e. r=dr=d). Denote by V=(Vt)t0V=(V_{t})_{t\geq 0} a process, starting at a constant level V0>0V_{0}>0. The process VV can be interpreted as the instantaneous variance process of XX but may also control the arrival rate of jumps of XX. We make the following assumptions on the process (X,V)(X,V) throughout the paper.

A1:

(X,V)(X,V) is a stochastically continuous, time-homogeneous Markov process with state-space D=×0D=\mathbb{R}\times\mathbb{R}_{\geqslant 0}, where 0:=[0,)\mathbb{R}_{\geqslant 0}:=[0,\infty).

A2:

The cumulant generating function Φt(u,w)\Phi_{t}(u,w) of (Xt,Vt)(X_{t},V_{t}) is of a particular affine form: there exist functions ϕ(t,u,w)\phi(t,u,w) and ψ(t,u,w)\psi(t,u,w) such that

Φt(u,w)\displaystyle\Phi_{t}(u,w) :=\displaystyle:= log𝖤[exp(uXt+wVt)|X0,V0]\displaystyle\log\mathsf{E}\left[\left.\exp(uX_{t}+wV_{t})\right|X_{0},V_{0}\right]
=\displaystyle= ϕ(t,u,w)+V0ψ(t,u,w)+X0u\displaystyle\phi(t,u,w)+V_{0}\psi(t,u,w)+X_{0}u

for all (t,u,w)0×2(t,u,w)\in\mathbb{R}_{\geqslant 0}\times\mathbb{C}^{2}, where the expectation exists.

Remarks.

(i) A1 and A2 make (X,V)(X,V) into an affine process in the sense of [DFS03].

(ii) A1 and A2 imply a homogeneity property of (X,V)(X,V): if the starting value (X0,V0)(X_{0},V_{0}) is shifted by (x,0)D(x,0)\in D, the law of the random variable (Xt,Vt)(X_{t},V_{t}) is shifted by the vector (x,0)(x,0) for any t0t\geq 0.

(iii) Assumptions A1 and A2 imply that the variance process VV is a one-dimensional strong Markov process in its own right.

(iv) The law of iterated expectations applied to Φt(u,w)\Phi_{t}(u,w) yields the flow-equations for ϕ\phi and ψ\psi (see [DFS03, Eq. (3.8)–(3.9)]):

(3) ϕ(t+s,u,w)=ϕ(t,u,w)+ϕ(s,u,ψ(t,u,w)),ψ(t+s,u,w)=ψ(s,u,ψ(t,u,w)),\begin{split}\phi(t+s,u,w)&=\phi(t,u,w)+\phi(s,u,\psi(t,u,w)),\\ \psi(t+s,u,w)&=\psi(s,u,\psi(t,u,w)),\end{split}

for all t,s0t,s\geq 0.

(v) It is shown in [KR11, Thm. 2.1] (see also [DFS03]) that if |ϕ(τ,u,η)|,|ψ(τ,u,η)|<|\phi(\tau,u,\eta)|,|\psi(\tau,u,\eta)|<\infty for (τ,u,η)(0,)×2(\tau,u,\eta)\in(0,\infty)\times\mathbb{C}^{2}, then for all t[0,τ)t\in[0,\tau) and ww\in\mathbb{C} such that wη\Re\,w\leq\Re\,\eta, the functions ϕ\phi and ψ\psi satisfy the generalized Riccati equations

(4a) tϕ(t,u,w)\displaystyle\partial_{t}\phi(t,u,w) =F(u,ψ(t,u,w)),ϕ(0,u,w)=0,\displaystyle=F(u,\psi(t,u,w)),\quad\phi(0,u,w)=0,
(4b) tψ(t,u,w)\displaystyle\partial_{t}\psi(t,u,w) =R(u,ψ(t,u,w)),ψ(0,u,w)=w,\displaystyle=R(u,\psi(t,u,w)),\quad\psi(0,u,w)=w,

where

(5) F(u,w):=tϕ(t,u,w)|t=0+,R(u,w):=tψ(t,u,w)|t=0+.F(u,w):=\left.\frac{\partial}{\partial t}\phi(t,u,w)\right|_{t=0+},\qquad R(u,w):=\left.\frac{\partial}{\partial t}\psi(t,u,w)\right|_{t=0+}.

Furthermore for all t[0,τ]t\in[0,\tau] we have |ϕ(t,u,w)|,|ψ(t,u,w)|<|\phi(t,u,w)|,|\psi(t,u,w)|<\infty.

(vi) If (X,V)(X,V) is a diffusion process, then ODEs (4) become classical Riccati. Note also that (4) follows from the flow equations (3) by differentiation with respect to ss.

(vii) ϕ\phi and ψ\psi can for small tt be expressed implicitly in terms of FF and RR as

ϕ(t,u,w)=0tF(u,ψ(s,u,w))dsandwψ(t,u,w)dηR(u,η)=t.\phi(t,u,w)=\int_{0}^{t}{F(u,\psi(s,u,w))\;\mathrm{d}s}\qquad\text{and}\qquad\int_{w}^{\psi(t,u,w)}{\frac{\mathrm{d}\eta}{R(u,\eta)}}=t\;.

The functions FF and RR, defined in (5), must be of Lévy-Khintchine form (see [DFS03]). In other words

(6a) F(u,w)\displaystyle F(u,w) =a2(u,w),(u,w)+b,(u,w)c\displaystyle=\left\langle\frac{a}{2}(u,w)^{\prime},(u,w)^{\prime}\right\rangle+\left\langle b,(u,w)^{\prime}\right\rangle-c
+D{0}(eξ,(u,w)1ωF(ξ),(u,w))m(dξ),\displaystyle+\int_{D\setminus\{0\}}{\left(\mathrm{e}^{\langle\xi,(u,w)^{\prime}\rangle}-1-\langle\omega_{F}(\xi),(u,w)^{\prime}\rangle\right)\,m(\mathrm{d}\xi)},
(6b) R(u,w)\displaystyle R(u,w) =α2(u,w),(u,w)+β,(u,w)γ\displaystyle=\left\langle\frac{\alpha}{2}(u,w)^{\prime},(u,w)^{\prime}\right\rangle+\left\langle\beta,(u,w)^{\prime}\right\rangle-\gamma
+D{0}(eξ,(u,w)1ωR(ξ),(u,w))μ(dξ),\displaystyle+\int_{D\setminus\{0\}}{\left(\mathrm{e}^{\langle\xi,(u,w)^{\prime}\rangle}-1-\langle\omega_{R}(\xi),(u,w)^{\prime}\rangle\right)\,\mu(\mathrm{d}\xi)},

where D=×0D=\mathbb{R}\times\mathbb{R}_{\geqslant 0}, ,\langle\cdot,\cdot\rangle is the inner product on 2\mathbb{R}^{2}, (u,w)(u,w)^{\prime} denotes transposition, ωF\omega_{F}, ωR\omega_{R} are suitable truncation functions, which we fix by defining

ωF(ξ)=(ξ11+ξ120)andωR(ξ)=(ξ11+ξ12ξ21+ξ22),whereξ=(ξ1ξ2),\omega_{F}(\xi)=\left(\begin{array}[]{@{}c@{}}\frac{\xi_{1}}{1+\xi_{1}^{2}}\\ 0\end{array}\right)\qquad\text{and}\qquad\omega_{R}(\xi)=\left(\begin{array}[]{@{}c@{}}\frac{\xi_{1}}{1+\xi_{1}^{2}}\\ \frac{\xi_{2}}{1+\xi_{2}^{2}}\end{array}\right),\quad\text{where}\quad\xi=\left(\begin{array}[]{@{}c@{}}\xi_{1}\\ \xi_{2}\end{array}\right),

and the parameters (a,α,b,β,m,μ)(a,\alpha,b,\beta,m,\mu) satisfy the following admissibility conditions:

  • a,αa,\alpha are positive semi-definite 2×22\times 2-matrices with a12=a21=a22=0a_{12}=a_{21}=a_{22}=0;

  • bDb\in D, β2\beta\in\mathbb{R}^{2} and c,γ0c,\gamma\in\mathbb{R}_{\geqslant 0};

  • mm and μ\mu are Lévy measures on DD and D{0}((ξ12+ξ2)1)m(dξ)<\int_{D\setminus\{0\}}{\left((\xi_{1}^{2}+\xi_{2})\wedge 1\right)\,m(\mathrm{d}\xi)}<\infty.

Assumptions A1 and A2, the generalized Riccati equations and the Lévy-Khintchine decomposition (6) lead to the following interpretation of FF and RR: FF characterizes the state-independent dynamics of the process (X,V)(X,V) while RR characterizes its state-dependent dynamics. The instantaneous characteristics of the Markov process (X,V)(X,V) are given as follows: a+Vαa+V\alpha the instantaneous covariance matrix, b+Vβb+V\beta the instantaneous drift, m(dξ)+Vμ(dξ)m(\mathrm{d}\xi)+V\mu(\mathrm{d}\xi) the instantaneous arrival rate of jumps with jump heights in dξ\mathrm{d}\xi and c+γVc+\gamma V the instantaneous killing rate.

The function χ\chi defined below plays a key role in the characterisation of the martingale property of the process S=exp(X)S=\exp(X).

Definition 1.

For each uu\in\mathbb{R} such that R(u,0)<R(u,0)<\infty, define χ(u)\chi(u) as

χ(u):=2R(u,w)|w=0:=Rw(u,w)|w=0.\chi(u):=\left.\partial_{2}R(u,w)\right|_{w=0}:=\left.\frac{\partial R}{\partial w}(u,w)\right|_{w=0}\;.
Remarks.

(i) The condition R(u,0)<R(u,0)<\infty implies that, for some δ>0\delta>0 the function wR(u,w)w\mapsto R(u,w) is convex on (δ,0](-\delta,0] and differentiable on (δ,0)(-\delta,0), since the process VV does not have negative jumps. Therefore χ(u)\chi(u) is a well-defined, possibly equal to ++\infty, convex function given by the limit of 2R(u,w)\partial_{2}R(u,w) as w0w\uparrow 0. It can be expressed explicitly as

χ(u)=α12u+β1+D{0}ξ2(euξ111+ξ22)μ(dξ),whereξ=(ξ1,ξ2).\chi(u)=\alpha_{12}u+\beta_{1}+\int_{D\setminus\{0\}}{\xi_{2}\left(\mathrm{e}^{u\xi_{1}}-\frac{1}{1+\xi_{2}^{2}}\right)\,\mu(\mathrm{d}\xi)},\qquad\text{where}\quad\xi^{\prime}=(\xi_{1},\xi_{2})\;.

(ii) The sufficient and necessary condition for SS to be conservative and a martingale, in terms of R,FR,F and χ\chi, is given in [KR11, Thm. 2.5]. A simple sufficient condition for these properties reads (see [KR11, Cor. 2.7]):

  1. \bullet

    if F(0,0)=R(0,0)=0F(0,0)=R(0,0)=0 and χ(0)<\chi(0)<\infty then S=exp(X)S=\exp(X) is conservative.

  2. \bullet

    if SS is conservative, F(1,0)=R(1,0)=0F(1,0)=R(1,0)=0 and χ(1)<\chi(1)<\infty, then S=exp(X)S=\exp(X) is a martingale.

Since SS serves as a forward price process under a risk-neutral measure 𝖯\mathsf{P} in an arbitrage-free asset pricing model, it has to be conservative and a martingale and hence we assume:

A3:

F(0,0)=R(0,0)=F(1,0)=R(1,0)=0F(0,0)=R(0,0)=F(1,0)=R(1,0)=0 and χ(0)+χ(1)<\chi(0)+\chi(1)<\infty.

In particular F(0,0)=R(0,0)=0F(0,0)=R(0,0)=0 in assumption A3 means that the instantaneous killing rates cc and γ\gamma in (6) are zero and the condition F(1,0)=R(1,0)=0F(1,0)=R(1,0)=0 is closely related to the functions ψ(,1,0)\psi(\cdot,1,0) and ϕ(,1,0)\phi(\cdot,1,0) being identically equal to zero (see the generalized Riccati equations in (4)), which implies the martingale property of S=exp(X)S=\exp(X). The following non-degeneracy assumption will guarantee the stochasticity of volatility of the process XX.

A4:

There exists some uu\in\mathbb{R}, such that R(u,0)0R(u,0)\neq 0.

Definition 2.

The process (X,V)(X,V) is a non-degenerate (resp. degenerate) affine stochastic volatility process if it satisfies assumptions A1 – A4 (resp. A1 – A3 and does not satisfy A4) and S=eXS=\mathrm{e}^{X} is the corresponding affine stochastic volatility model.

Remark.

Assumption A4 excludes the degenerate case where the distribution of XX does not depend on the volatility state V0V_{0}. Indeed, if A4 is not satisfied, i.e. R(,0)0R(\cdot,0)\equiv 0, then (4) implies that ψ(t,u,0)=0\psi(t,u,0)=0 and ϕ(t,u,0)=tF(u,0)\phi(t,u,0)=tF(u,0) for all (t,u)0×(t,u)\in\mathbb{R}_{\geqslant 0}\times\mathbb{C} where the expectation in assumption A2 exists. Hence if A4 does not hold, then A2,  (6a) and the characterisation theorem for regular affine processes [DFS03, Theorem 2.7] imply that S=eXS=\mathrm{e}^{X} is an exponential Lévy model. In particular the class of affine stochastic volatility models includes the Black-Scholes model as a degenerate case.

The following proposition describes certain properties of FF and RR that will play a crucial role in Section 4.1.

Proposition 3.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility model and let the sets 𝒟F={(u,w)2:F(u,w)<}\mathcal{D}_{F}=\left\{(u,w)\in\mathbb{R}^{2}:F(u,w)<\infty\right\} and 𝒟R={(u,w)2:R(u,w)<}\mathcal{D}_{R}=\left\{(u,w)\in\mathbb{R}^{2}:R(u,w)<\infty\right\} be the effective domains of the functions FF and RR respectively. Then the following holds:

  1. (A)

    FF and RR are lower semicontinuous convex functions, which are continuously differentiable in the interiors 𝒟F\mathcal{D}_{F}^{\circ} and 𝒟R\mathcal{D}_{R}^{\circ} (in 2\mathbb{R}^{2}), and their effective domains 𝒟F\mathcal{D}_{F} and 𝒟R\mathcal{D}_{R} are also convex;

  2. (B)

    FF and RR are either affine or strictly convex functions when restricted to one-dimensional affine subspaces of 2\mathbb{R}^{2}.

Proof.

The Lévy-Khintchine representation for FF and RR in (6) implies that they are cumulant generating functions of some (infinitely divisible) random vectors taking values in 2\mathbb{R}^{2}. Hölder’s inequality yields that FF and RR are convex. The dominated convergence theorem and the representation in (6) implies that FF and RR are analytic in 𝒟F\mathcal{D}_{F}^{\circ} and 𝒟R\mathcal{D}_{R}^{\circ} respectively. Fatou’s lemma implies that the functions FF and RR are lower semicontinuous. Since FF and RR are cumulant generating functions, the second derivative of their restriction to an affine subspace is either identically zero or strictly positive everywhere (each affine subspace in 2\mathbb{R}^{2} corresponds to a random variable which takes values in \mathbb{R} and may or may not be constant almost surely). This concludes the proof. ∎

2.1. SDE representation of affine stochastic volatility processes

In order to define an affine stochastic volatility model one needs to choose admissible parameters (a,α,b,β,m,μ)(a,\alpha,b,\beta,m,\mu) such that the corresponding process (X,V)(X,V), which exists by [DFS03, Thm. 2.7], satisfies assumptions A1 – A3 (note that c=γ=0c=\gamma=0 by A3 and will henceforth be ignored). This procedure yields a semigroup, and hence the law, of the Markov process (X,V)(X,V) which is in principle sufficient for option pricing. However path-wise descriptions of the pricing models in financial markets are widely used as they add to the intuitive understanding of the properties of the model. In the rest of this section we briefly describe a path-wise construction of the process (X,V)(X,V), given in [DL06], and relate it to the most popular affine stochastic volatility models used in derivatives pricing.

Assume that the parameters (a,α,b,β,m,μ)(a,\alpha,b,\beta,m,\mu) are admissible and suppose in addition that the tails (i.e. the large jumps) of mm and μ\mu satisfy:

(7) ξ>1ξm(dξ)<andξ>1ξμ(dξ)<\displaystyle\int_{\|\xi\|>1}{\|\xi\|\,m(\mathrm{d}\xi)}<\infty\qquad\text{and}\qquad\int_{\|\xi\|>1}{\|\xi\|\,\mu(\mathrm{d}\xi)}<\infty

where ξ=ξ,ξ.\|\xi\|=\langle\xi,\xi\rangle. Let

(8) b~1=b1+D{0}ξ131+ξ12m(dξ),b~2=b2,β~i=βi+D{0}ξi31+ξi2μ(dξ)fori=1,2.\begin{split}\widetilde{b}_{1}&=b_{1}+\int_{D\setminus\{0\}}\frac{\xi_{1}^{3}}{1+\xi_{1}^{2}}\,m(\mathrm{d}\xi),\quad\widetilde{b}_{2}=b_{2},\\ \widetilde{\beta}_{i}&=\beta_{i}+\int_{D\setminus\{0\}}\frac{\xi_{i}^{3}}{1+\xi_{i}^{2}}\,\mu(\mathrm{d}\xi)\quad\text{for}\quad i=1,2.\end{split}

Note that the integrals in (8) are finite by (7) and the parameters (a,α,b~,β~,m,μ)(a,\alpha,\widetilde{b},\widetilde{\beta},m,\mu) are admissible with appropriate truncation functions.

Let (Ω,,(t)t0,𝖯)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},\mathsf{P}) be a filtered probability space equipped with

  1. \bullet

    a three-dimensional standard Brownian motion (B0,B1,B2)(B^{0},B^{1},B^{2}),

  2. \bullet

    a Poisson random measure N0(ds,dξ)N_{0}(\mathrm{d}s,\mathrm{d}\xi) on 0×D\mathbb{R}_{\geqslant 0}\times D with compensator dsm(dξ)\mathrm{d}s\,m(\mathrm{d}\xi),

  3. \bullet

    a Poisson random measure N1(ds,du,dξ)N_{1}(\mathrm{d}s,\mathrm{d}u,\mathrm{d}\xi) on 02×D\mathbb{R}_{\geqslant 0}^{2}\times D with compensator dsduμ(dξ)\mathrm{d}s\,\mathrm{d}u\,\mu(\mathrm{d}\xi),

where as usual D=×0D=\mathbb{R}\times\mathbb{R}_{\geqslant 0} denotes the state-space of the model. Let

N~0(ds,dξ)=N0(ds,dξ)dsm(dξ)andN~1(ds,du,dξ)=N1(ds,du,dξ)dsduμ(dξ)\widetilde{N}_{0}(\mathrm{d}s,\mathrm{d}\xi)=N_{0}(\mathrm{d}s,\mathrm{d}\xi)-\mathrm{d}s\,m(\mathrm{d}\xi)\quad\text{and}\quad\widetilde{N}_{1}(\mathrm{d}s,\mathrm{d}u,\mathrm{d}\xi)=N_{1}(\mathrm{d}s,\mathrm{d}u,\mathrm{d}\xi)-\mathrm{d}s\,\mathrm{d}u\,\mu(\mathrm{d}\xi)

be the compensated Poisson random measures and let σ\sigma be a 2×22\times 2-matrix such that σσ=α\sigma\sigma^{\top}=\alpha. Theorem 6.2 in [DL06] implies that the system of SDEs

dXt\displaystyle\mathrm{d}X_{t} =\displaystyle= (b~1+Vtβ~1)dt+a11dBt0+Vtσ11dBt1+Vtσ12dBt2+\displaystyle\left(\widetilde{b}_{1}+V_{t}\widetilde{\beta}_{1}\right)\mathrm{d}t+\sqrt{a_{11}}\mathrm{d}B_{t}^{0}+\sqrt{V_{t}}\sigma_{11}\mathrm{d}B_{t}^{1}+\sqrt{V_{t}}\sigma_{12}\mathrm{d}B_{t}^{2}+
D{0}ξ1N~0(dt,dξ)+D{0}0Vtξ1N~1(dt,du,dξ),\displaystyle\int_{D\setminus\{0\}}\xi_{1}\widetilde{N}_{0}(\mathrm{d}t,\mathrm{d}\xi)+\int_{D\setminus\{0\}}\int_{0}^{V_{t-}}\xi_{1}\widetilde{N}_{1}(\mathrm{d}t,\mathrm{d}u,\mathrm{d}\xi),
dVt\displaystyle\mathrm{d}V_{t} =\displaystyle= (b~2+Vtβ~2)dt+Vtσ21dBt1+Vtσ22dBt2+\displaystyle\left(\widetilde{b}_{2}+V_{t}\widetilde{\beta}_{2}\right)\mathrm{d}t+\sqrt{V_{t}}\sigma_{21}\mathrm{d}B_{t}^{1}+\sqrt{V_{t}}\sigma_{22}\mathrm{d}B_{t}^{2}+
D{0}ξ2N0(dt,dξ)+D{0}0Vtξ2N~1(dt,du,dξ),\displaystyle\int_{D\setminus\{0\}}\xi_{2}N_{0}(\mathrm{d}t,\mathrm{d}\xi)+\int_{D\setminus\{0\}}\int_{0}^{V_{t-}}\xi_{2}\widetilde{N}_{1}(\mathrm{d}t,\mathrm{d}u,\mathrm{d}\xi),

with initial condition (X0,V0)×(0,)(X_{0},V_{0})\in\mathbb{R}\times(0,\infty), has a unique strong solution (X,V)(X,V) that is an affine Markov process with admissible parameters (a,α,b,β,m,μ)(a,\alpha,b,\beta,m,\mu).

Remarks.

(i) The change of parameters bb and β\beta introduced in (8) is inessential. Its function is to establish the notational compatibility with [DL06].

(ii) The integrals in (2.1)–(2.1) against N~1\widetilde{N}_{1} are taken over a random set whose dsduμ(dξ)\mathrm{d}s\,\mathrm{d}u\,\mu(\mathrm{d}\xi)-volume is proportional to VtV_{t-}. This, together with the structure of the Poisson random measure N1N_{1}, reinforces the intuition that the jumps of the process (X,V)(X,V) that correspond to the integral term with respect to N~1\widetilde{N}_{1} have random intensity which is proportional to VV.

2.2. Examples of affine stochastic volatility models

We now describe some of the affine stochastic volatility models that are of interest in the financial markets and can be obtained as solutions of the special cases of SDE (2.1)–(2.1).

2.2.1. Heston model

The log-price XX and the stochastic variance process VV are given under the risk-neutral measure by the SDE

dXt\displaystyle\mathrm{d}X_{t} =Vt2dt+VtdWt1,\displaystyle=-\frac{V_{t}}{2}\,\mathrm{d}t+\sqrt{V_{t}}\,\mathrm{d}W^{1}_{t},
dVt\displaystyle\mathrm{d}V_{t} =λ(Vtθ)dt+ζVtdWt2,\displaystyle=-\lambda(V_{t}-\theta)\,\mathrm{d}t+\zeta\sqrt{V_{t}}\,\mathrm{d}W^{2}_{t},\,

where W1,W2W^{1},W^{2} are Brownian motions with correlation parameter ρ(1,1)\rho\in(-1,1), and ζ,λ,θ>0\zeta,\lambda,\theta>0 (see [Hes93]). The affine characteristics of the model are

(11a) F(u,w)\displaystyle F(u,w) =λθw,\displaystyle=\lambda\theta w,
(11b) R(u,w)\displaystyle R(u,w) =12(u2u)+ζ22w2λw+uwρζ.\displaystyle=\frac{1}{2}(u^{2}-u)+\frac{\zeta^{2}}{2}w^{2}-\lambda w+uw\rho\zeta.

It is easily seen that χ\chi is given by

(12) χ(u)=ρζuλ\chi(u)=\rho\zeta u-\lambda

and it is trivial to check that A1 – A4 are satisfied.

2.2.2. Heston model with state-independent jumps

Let JJ be a pure-jump Lévy process independent of the correlated Brownian motions W1W^{1} and W2W^{2}. The Heston-with-jumps model is defined by the SDEs

dXt\displaystyle\mathrm{d}X_{t} =(δVt2)dt+VtdWt1+dJt,\displaystyle=\left(\delta-\frac{V_{t}}{2}\right)\,\mathrm{d}t+\sqrt{V_{t}}\,\mathrm{d}W^{1}_{t}+\mathrm{d}J_{t},
dVt\displaystyle\mathrm{d}V_{t} =λ(Vtθ)dt+ζVtdWt2,\displaystyle=-\lambda(V_{t}-\theta)\,\mathrm{d}t+\zeta\sqrt{V_{t}}\,\mathrm{d}W^{2}_{t},

where ζ,λ,θ>0\zeta,\lambda,\theta>0 and δ\delta\in\mathbb{R}. Assume that JJ is a spectrally negative Lévy process with characteristic exponent 1tlog𝖤[euJt]=(,0)(euξ11uξ1/(1+ξ12))ν(dξ1)\frac{1}{t}\log\mathsf{E}[\mathrm{e}^{uJ_{t}}]=\int_{(-\infty,0)}(\mathrm{e}^{u\xi_{1}}-1-u\xi_{1}/(1+\xi_{1}^{2}))\,\nu(\mathrm{d}\xi_{1}). Since JJ only jumps down, this captures the generic situation in the modelling of equity markets. Assume further that the jumps of JJ are integrable (i.e. (,0)|ξ1|ν(dξ1)<\int_{(-\infty,0)}|\xi_{1}|\,\nu(\mathrm{d}\xi_{1})<\infty). In order to identify the coefficients in (2.1)–(2.1), we compensate JJ so that it becomes a martingale and can hence be expressed as an integral against the compensated Poisson random measure N~0\widetilde{N}_{0}. This implies

b~1=δ+(,0)ξ131+ξ12ν(dξ1).\widetilde{b}_{1}=\delta+\int_{(-\infty,0)}\frac{\xi_{1}^{3}}{1+\xi_{1}^{2}}\,\nu(\mathrm{d}\xi_{1}).

It is easily seen that a=0a=0, α11=1\alpha_{11}=1, α12=ρζ\alpha_{12}=\rho\zeta, α22=ζ2\alpha_{22}=\zeta^{2}, μ0\mu\equiv 0 and m(dξ)=(νδ0)(dξ)m(\mathrm{d}\xi)=(\nu\otimes\delta_{0})(\mathrm{d}\xi), where δ0\delta_{0} is the Dirac delta measure. Therefore b1=δb_{1}=\delta, b2=λθb_{2}=\lambda\theta, β1=1/2\beta_{1}=-1/2 and β2=λ\beta_{2}=-\lambda. The martingale condition (F(1,0)=0F(1,0)=0) implies δ=(,0)(eξ11ξ1/(1+ξ12))ν(dξ1)\delta=-\int_{(-\infty,0)}\left(\mathrm{e}^{\xi_{1}}-1-\xi_{1}/(1+\xi_{1}^{2})\right)\,\nu(\mathrm{d}\xi_{1}) and the affine form of the model is given by

(13a) F(u,w)\displaystyle F(u,w) =λθw+κ~(u),\displaystyle=\lambda\theta w+\widetilde{\kappa}(u),
(13b) R(u,w)\displaystyle R(u,w) =12(u2u)+ζ22w2λw+uwρζ,\displaystyle=\frac{1}{2}(u^{2}-u)+\frac{\zeta^{2}}{2}w^{2}-\lambda w+uw\rho\zeta,

where κ~(u)\widetilde{\kappa}(u) is the compensated cumulant generating function of the jump part, i.e.

(14) κ~(u)=(,0)(eξ1u1u(eξ11))ν(dξ1).\widetilde{\kappa}(u)=\int_{(-\infty,0)}\left(\mathrm{e}^{\xi_{1}u}-1-u\left(\mathrm{e}^{\xi_{1}}-1\right)\right)\,\nu(\mathrm{d}\xi_{1}).

2.2.3. A model of Bates with state-dependent jumps

We consider the model given by

dXt\displaystyle\mathrm{d}X_{t} =(12+δ)Vtdt+VtdWt1+{0}ξ1N~(Vt,dt,dξ1),\displaystyle=-\left(\frac{1}{2}+\delta\right)V_{t}\,\mathrm{d}t+\sqrt{V_{t}}\,\mathrm{d}W^{1}_{t}+\int_{\mathbb{R}\setminus\{0\}}{\xi_{1}\,\widetilde{N}(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1})},
dVt\displaystyle\mathrm{d}V_{t} =λ(Vtθ)dt+ζVtdWt2,\displaystyle=-\lambda(V_{t}-\theta)\,\mathrm{d}t+\zeta\sqrt{V_{t}}\,\mathrm{d}W^{2}_{t},

where as before λ,θ,ζ>0\lambda,\theta,\zeta>0, δ\delta\in\mathbb{R} and the Brownian motions W1W^{1} and W2W^{2} are correlated with correlation ρ(1,1)\rho\in(-1,1). The jump component is given by N~(Vt,dt,dξ1)=N(Vt,dt,dξ1)n(Vt,dt,dξ1)\widetilde{N}(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1})=N(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1})-n(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1}), where N(Vt,dt,dξ1)N(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1}) is a Poisson random measure independent of W1W^{1} and W2W^{2} with intensity measure n(Vt,dt,dξ1)n(V_{t},\mathrm{d}t,\mathrm{d}\xi_{1}) of the state-dependent form Vtν(dξ1)dtV_{t}\nu(\mathrm{d}\xi_{1})\mathrm{d}t. Here ν(dξ1)\nu(\mathrm{d}\xi_{1}) denotes a Lévy measure on {0}\mathbb{R}\setminus\{0\}. A model of this kind has been proposed in [Bat00] to explain the time-variation of jump-risk implicit in observed option prices.

As in Section 2.2.2, we assume that the support of ν(dξ1)\nu(\mathrm{d}\xi_{1}) is contained in (,0)(-\infty,0) and that the inequality (,0)|ξ1|ν(dξ1)<\int_{(-\infty,0)}|\xi_{1}|\,\nu(\mathrm{d}\xi_{1})<\infty is satisfied. We can identify the parameters in (2.1)–(2.1) as a=0a=0, α11=1\alpha_{11}=1, α12=ρζ\alpha_{12}=\rho\zeta, α22=ζ2\alpha_{22}=\zeta^{2}, β~1=1/2δ\widetilde{\beta}_{1}=-1/2-\delta, β~2=λ\widetilde{\beta}_{2}=-\lambda, b~1=0\widetilde{b}_{1}=0, b~2=b2=λθ\widetilde{b}_{2}=b_{2}=\lambda\theta, m0m\equiv 0 and μ(dξ)=(νδ0)(dξ)\mu(\mathrm{d}\xi)=(\nu\otimes\delta_{0})(\mathrm{d}\xi), where δ0\delta_{0} is the Dirac delta concentrated at 0. Hence we find

β1=12δ(,0)ξ131+ξ12ν(dξ1)\beta_{1}=-\frac{1}{2}-\delta-\int_{(-\infty,0)}\frac{\xi_{1}^{3}}{1+\xi_{1}^{2}}\,\nu(\mathrm{d}\xi_{1})

and β2=λ\beta_{2}=-\lambda, b1=0b_{1}=0. The functions FF and RR for the Bates model are

(15a) F(u,w)\displaystyle F(u,w) =λθw,\displaystyle=\lambda\theta w,
(15b) R(u,w)\displaystyle R(u,w) =12(u2u)+ζ22w2λw+uwρζ+κ~(u),\displaystyle=\frac{1}{2}(u^{2}-u)+\frac{\zeta^{2}}{2}w^{2}-\lambda w+uw\rho\zeta+\widetilde{\kappa}(u),

where κ~(u)=(,0)(eξ1u1u(eξ11))ν(dξ1)\widetilde{\kappa}(u)=\int_{(-\infty,0)}\left(\mathrm{e}^{\xi_{1}u}-1-u\left(\mathrm{e}^{\xi_{1}}-1\right)\right)\,\nu(\mathrm{d}\xi_{1}) and the martingale property (R(1,0)=0R(1,0)=0) was used to determine the value of the parameter δ=(,0)(eξ11ξ1)ν(dξ1)\delta=\int_{(-\infty,0)}\left(\mathrm{e}^{\xi_{1}}-1-\xi_{1}\right)\,\nu(\mathrm{d}\xi_{1}). It is clear that χ(u)=ρζuλ\chi(u)=\rho\zeta u-\lambda and that A1 – A4 are satisfied.

2.2.4. The Barndorff-Nielsen-Shephard (BNS) model

The BNS model was introduced in [BNS01] as a model for asset pricing. Under a risk-neutral measure, it can be defined by the following SDE

dXt\displaystyle\mathrm{d}X_{t} =(δ12Vt)dt+VtdWt+ρdJλt,\displaystyle=(\delta-\frac{1}{2}V_{t})\mathrm{d}t+\sqrt{V_{t}}\,\mathrm{d}W_{t}+\rho\,\mathrm{d}J_{\lambda t},
dVt\displaystyle\mathrm{d}V_{t} =λVtdt+dJλt,\displaystyle=-\lambda V_{t}\,\mathrm{d}t+\mathrm{d}J_{\lambda t},

where λ>0\lambda>0, ρ<0\rho<0 and (Jt)t0(J_{t})_{t\geq 0} is a Lévy subordinator with the Lévy measure ν\nu, i.e. a pure jump Lévy process that increases a.s. The cumulant generating function κ(u)\kappa(u) of (Jt)t0(J_{t})_{t\geq 0} takes the form

(16) κ(u)=(0,)(euξ21)ν(dξ2).\displaystyle\kappa(u)=\int_{(0,\infty)}(\mathrm{e}^{u\xi_{2}}-1)\,\nu(\mathrm{d}\xi_{2}).

To conform with (7) we further assume that (0,)ξ2ν(dξ2)<\int_{(0,\infty)}\xi_{2}\,\nu(\mathrm{d}\xi_{2})<\infty. The drift δ\delta will be determined by the martingale condition for SS. The time-scaling JλtJ_{\lambda t} is introduced in [BNS01] to make the invariant distribution of the variance process independent of λ\lambda. The distinctive features of the BNS model are that the variance process has no diffusion component, i.e. moves purely by jumps, and the negative correlation between variance and price movements is achieved by simultaneous jumps in VV and XX.

It follows from (2.1)–(2.1) and the SDE above that a=α12=α22=0a=\alpha_{12}=\alpha_{22}=0, α11=1\alpha_{11}=1, μ0\mu\equiv 0 and

m(dξ)=I{ξ1=ρξ2}λν(dξ2),m(\mathrm{d}\xi)=I_{\{\xi_{1}=\rho\xi_{2}\}}\lambda\nu(\mathrm{d}\xi_{2}),

where I{ξ1=ρξ2}I_{\{\xi_{1}=\rho\xi_{2}\}} denotes the indicator function of the half-line ξ1=ρξ2\xi_{1}=\rho\xi_{2} in D{0}D\setminus\{0\}. Therefore it follows that β~1=β1=1/2\widetilde{\beta}_{1}=\beta_{1}=-1/2, β~2=β2=λ\widetilde{\beta}_{2}=\beta_{2}=-\lambda, b~2=b2=0\widetilde{b}_{2}=b_{2}=0, b~1=δ+λ(0,)ρξ2ν(dξ2)\widetilde{b}_{1}=\delta+\lambda\int_{(0,\infty)}\rho\xi_{2}\,\nu(\mathrm{d}\xi_{2}) and

b1=δ+λ(0,)ρξ21+(ρξ2)2ν(dξ2).b_{1}=\delta+\lambda\int_{(0,\infty)}\frac{\rho\xi_{2}}{1+(\rho\xi_{2})^{2}}\,\nu(\mathrm{d}\xi_{2}).

The definition of FF in (6) and the martingale condition F(1,0)=0F(1,0)=0 imply that we need to define δ=λκ(ρ)\delta=-\lambda\kappa(\rho), where κ\kappa is the cumulant generating function of JJ given in (16). The BNS model is an affine stochastic volatility model with FF and RR given by

(17a) F(u,w)\displaystyle F(u,w) =λκ(w+ρu)uλκ(ρ),\displaystyle=\lambda\kappa(w+\rho u)-u\lambda\kappa(\rho),
(17b) R(u,w)\displaystyle R(u,w) =12(u2u)λw.\displaystyle=\frac{1}{2}(u^{2}-u)-\lambda w.

We have χ(u)=λ\chi(u)=-\lambda and the assumptions A1 – A4 are clearly satisfied.

3. Large deviation principle and the Gärtner-Ellis theorem

In this section we give a brief review of the key concepts of large deviations for a family of (possibly dependent) random variables (Zt)t1(Z_{t})_{t\geq 1} and state a version of the Gärtner-Ellis theorem (see Theorem 6) that will be used to obtain the asymptotic behaviour of the option prices and implied volatilities. A general reference for all the concepts in this section is [DZ98, Section 2.3].

Let ZtZ_{t} take values in \mathbb{R} and recall that I:(,]I:\mathbb{R}\to(-\infty,\infty] is lower semicontinuous if {x:I(x)α}\{x:I\left(x\right)\leq\alpha\} is closed in \mathbb{R} for any α\alpha\in\mathbb{R} (intuitively for any x0x_{0}\in\mathbb{R} the values of II near x0x_{0} are either close to I(x0)I(x_{0}) or greater than I(x0)I(x_{0})). A nonnegative lower semicontinuous function II is called a rate function. If in addition {x:I(x)α}\{x:I\left(x\right)\leq\alpha\} is compact for any α\alpha\in\mathbb{R}, then II is a good rate function.

Definition 4.

The family (Zt)t1(Z_{t})_{t\geq 1} satisfies the large deviation principle (LDP) with the rate function II if for every Borel set BB\subset\mathbb{R} we have

inf{I(x):xB}lim inft1tlog𝖯[ZtB]lim supt1tlog𝖯[ZtB]inf{I(x):xB¯},-\inf\{I(x):x\in B^{\circ}\}\leq\liminf_{t\to\infty}\frac{1}{t}\log\mathsf{P}\left[Z_{t}\in B\right]\leq\limsup_{t\to\infty}\frac{1}{t}\log\mathsf{P}\left[Z_{t}\in B\right]\leq-\inf\left\{I(x):x\in\overline{B}\right\},

with the convention inf=\inf\emptyset=\infty (the interior BB^{\circ} and closure B¯\overline{B} are relative to the topology of \mathbb{R}).

An important consequence of Definition 4 is that if (Zt)t1(Z_{t})_{t\geq 1} satisfies LDP and II is continuous on B¯\overline{B}, then limtt1log𝖯[ZtB]=inf{I(x):xB}\lim_{t\to\infty}t^{-1}\log\mathsf{P}\left[Z_{t}\in B\right]=-\inf\{I(x):x\in B\}.

The Gärtner-Ellis theorem (Theorem 6) gives sufficient conditions for (Zt)t1(Z_{t})_{t\geq 1} to satisfy the LDP and in that case describes the rate function. Let ΛtZ(u):=log𝖤[euZt]\Lambda_{t}^{Z}(u):=\log\mathsf{E}\left[\mathrm{e}^{uZ_{t}}\right] be a cumulant generating function. Assume that for every uu\in\mathbb{R}

(18) Λ(u)\displaystyle\Lambda(u) :=\displaystyle:= limt1tΛtZ(tu)exists in [,]and0𝒟Λ,\displaystyle\lim_{t\to\infty}\frac{1}{t}\Lambda_{t}^{Z}(tu)\quad\text{exists in }[-\infty,\infty]\qquad\text{and}\qquad 0\in\mathcal{D}_{\Lambda}^{\circ},

where 𝒟Λ:={u:Λ(u)<}\mathcal{D}_{\Lambda}:=\{u\in\mathbb{R}\>:\>\Lambda(u)<\infty\} is the effective domain of Λ\Lambda and 𝒟Λ\mathcal{D}_{\Lambda}^{\circ} is its interior in \mathbb{R}. Since ΛtZ\Lambda_{t}^{Z} is convex (by the Hölder inequality) for every tt, the limit Λ\Lambda is also convex by [Roc70, Theorem 10.8] and the set 𝒟Λ\mathcal{D}_{\Lambda} is an interval. Since Λ(0)=0\Lambda(0)=0, convexity of Λ\Lambda and 0𝒟Λ0\in\mathcal{D}_{\Lambda}^{\circ} imply Λ(u)>\Lambda(u)>-\infty for all uu\in\mathbb{R}. Furthermore the convexity implies that Λ\Lambda is continuous on 𝒟Λ\mathcal{D}_{\Lambda}^{\circ}. The statement in (18) is an important assumption of Gärtner-Ellis theorem (Theorem 6 below), which in particular implies 𝒟Λ\mathcal{D}_{\Lambda}^{\circ}\neq\emptyset. However the converse does not hold in general, i.e. if 0 is a boundary point of a domain 𝒟Λ\mathcal{D}_{\Lambda} with non-empty interior, LDP may still hold true.

A further property of the function Λ:(,]\Lambda:\mathbb{R}\to(-\infty,\infty], which arises as an assumption in Theorem 6, is essential smoothness.

Definition 5.

A convex function Λ:(,]\Lambda:\mathbb{R}\to(-\infty,\infty] is essentially smooth if

  • (a)

    𝒟Λ\mathcal{D}_{\Lambda}^{\circ} is non-empty;

  • (b)

    Λ\Lambda is differentiable in 𝒟Λ\mathcal{D}^{\circ}_{\Lambda};

  • (c)

    Λ\Lambda is steep, in other words it satisfies limn|Λ(un)|=\lim_{n\to\infty}|\Lambda^{\prime}(u_{n})|=\infty for every sequence (un)n(u_{n})_{n\in\mathbb{N}} in 𝒟Λ\mathcal{D}^{\circ}_{\Lambda} that converges to a boundary point of 𝒟Λ\mathcal{D}^{\circ}_{\Lambda}.

The Fenchel-Legendre transform (or convex dual) Λ\Lambda^{*} of Λ\Lambda is defined by the formula

(19) Λ(x)\displaystyle\Lambda^{*}(x) :=\displaystyle:= sup{uxΛ(u):u}forx\displaystyle\sup\{ux-\Lambda(u)\>:\>u\in\mathbb{R}\}\quad\text{for}\quad x\in\mathbb{R}

with an effective domain 𝒟Λ:={x:Λ(x)<}\mathcal{D}_{\Lambda^{*}}:=\{x\in\mathbb{R}\>:\>\Lambda^{*}(x)<\infty\}. The following properties are immediate from the definition:

  • (i)

    0Λ(x)0\leq\Lambda^{*}(x)\leq\infty for all xx\in\mathbb{R}, since Λ(0)=0\Lambda(0)=0;

  • (ii)

    Λ(x)=sup{uxΛ(u):u𝒟Λ}\Lambda^{*}(x)=\sup\{ux-\Lambda(u):u\in\mathcal{D}_{\Lambda}\} for all xx\in\mathbb{R} and hence Λ\Lambda^{*} is convex in the interval 𝒟Λ\mathcal{D}_{\Lambda^{*}} and continuous in the interior 𝒟Λ\mathcal{D}_{\Lambda^{*}}^{\circ};

  • (iii)

    Λ\Lambda^{*} is lower semicontinuous on \mathbb{R} as it is a supremum of continuous (in fact linear) functions. Hence the level sets {x:Λ(x)α}\{x:\Lambda^{*}(x)\leq\alpha\} are closed.

In general 𝒟Λ\mathcal{D}_{\Lambda^{*}} can be strictly contained in \mathbb{R} and Λ\Lambda^{*} can be discontinuous at the boundary of 𝒟Λ\mathcal{D}_{\Lambda^{*}} (see [DZ98, Section 2.3] for elementary examples of such rate functions). Assumption (18) implies that for any δ>0\delta>0, such that (δ,δ)𝒟Λ(-\delta,\delta)\subset\mathcal{D}_{\Lambda}^{\circ}, and c=sup{Λ(u):u[δ,δ]}c=\sup\{\Lambda(u):u\in[-\delta,\delta]\} we have

(20) Λ(x)sup{uxΛ(u):u[δ,δ]}δ|x|c.\displaystyle\Lambda^{*}(x)\quad\geq\quad\sup\{ux-\Lambda(u):u\in[-\delta,\delta]\}\quad\geq\quad\delta|x|-c.

Hence the set {x:Λ(x)α}\{x:\Lambda^{*}(x)\leq\alpha\} is compact for any α\alpha\in\mathbb{R} and therefore Λ\Lambda^{*} is a good rate function.

Remarks.

(A) If Λ\Lambda is strictly convex, differentiable on 𝒟Λ\mathcal{D}_{\Lambda}^{\circ} and steep, which is the case in the applications in this paper, then 𝒟Λ=\mathcal{D}_{\Lambda^{*}}=\mathbb{R} and for each xx\in\mathbb{R} the equation Λ(u)=x\Lambda^{\prime}(u)=x has a unique solution uxu_{x} in 𝒟Λ\mathcal{D}_{\Lambda}^{\circ}. Furthermore the formula

(21) Λ(x)\displaystyle\Lambda^{*}(x) =\displaystyle= xuxΛ(ux)\displaystyle xu_{x}-\Lambda(u_{x})

holds. This reduces the computation of Λ(x)\Lambda^{*}(x) to finding the unique root of the equation Λ(u)=x\Lambda^{\prime}(u)=x, where the strictly increasing function Λ\Lambda^{\prime} is in most applications known in closed form.
(B) If (Zt)t1(Z_{t})_{t\geq 1} satisfies (18) and the function Λ\Lambda satisfies the assumptions of Remark (A) and is twice differentiable with Λ′′(u)>0\Lambda^{\prime\prime}(u)>0 for all u𝒟Λu\in\mathcal{D}_{\Lambda}^{\circ}, then (21) implies that the Fenchel-Legendre transform Λ\Lambda^{*} is differentiable with the derivative

(22) (Λ)(x)\displaystyle\left(\Lambda^{*}\right)^{\prime}(x) =\displaystyle= (Λ)1(x) for all x.\displaystyle\left(\Lambda^{\prime}\right)^{-1}(x)\quad\text{ for all }\quad x\in\mathbb{R}.

In particular (22) implies that Λ\Lambda^{*} is strictly convex on \mathbb{R} and that its global minimum is attained at the unique point xx^{*} given by

x\displaystyle x^{*} =\displaystyle= Λ(0).\displaystyle\Lambda^{\prime}(0).

We state a simple version of the Gärtner-Ellis theorem (for the proof see [DZ98, Section 2.3]).

Theorem 6.

Let (Zt)t1(Z_{t})_{t\geq 1} be a family of random variables that satisfies assumption (18) with the limiting cumulant generating function Λ:(,]\Lambda:\mathbb{R}\to(-\infty,\infty]. If Λ\Lambda is essentially smooth and lower semicontinuous, then the LDP holds for (Zt)t1(Z_{t})_{t\geq 1} with the good rate function Λ\Lambda^{*}.

4. Limiting cumulant generating function in affine stochastic volatility models

4.1. Non-degenerate affine stochastic volatility processes

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process (see Definition 2). The goal of the present section is to describe the limiting cumulant generating function hh of the family of variables (Xt/t)t1(X_{t}/t)_{t\geq 1}, defined by

(23) h(u)\displaystyle h(u) :=\displaystyle:= limt1tlog𝖤[euXt]\displaystyle\lim_{t\to\infty}\frac{1}{t}\log\mathsf{E}\left[\mathrm{e}^{uX_{t}}\right]

for every uu\in\mathbb{R} where the limit in (23) exists as an extended real number. The function hh will determine the limiting implied volatility smile of the model S=eXS=\mathrm{e}^{X}. To ensure that hh is finite on an interval that contains [0,1][0,1], which is key for establishing the LDP, a further assumption will be required:

A5:

χ(0)<0\chi(0)<0 and χ(1)<0\chi(1)<0, where χ\chi is given in Definition 1.

This assumption will also imply that hh can be uniquely extended to a cumulant generating function of an infinitely divisible random variable.

In order to apply the Gärtner-Ellis theorem in our setting, we need to answer the following three questions: is hh well-defined as an extended real number by (23) for every uu\in\mathbb{R}, does the effective domain 𝒟h\mathcal{D}_{h} contain [0,1][0,1] in its interior and is hh essentially smooth? Answers to these questions play a crucial role in establishing the large deviation principle, via Theorem 6, for affine stochastic volatility models. Theorem 10 and Corollary 11, proved in this section, provide easy to check sufficient conditions for the affirmative answers to hold.

It is shown in  [KR11] that the function hh can be obtained from the functions FF and RR without the explicit knowledge of ϕ\phi and ψ\psi (see Section 2 for definition of ψ,ϕ\psi,\phi). Lemma 7 and Theorem 8, taken from [KR11, Lemma 3.2 and Theorem 3.4], describe certain properties of the limiting cumulant generating function hh, which are needed in Section 5 but are insufficient to guarantee the essential smoothness of hh. The main contribution of the present section is Theorem 10, which identifies sufficient conditions for the process (X,V)(X,V) that imply essential smoothness of the function hh. The conditions in Theorem 10 are easy to apply to the models of Section 2.2, which will allow us to find their limiting implied volatility smiles.

Lemma 7.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies assumption A5. Then there exist a maximal interval \mathcal{I} and a unique convex function w:w:\mathcal{I}\to\mathbb{R} such that wC()C1()w\in C(\mathcal{I})\cap C^{1}(\mathcal{I}^{\circ}) and

R(u,w(u))=0for allu,R(u,w(u))=0\qquad\text{for all}\quad u\in\mathcal{I},

where RR is given in (5) (see also (6)). Furthermore we have

  1. (a)

    [0,1][0,1]\subseteq\mathcal{I} and 2R(u,w(u))<0\partial_{2}R(u,w(u))<0 for all uu\in\mathcal{I}^{\circ};

  2. (b)

    w(0)=w(1)=0w(0)=w(1)=0 and w(u)<0w(u)<0 for all u(0,1)u\in(0,1);

  3. (c)

    w(u)>0w(u)>0 for all u[0,1]u\in\mathcal{I}\setminus[0,1].

Remarks.

(i) The proof of Lemma 7 in [KR11] is based on the analysis of the qualitative properties of the generalized Riccati equations in (4).

(ii) The function uw(u)u\mapsto w(u) from Lemma 7 can be extended naturally to a lower semicontinuous function w:(,]w:\mathbb{R}\to(-\infty,\infty] by w()=w\left(\mathbb{R}\setminus\mathcal{I}\right)=\infty. Then the extension, again denoted by w(u)w(u), has the following properties:

  • ww is convex with effective domain 𝒟w=\mathcal{D}_{w}=\mathcal{I} and ={u:w(u)<}\mathcal{I}=\{u\in\mathbb{R}:w(u)<\infty\};

  • the maximality of \mathcal{I} implies that for uu\in\mathbb{R}\setminus\mathcal{I} there exists no ww^{*}\in\mathbb{R} such that R(u,w)=0R(u,w^{*})=0.

The next theorem, proved in [KR11, Theorem 3.4], describes further properties of the function uw(u)u\mapsto w(u) and specifies its relationship to the limiting cumulant generating function hh defined in (23).

Theorem 8.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies assumption A5 and let w(u)w(u) be given by Lemma 7. Then the function h(u)h(u) defined in (23) satisfies

(24) h(u)=F(u,w(u))for anyu𝒥:={s:F(s,w(s))<},\displaystyle h(u)=F(u,w(u))\quad\text{for any}\quad u\in\mathcal{J}:=\{s\in\mathcal{I}:F(s,w(s))<\infty\},

where FF is defined in (5) (see also (6)). Furthermore the inclusions hold: [0,1]𝒥[0,1]\subseteq\mathcal{J}\subseteq\mathcal{I}. The functions w(u)w(u) and h(u)h(u) can be extended uniquely to cumulant generating functions of infinitely divisible random variables and

(25a) limtψ(t,u,0)\displaystyle\lim_{t\to\infty}\psi(t,u,0) =w(u)for allu;\displaystyle=w(u)\quad\text{for all}\quad u\in\mathcal{I};
(25b) limt1tϕ(t,u,0)\displaystyle\lim_{t\to\infty}\frac{1}{t}\phi(t,u,0) =h(u)for allu𝒥.\displaystyle=h(u)\quad\text{for all}\quad u\in\mathcal{J}.
Remark.

Since w(u)w(u) and h(u)h(u) can be extended to cumulant generating functions of some (infinitely divisible) random variables it follows that:

  • ww (resp. hh) is continuously differentiable in the interior of \mathcal{I} (resp. 𝒥\mathcal{J});

  • either h′′(u)>0h^{\prime\prime}(u)>0 for all u𝒥u\in\mathcal{J}^{\circ} or h(u)=0h(u)=0 for all uu\in\mathbb{R} (this follows from (24), (b) in Lemma 7 and assumption A3).

We say that RR explodes at the boundary if limnR(un,wn)=\lim_{n\to\infty}R(u_{n},w_{n})=\infty for any sequence ((un,wn))n\left((u_{n},w_{n})\right)_{n\in\mathbb{N}} in the interior 𝒟R\mathcal{D}_{R}^{\circ} that tends to a point in the boundary of 𝒟R\mathcal{D}_{R} (both the boundary and the interior of 𝒟R\mathcal{D}_{R} are relative to the topology of 2\mathbb{R}^{2}) or equivalently 𝒟R\mathcal{D}_{R} is open. By Proposition 3 (A), the gradient F=(1F,2F)\nabla F=(\partial_{1}F,\partial_{2}F) is continuous on 𝒟F\mathcal{D}_{F}^{\circ}. Analogously to the one-dimensional case (see (c) in Definition 5), we say that FF is steep if limnF(un,wn)=\lim_{n\to\infty}\|\nabla F(u_{n},w_{n})\|=\infty for any sequence ((un,wn))n\left((u_{n},w_{n})\right)_{n\in\mathbb{N}} in the interior 𝒟F\mathcal{D}_{F}^{\circ} that tends to a point in the boundary of 𝒟F\mathcal{D}_{F}. It is clear that if FF explodes at the boundary, it is also steep but the converse may not be true.

Before we state and prove the main results of this section (Theorem 10 and Corollary 11), we establish Lemma 9, which states that in an affine stochastic volatility model, the limiting cumulant generating function hh cannot be identically equal to zero. This property will play an important role in understanding the limiting behaviour of the implied volatility smile (see e.g. Theorem 13).

Lemma 9.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies assumption A5 and let hh be given by (23). Assume further that the interior 𝒟F\mathcal{D}_{F}^{\circ} of the effective domain of FF contains the set {(0,0),(1,0)}\{(0,0),(1,0)\} and that F(u,w)0F(u,w)\neq 0 for some (u,v)𝒟F(u,v)\in\mathcal{D}_{F}. Then h(u)>0h(u)>0 for all u𝒥[0,1]u\in\mathcal{J}\setminus[0,1] and h(u)<0h(u)<0 for all u(0,1)u\in(0,1). Furthermore we have h′′(u)>0h^{\prime\prime}(u)>0 for all u𝒥u\in\mathcal{J}^{\circ}.

Proof.

Note that since hh can be extended to a cumulant generating function of a random variable by Theorem 8, it is smooth in 𝒥\mathcal{J}^{\circ}. Since hh is either identically equal to zero or strictly convex on 𝒥\mathcal{J} by the remark following Theorem 8, the statement h′′(u)>0h^{\prime\prime}(u)>0 for all u𝒥u\in\mathcal{J}^{\circ} follows if we prove that h(u)<0h(u)<0 for some u(0,1)u\in(0,1).

The function uF(u,0)u\mapsto F(u,0) is convex by (B) of Proposition 3. Furthermore it is either (I) strictly convex or (II) identically equal to zero (by A3). We analyse both cases.

(I) Strict convexity and A3 imply that for u(0,1)u\in(0,1) we have F(u,0)<0F(u,0)<0. The same argument implies that for u[0,1]u\in\mathbb{R}\setminus[0,1], such that (u,0)𝒟F(u,0)\in\mathcal{D}_{F}^{\circ}, the inequality F(u,0)>0F(u,0)>0 holds. The Lévy-Khintchine representation of FF in (6) implies that

(26) 2F(u,w)\displaystyle\partial_{2}F(u,w) =\displaystyle= b2+D{0}ξ2euξ1+wξ2m(dξ)\displaystyle b_{2}+\int_{D\setminus\{0\}}\xi_{2}\,\mathrm{e}^{u\xi_{1}+w\xi_{2}}\,m(\mathrm{d}\xi)

for any point in the interior of the effective domain 𝒟F\mathcal{D}_{F}. It is clear from (26) that 2F0\partial_{2}F\geq 0 on 𝒟F\mathcal{D}_{F}^{\circ}. Lemma 7 implies that for u(0,1)u\in(0,1) we have w(u)<0w(u)<0. Identity (24) in Theorem 8 yields

h(u)=F(u,w(u))=F(u,0)w(u)02F(u,z)dzF(u,0)<0.h(u)=F(u,w(u))=F(u,0)-\int_{w(u)}^{0}\partial_{2}F(u,z)\,\mathrm{d}z\leq F(u,0)<0.

The last inequality follows from the strict convexity of uF(u,0)u\mapsto F(u,0). If u𝒥[0,1]u\in\mathcal{J}^{\circ}\setminus[0,1], then (u,0)𝒟F(u,0)\in\mathcal{D}_{F}^{\circ} and an analogous argument implies that h(u)>0h(u)>0. The inequality at the boundary points of the interval 𝒥\mathcal{J} follows from the convexity of hh.

(II) Assume now that uF(u,0)u\mapsto F(u,0) is identically equal to zero. For any (u,0)(u,0) in the interior of the effective domain of FF, the Lévy-Khintchine representation of FF in (6) yields

12F(u,0)=a11+D{0}ξ12euξ1m(dξ)=0.\partial_{1}^{2}F(u,0)=a_{11}+\int_{D\setminus\{0\}}\xi_{1}^{2}\,\mathrm{e}^{u\xi_{1}}\,m(\mathrm{d}\xi)=0.

This implies a11=0a_{11}=0 and m(dξ)=(δ0ν)(dξ)m(\mathrm{d}\xi)=(\delta_{0}\otimes\nu)(\mathrm{d}\xi), where ν(dξ2)\nu(\mathrm{d}\xi_{2}) is a Lévy measure on (0,)(0,\infty) with integrable small jumps and δ0\delta_{0} is the Dirac delta. The condition F(1,0)=0F(1,0)=0 in A3 and the representation of FF in (6) yield b1=0b_{1}=0. Hence we have

F(u,w)=b2w+0{0}(ewξ21)ν(dξ2).F(u,w)=b_{2}w+\int_{\mathbb{R}_{\geqslant 0}\setminus\{0\}}\left(\mathrm{e}^{w\xi_{2}}-1\right)\,\nu(\mathrm{d}\xi_{2}).

Since by assumption there exists (u,v)𝒟F(u,v)\in\mathcal{D}_{F} such that F(u,w)0F(u,w)\neq 0, either b2>0b_{2}>0 or ν0\nu\neq 0 holds. Therefore identity (24) in Theorem 8, Lemma 7 and this representation of FF conclude the proof. ∎

Remark.

The assumption {(0,0),(1,0)}𝒟F\{(0,0),(1,0)\}\subset\mathcal{D}_{F}^{\circ} in Lemma 9 ensures that the interiors of the effective domains of FF and hh are non-empty. It may not be necessary for Lemma 9 to hold. However, the assumption is crucial in Theorem 10 and hence does not restrict the applicability of Lemma 9 in our setting.

Theorem 10.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies assumption A5 and suppose that the function wF(0,w)w\mapsto F(0,w), where FF is defined in (5), is not identically equal to zero. If RR explodes at the boundary (i.e. 𝒟R\mathcal{D}_{R} is open), FF is steep and {(0,0),(1,0)}𝒟F\{(0,0),(1,0)\}\subset\mathcal{D}_{F}^{\circ}, then the function h(u)h(u) is well-defined by (23) as an extended real number for every uu\in\mathbb{R} and its effective domain is given by 𝒟h=𝒥\mathcal{D}_{h}=\mathcal{J} (see (24) for the definition of interval 𝒥\mathcal{J}). Furthermore hh is essentially smooth and the set {0,1}\{0,1\} is contained in the interior 𝒟h\mathcal{D}_{h}^{\circ} (relative to \mathbb{R}) of 𝒟h\mathcal{D}_{h}.

Corollary 11.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies assumption A5 and assume that wF(0,w)w\mapsto F(0,w) is not identically equal to zero. If either of the following conditions holds

  1. (i)

    μ\mu has exponential moments of all orders, FF is steep, and 𝒟F\mathcal{D}_{F}^{\circ} contains (0,0)(0,0) and (1,0)(1,0),

  2. (ii)

    (X,V)(X,V) is a diffusion,

then the function hh is well-defined by (23) for every uu\in\mathbb{R} with effective domain 𝒟h=𝒥\mathcal{D}_{h}=\mathcal{J}. Moreover hh is essentially smooth and {0,1}𝒟h\{0,1\}\subset\mathcal{D}_{h}^{\circ}.

Proof of Corollary 11. Note that either of the conditions (i) or (ii) implies that 𝒟R=2\mathcal{D}_{R}=\mathbb{R}^{2} and hence 𝒟R\mathcal{D}_{R} is open. Therefore (i) and the assumptions of Corollary 11 imply the assumptions of Theorem 10. If (ii) holds, then (X,V)(X,V) is a diffusion and

F(u,w)=a11u22+b1u+b2wwitha11,b20andb1.F(u,w)=a_{11}\frac{u^{2}}{2}+b_{1}u+b_{2}w\qquad\text{with}\quad a_{11},b_{2}\geq 0\quad\text{and}\quad b_{1}\in\mathbb{R}.

Clearly 𝒟F=2\mathcal{D}_{F}^{\circ}=\mathbb{R}^{2} contains the set {(0,0),(1,0)}\{(0,0),(1,0)\} and FF is steep if b2b_{2} is non-zero. In the case b2=0b_{2}=0, the map wF(0,w)w\mapsto F(0,w) is identically equal to zero, which contradicts the assumption in Corollary 11. Thus Corollary 11 follows from Theorem 10 \square

Proof of Theorem 10. The proof of this theorem is in two steps. In step (I) we show that {0,1}𝒥\{0,1\}\subset\mathcal{J}^{\circ} and that, if we extend h|𝒥h|_{\mathcal{J}} by ++\infty to 𝒥\mathbb{R}\setminus\mathcal{J}, we obtain an essentially smooth convex function. In step (II) of the proof we show that the limit in definition (23) exists for any uu\in\mathbb{R} as an extended real number and that definition of hh in (23) agrees for every uu\in\mathbb{R} with the extension of h|𝒥h|_{\mathcal{J}} from the first part of the proof.

Step (I). Throughout this step we abuse notation by using hh to denote the extension of h|𝒥h|_{\mathcal{J}} to \mathbb{R} described above. Theorem 8 and the remark following it imply that hh is essentially smooth (see Definition 5) if it is steep. We will prove the steepness of hh at the right endpoint u+=sup{u:u𝒥}u_{+}=\sup\{u:u\in\mathcal{J}\} of the interval 𝒥\mathcal{J} and show that 1𝒥1\in\mathcal{J}^{\circ}. The left endpoint u=inf{u:u𝒥}u_{-}=\inf\{u:u\in\mathcal{J}\} and the fact 0𝒥0\in\mathcal{J}^{\circ} can be treated by a completely symmetrical argument.

Let (un)n(u_{n})_{n\in\mathbb{N}} be a sequence in 𝒥\mathcal{J}^{\circ} converging to u+u_{+}. We use the shorthand notation wn=w(un)w_{n}=w(u_{n}) and w+=limnwnw_{+}=\lim_{n\to\infty}w_{n}, where uw(u)u\mapsto w(u) is the function given in Lemma 7 (note that the limit w+w_{+} exists but may be infinite since uw(u)u\mapsto w(u) is a cumulant generating function of a random variable and 𝒥=𝒟w\mathcal{J}\subset\mathcal{I}=\mathcal{D}_{w}). Since uw(u)u\mapsto w(u) is convex on 𝒟w\mathcal{D}_{w}, the value w+w_{+} is independent of the choice of sequence (un)n(u_{n})_{n\in\mathbb{N}}.

Claim 1. The inequalities u+>1u_{+}>1 and w+>0w_{+}>0 hold.
Indeed, since R(1,0)=0R(1,0)=0 by assumption A3, we get that (1,0)𝒟R=𝒟R(1,0)\in\mathcal{D}_{R}=\mathcal{D}_{R}^{\circ}. Assume now that u+=1u_{+}=1. Then by Lemma 7 we have w+=0w_{+}=0 and (u+,w+)𝒟R(u_{+},w_{+})\in\mathcal{D}_{R}^{\circ}. Since RR is continuously differentiable in 𝒟R\mathcal{D}_{R}^{\circ} and 2R(1,0)=χ(1)<0\partial_{2}R(1,0)=\chi(1)<0 by assumption A5, the implicit function theorem and Lemma 7 imply that u+u_{+} is in the set 𝒟w=\mathcal{D}_{w}^{\circ}=\mathcal{I}^{\circ}. Since (1,0)𝒟F(1,0)\in\mathcal{D}_{F}^{\circ}, there exists uu\in\mathcal{I}^{\circ} such that u>u+u>u_{+} and (u,0)𝒟F(u,0)\in\mathcal{D}_{F}^{\circ}. Identity (24) in Theorem 8 therefore implies that h(u)<h(u)<\infty, which contradicts the definition of u+u_{+}. Therefore u+>1u_{+}>1. Lemma 7 implies that the sequence (wn)n(w_{n})_{n\in\mathbb{N}} is eventually (certainly when un>1u_{n}>1) non-decreasing and strictly positive. This yields that w+>0w_{+}>0 and the claim follows.

Discarding finitely many elements we may assume that un>1u_{n}>1 and wn>0w_{n}>0 for all nn. If u+u_{+} is infinite, it is not in the boundary of 𝒥\mathcal{J} and the steepness of hh follows. If u+u_{+} is finite but w+w_{+} is infinite, identity (24) and the assumption that wF(0,w)w\mapsto F(0,w) is non-zero imply limnh(un)=\lim_{n\to\infty}h(u_{n})=\infty. The steepness of hh follows from the convexity of hh. Therefore in the rest of the proof we can assume

(27) u+(1,)andw+(0,)u_{+}\in(1,\infty)\qquad\text{and}\qquad w_{+}\in(0,\infty)

without loss of generality.

Claim 2. The following statements hold true:

  1. (a)

    if uu\in\mathcal{I}^{\circ}, where \mathcal{I} is defined in Lemma 7, then (u,w(u))𝒟R(u,w(u))\in\mathcal{D}_{R}^{\circ} and

    (28) 0=1R(u,w(u))+2R(u,w(u))w(u);0=\partial_{1}R(u,w(u))+\partial_{2}R(u,w(u))w^{\prime}(u);
  2. (b)

    if u𝒥(1,)u\in\mathcal{J}^{\circ}\cap(1,\infty), where 𝒥\mathcal{J} is defined in Theorem 8, then (u,w(u))𝒟F(u,w(u))\in\mathcal{D}_{F}^{\circ} and

    (29) h(u)=1F(u,w(u))+2F(u,w(u))w(u).h^{\prime}(u)=\partial_{1}F(u,w(u))+\partial_{2}F(u,w(u))w^{\prime}(u).

The statement in (a) follows from Lemma 7, assumption 𝒟R=𝒟R\mathcal{D}_{R}=\mathcal{D}_{R}^{\circ} and the chain rule. To prove the first statement in (b), note that u𝒥(1,)u\in\mathcal{J}^{\circ}\cap(1,\infty)\subset\mathcal{I}^{\circ} and hence [0,1]\mathcal{I}^{\circ}\setminus[0,1]\neq\emptyset. Lemma 7 therefore implies that the function w:𝒥w:\mathcal{J}\to\mathbb{R} is strictly convex with w(0)=w(1)=0w(0)=w(1)=0 and therefore strictly increasing on 𝒥(1,)\mathcal{J}^{\circ}\cap(1,\infty). Pick u𝒥(1,)u^{\prime}\in\mathcal{J}^{\circ}\cap(1,\infty) such that u>uu^{\prime}>u and note that (u,w(u))𝒟F(u^{\prime},w(u^{\prime}))\in\mathcal{D}_{F} (by the definition of 𝒥\mathcal{J}) and (u,0)𝒟F(u^{\prime},0)\in\mathcal{D}_{F} (by representation (6) and 0w(u)0\leq w(u^{\prime})). Assumption (1,0)𝒟F(1,0)\in\mathcal{D}_{F} in the theorem and the fact w(u)<w(u)w(u)<w(u^{\prime}) imply that the point (u,w(u))(u,w(u)) lies in the interior of the triangle with vertices (u,w(u)),(1,0),(u,0)(u^{\prime},w(u^{\prime})),(1,0),(u^{\prime},0) in the convex set 𝒟F\mathcal{D}_{F}. Therefore (u,w(u))𝒟F(u,w(u))\in\mathcal{D}_{F}^{\circ}. Equality (29) follows by the chain rule. This proves the claim.

Claim 3. The following holds for any strictly increasing sequence (un)n(u_{n})_{n\in\mathbb{N}} with limit u+u_{+}:

  1. (a)

    if u+=sup𝒥=supu_{+}=\sup\mathcal{J}=\sup\mathcal{I}, then

    |w(un)|asn;|w^{\prime}(u_{n})|\to\infty\qquad\text{as}\quad n\to\infty;
  2. (b)

    if u+=sup𝒥<supu_{+}=\sup\mathcal{J}<\sup\mathcal{I}, then

    F(un,wn)asn.\|\nabla F(u_{n},w_{n})\|\to\infty\qquad\text{as}\quad n\to\infty.

To prove the claim, assume that the conclusion of (a) does not hold. Since the sequence (w(un))n(w^{\prime}(u_{n}))_{n\in\mathbb{N}} is non-decreasing by Lemma 7, there exists a finite positive number, denoted by w(u+)>0w^{\prime}(u_{+})>0, such that limnw(un)=w(u+)\lim_{n\to\infty}w^{\prime}(u_{n})=w^{\prime}(u_{+}). Claim 2(a), applied to u=unu=u_{n}, implies (un,wn)𝒟R(u_{n},w_{n})\in\mathcal{D}_{R}^{\circ} for all nn\in\mathbb{N} and hence by (27) (u+,w+)(u_{+},w_{+}) is in the closure of 𝒟R\mathcal{D}_{R}. However (u+,w+)(u_{+},w_{+}) cannot be in the boundary of 𝒟R\mathcal{D}_{R} since RR explodes at the boundary by assumption and it holds limnR(un,wn)=0\lim_{n\to\infty}R(u_{n},w_{n})=0 (recall that R(un,wn)=0R(u_{n},w_{n})=0 for all nn\in\mathbb{N}). Therefore (u+,w+)𝒟R(u_{+},w_{+})\in\mathcal{D}_{R}^{\circ}. The derivatives 1R,2R\partial_{1}R,\partial_{2}R are hence continuous at (u+,w+)(u_{+},w_{+}) and, in the limit as nn\to\infty, formula (28) and the fact w(u+)>0w^{\prime}(u_{+})>0 imply

2R(u+,w+)=1R(u+,w+)w(u+).\partial_{2}R(u_{+},w_{+})=-\frac{\partial_{1}R(u_{+},w_{+})}{w^{\prime}(u_{+})}.

Therefore either 0=1R(u+,w+)=2R(u+,w+)0=\partial_{1}R(u_{+},w_{+})=\partial_{2}R(u_{+},w_{+}) or both partial derivatives at (u+,w+)(u_{+},w_{+}) are non-zero. Suppose the former. For an arbitrary u(0,1)u\in(0,1), the convexity of RR yields

R(u,0)=R(u+,w+)R(u,0)R(u+,w+)(u+u,w+)=0.-R(u,0)=R(u_{+},w_{+})-R(u,0)\leq\nabla R(u_{+},w_{+})\cdot(u_{+}-u,w_{+})^{\prime}=0.

Since R(u,0)<0R(u,0)<0 (see assumptions A3 and A4), this leads to a contradiction. Hence 1R(u+,w+)\partial_{1}R(u_{+},w_{+}) and 2R(u+,w+)\partial_{2}R(u_{+},w_{+}) are non-zero and related by the equality above. By the implicit function theorem there exists an open interval NN containing u+u_{+} and a function w~:N\widetilde{w}:N\to\mathbb{R}, such that R(u,w~(u))=0R(u,\widetilde{w}(u))=0 for all uNu\in N. This contradicts the maximality of \mathcal{I} and proves Claim 3(a). Note that under assumption of Claim 3(b), (u+,w+)(u_{+},w_{+}) must be a boundary point of 𝒟F\mathcal{D}_{F}. Since FF is steep this implies F(un,wn)\|\nabla F(u_{n},w_{n})\|\to\infty and the claim follows.

Theorem 10 follows easily if F(un,wn)\|\nabla F(u_{n},w_{n})\|\to\infty as nn\to\infty. Indeed, assumption (27) and Lemma 7 imply that the sequence w(un)>0w^{\prime}(u_{n})>0 is strictly increasing and positive for all large nn. Since F(0,0)=F(1,0)=0F(0,0)=F(1,0)=0, Proposition 3 (B) implies that 1F(1,0)0\partial_{1}F(1,0)\geq 0. The Lévy-Khintchine representation of FF in (6) implies 2F(u,w)0\partial_{2}F(u,w)\geq 0 for all (u,w)𝒟F(u,w)\in\mathcal{D}_{F}. Since the gradient of the convex function FF is monotone on 𝒟F\mathcal{D}_{F}^{\circ} and (un,wn),(1,0)𝒟F(u_{n},w_{n}),(1,0)\in\mathcal{D}_{F}^{\circ} for all nn, we find

1F(un,wn)(un1)+2F(un,wn)wn1F(1,0)(un1)+2F(1,0)wn0.\partial_{1}F(u_{n},w_{n})(u_{n}-1)+\partial_{2}F(u_{n},w_{n})w_{n}\geq\partial_{1}F(1,0)(u_{n}-1)+\partial_{2}F(1,0)w_{n}\geq 0.

Therefore by (29) we obtain

(30) h(un)2F(un,wn)(w(un)wnun1).h^{\prime}(u_{n})\geq\partial_{2}F(u_{n},w_{n})\left(w^{\prime}(u_{n})-\frac{w_{n}}{u_{n}-1}\right).

If |2F(un,wn)||\partial_{2}F(u_{n},w_{n})|\to\infty as nn\to\infty, the steepness of hh at u+u_{+} follows from (27), (30) and the fact that w(un)w^{\prime}(u_{n}) is strictly positive and increasing. If |1F(un,wn)||\partial_{1}F(u_{n},w_{n})|\to\infty as nn\to\infty, then, since 2F0\partial_{2}F\geq 0 on 𝒟F\mathcal{D}_{F}^{\circ}, formula (29) implies Theorem 10.

If F(un,wn)\|\nabla F(u_{n},w_{n})\| does not tend to infinity as nn\to\infty, the following facts hold: (u+,w+)𝒟F(u_{+},w_{+})\in\mathcal{D}_{F}^{\circ} (since (u+,w+)(u_{+},w_{+}) is in the closure of 𝒟F\mathcal{D}_{F} and FF is steep), |w(un)||w^{\prime}(u_{n})|\to\infty as nn\to\infty (by Claim 3) and 2F(u,w)0\partial_{2}F(u,w)\geq 0 for all (u,w)𝒟F(u,w)\in\mathcal{D}_{F}^{\circ} (by Lévy-Khintchine representation (6) of FF). The next claim plays a key role in the proof of steepness of hh.
Claim 4. If F(un,wn)\|\nabla F(u_{n},w_{n})\| does not tend to infinity as nn\to\infty, then 2F(u+,w+)>0\partial_{2}F(u_{+},w_{+})>0. In particular there exists δ>0\delta>0 such that 2F(un,wn)>δ\partial_{2}F(u_{n},w_{n})>\delta for all large nn\in\mathbb{N}.

Note first that 2F(u+,w+)\partial_{2}F(u_{+},w_{+}) is well-defined since (u+,w+)𝒟F(u_{+},w_{+})\in\mathcal{D}_{F}^{\circ}. If 2F(u+,w+)=0\partial_{2}F(u_{+},w_{+})=0, differentiation under the integral in (6) implies that b2=0b_{2}=0 and the support of mm is contained in the set ×{0}\mathbb{R}\times\{0\}. This would imply that F(0,w)=0F(0,w)=0 for all ww\in\mathbb{R}, which contradicts the assumption in the theorem. Hence the claim follows.

To conclude the proof of Step (I), it remains to note that equality (29) applied at unu_{n} together with Claim 4 yield the steepness of hh in the case F(un,wn)\|\nabla F(u_{n},w_{n})\| does not tend to infinity.

Step (II). We now prove that for any u𝒥u\in\mathbb{R}\setminus\mathcal{J}, the limit in (23) is equal to ++\infty. This will conclude the proof of Theorem 10.

Let tn0t_{n}\downarrow 0 and define hn(u)=1tnlog𝖤euXtnh_{n}(u)=\frac{1}{t_{n}}\log\mathsf{E}{\mathrm{e}^{uX_{t_{n}}}} for all uu\in\mathbb{R}. We know that limnhn(u)=h(u)\lim_{n\to\infty}h_{n}(u)=h(u) for all u𝒥u\in\mathcal{J}. Moreover by Step (I), h(u)h(u) is steep at the boundary of 𝒥\mathcal{J} and 0𝒥0\in\mathcal{J}^{\circ}. Since XtX_{t} is infinitely divisible for all t0t\geq 0 (see [DFS03, Theorem 2.15]), there exist random variables X^n\widehat{X}_{n} such that hn(u)=log𝖤euX^nh_{n}(u)=\log\mathsf{E}{\mathrm{e}^{u\widehat{X}_{n}}} (i.e. hnh_{n} is the cumulant generating function of X^n\widehat{X}_{n}). Therefore there exists a random variable XX such that X^nX\widehat{X}_{n}\to X in distribution and, if we define H(u)=log𝖤euXH(u)=\log\mathsf{E}{\mathrm{e}^{uX}} for all uu\in\mathbb{R}, the equality H(u)=h(u)H(u)=h(u) holds on 𝒥\mathcal{J}. Since HH is a cumulant generating function, it is lower semicontinuous and convex, and in particular continuously differentiable in the interior of its effective domain 𝒟H\mathcal{D}_{H}. But hh is steep and hence non-differentiable at the boundary of 𝒥\mathcal{J}. Therefore it follows that 𝒟H=𝒥\mathcal{D}_{H}=\mathcal{J} and H(u)=H(u)=\infty for all u𝒥u\in\mathbb{R}\setminus\mathcal{J}. However for all u𝒥u\in\mathbb{R}\setminus\mathcal{J}, the Skorokhod representation theorem and Fatou’s lemma imply

lim infn𝖤euX^n𝖤euX=eH(u)=.\liminf_{n\to\infty}\mathsf{E}{\mathrm{e}^{u\widehat{X}_{n}}}\geq\mathsf{E}{\mathrm{e}^{uX}}=\mathrm{e}^{H(u)}=\infty.

Hence the equality limn1tnlog𝖤euXtn=\lim_{n\to\infty}\frac{1}{t_{n}}\log\mathsf{E}{\mathrm{e}^{uX_{t_{n}}}}=\infty holds for u𝒥u\in\mathbb{R}\setminus\mathcal{J}. This concludes the proof of Theorem 10 \square

4.2. Degenerate affine stochastic volatility models

The remark following Definition 2 implies that in the case of a degenerate affine stochastic volatility process (X,V)(X,V), the model S=eXS=\mathrm{e}^{X} is an exponential Lévy model (note also that A5 in this setting fails). Therefore Definition 2 and (6) imply that the characteristic exponent h(u):=F(u,0)h(u):=F(u,0) of XX possesses a Lévy-Khintchine characteristic triplet (δ,σ2,ν)(\delta,\sigma^{2},\nu), where δ,σ\delta,\sigma\in\mathbb{R} and ν\nu a Lévy measure on {0}\mathbb{R}\setminus\{0\}, and satisfies

h(u)\displaystyle h(u) =\displaystyle= log𝖤[exp(uX1)|X0=0]\displaystyle\log\mathsf{E}\left[\exp\left(uX_{1}\right)|X_{0}=0\right]
=\displaystyle= uδ+12σ2u2+{0}(euξ11uξ1ξ12+1)ν(dξ1)\displaystyle u\delta+\frac{1}{2}\sigma^{2}u^{2}+\int_{\mathbb{R}\setminus\{0\}}\left(\mathrm{e}^{u\xi_{1}}-1-u\frac{\xi_{1}}{\xi_{1}^{2}+1}\right)\nu(\mathrm{d}\xi_{1})

for all uu\in\mathbb{C} where the expectation exists. The independence and stationarity of the increments of XX imply that SS is a martingale if and only if h(1)=0h(1)=0, which is, in terms of the characteristic triplet (δ,σ2,ν)(\delta,\sigma^{2},\nu), equivalent to (1,)eξ1ν(dξ1)<\int_{(1,\infty)}\mathrm{e}^{\xi_{1}}\nu(\mathrm{d}\xi_{1})<\infty and

(32) δ=12σ2{0}(eξ11ξ1ξ12+1)ν(dξ1).\displaystyle\delta=-\frac{1}{2}\sigma^{2}-\int_{\mathbb{R}\setminus\{0\}}\left(\mathrm{e}^{\xi_{1}}-1-\frac{\xi_{1}}{\xi_{1}^{2}+1}\right)\nu(\mathrm{d}\xi_{1}).

The limiting cumulant generating function for the family of random variables (Xt/t)t1(X_{t}/t)_{t\geq 1}, defined by the limit in (18), is in the case when XX is a Lévy process given trivially by hh in (4.2), which therefore also coincides with definition (23). The martingale condition for S=eXS=\mathrm{e}^{X} and the convexity of hh imply that [0,1][0,1] is contained in the effective domain 𝒟h\mathcal{D}_{h}. In the case of affine stochastic volatility models we had to establish Theorem 10 to obtain sufficient condition for the set {0,1}\{0,1\} to be contained in the interior 𝒟h\mathcal{D}_{h}^{\circ} of the effective domain of hh. In the setting of Lévy processes it is well known (see e.g. [Sat99, Theorem 25.17]) that {0,1}𝒟h\{0,1\}\subset\mathcal{D}_{h}^{\circ} if and only if

(33) (,1)eu0ξ1ν(dξ1)+(1,)eu1ξ1ν(dξ1)<for someu0<0andu1>1.\displaystyle\int_{(-\infty,-1)}\mathrm{e}^{u_{0}\xi_{1}}\nu\left(\mathrm{d}\xi_{1}\right)+\int_{(1,\infty)}\mathrm{e}^{u_{1}\xi_{1}}\nu\left(\mathrm{d}\xi_{1}\right)<\infty\qquad\text{for some}\quad u_{0}<0\quad\text{and}\quad u_{1}>1.

Condition (33) implies that the interior of the effective domain of hh is of the form 𝒟h=(u,u+)\mathcal{D}_{h}^{\circ}=(u_{-},u_{+}) for some u[,0)u_{-}\in[-\infty,0) and u+(1,]u_{+}\in(1,\infty]. It is therefore clear that hh is steep if and only if

(34) (,1)|ξ1|eξ1uν(dξ1)=\displaystyle\int_{(-\infty,-1)}|\xi_{1}|\mathrm{e}^{\xi_{1}u_{-}}\nu\left(\mathrm{d}\xi_{1}\right)=\infty and (1,)ξ1eξ1u+ν(dξ1)=,\displaystyle\int_{(1,\infty)}\xi_{1}\mathrm{e}^{\xi_{1}u_{+}}\nu\left(\mathrm{d}\xi_{1}\right)=\infty,

where the integrals are taken to be infinite if the integrands take infinite value for some finite ξ1\xi_{1} (e.g. if u=u_{-}=-\infty or u+=u_{+}=\infty). Note also that under assumption (32), the Lévy process XX is non-constant if and only if there is a Brownian component (i.e. σ2>0\sigma^{2}>0) or its paths are discontinuous (i.e. ν0\nu\neq 0). Hence the equality

h′′(u)=σ2+{0}ξ12euξ1ν(dξ1),u𝒟h,h^{\prime\prime}(u)=\sigma^{2}+\int_{\mathbb{R}\setminus\{0\}}\xi_{1}^{2}\,\mathrm{e}^{u\xi_{1}}\,\nu(\mathrm{d}\xi_{1}),\qquad u\in\mathcal{D}_{h}^{\circ},

implies h′′(u)>0h^{\prime\prime}(u)>0 for all u𝒟hu\in\mathcal{D}_{h}^{\circ}. These arguments therefore imply Proposition 12, which is the analogue of Theorem 10 for Lévy processes.

Proposition 12.

Let XX be a non-constant Lévy process (i.e. the first component of a degenerate affine stochastic volatility process) with state-space \mathbb{R}, characteristic triplet (δ,σ2,ν)(\delta,\sigma^{2},\nu) and the characteristic exponent hh given by (4.2). Assume further that conditions (32), (33) and (34) are satisfied. Then the interior 𝒟h\mathcal{D}_{h}^{\circ} of the effective domain of hh is an interval (u,u+)(u_{-},u_{+}), where uu+u_{-}\leq u_{+} are extended real numbers, hh is a convex essentially smooth limiting cumulant generating function for the family (Xt/t)t1(X_{t}/t)_{t\geq 1} and the set {0,1}\{0,1\} is contained in the interior of 𝒟h\mathcal{D}_{h}. Furthermore, hh is smooth on 𝒟h\mathcal{D}_{h}^{\circ} and h′′(u)>0h^{\prime\prime}(u)>0 for all u𝒟hu\in\mathcal{D}_{h}^{\circ}.

5. Rate functions and the option prices far from maturity

In this section we describe the limiting behaviour of a family of European options under an affine stochastic volatility model S=eXS=\mathrm{e}^{X}. These results will be used in Section 6 to prove the formulae for the limiting implied volatility smile.

In order to understand the limits of the vanilla option prices far from maturity in an affine stochastic volatility model SS, we will need to apply the large deviation principle for the family (Xt/t)t1(X_{t}/t)_{t\geq 1} under a risk-neutral measure 𝖯\mathsf{P} and under the measure 𝖯~\widetilde{\mathsf{P}}, known as the share measure.111The name stems from the fact that under 𝖯~\widetilde{\mathsf{P}} the numeraire asset is the risky security S=eXS=\mathrm{e}^{X}. Recall that for every t0t\geq 0 the measure 𝖯~\widetilde{\mathsf{P}} is equivalent to 𝖯\mathsf{P} on the σ\sigma-field t\mathcal{F}_{t} and the Radon-Nikodym derivative is given by

d𝖯~d𝖯|t=eXt.\frac{\mathrm{d}\widetilde{\mathsf{P}}}{\mathrm{d}\mathsf{P}}\Big{\lvert}_{\mathcal{F}_{t}}\>=\>\mathrm{e}^{X_{t}}.

The limiting cumulant generating function for (Xt/t)t0(X_{t}/t)_{t\geq 0} under 𝖯~\widetilde{\mathsf{P}} is defined by

(35) h~(u)\displaystyle\widetilde{h}(u) :=\displaystyle:= limt1tlog𝖤~[euXt].\displaystyle\lim_{t\to\infty}\frac{1}{t}\log\widetilde{\mathsf{E}}\left[\mathrm{e}^{uX_{t}}\right].

The function h~\widetilde{h} and its effective domain 𝒟h~\mathcal{D}_{\widetilde{h}} satisfy

(36) h~(u)=h(u+1)and𝒟h~={u: 1+u𝒟h},\displaystyle\widetilde{h}(u)=h(u+1)\qquad\text{and}\qquad\mathcal{D}_{\widetilde{h}}=\{u\in\mathbb{R}\>:\>1+u\in\mathcal{D}_{h}\},

where hh is the limiting cumulant generating function defined in (23) and 𝒟h\mathcal{D}_{h} is its effective domain. Note that 0𝒟h~0\in\mathcal{D}_{\widetilde{h}}^{\circ} if and only if 1𝒟h1\in\mathcal{D}_{h}^{\circ}. The identity in (36) implies the following relationship between the Fenchel-Legendre transforms (see (19) for the definition) of hh and h~\widetilde{h}:

(37) h~(x)=h(x)xfor allx.\widetilde{h}^{*}(x)=h^{*}(x)-x\quad\text{for all}\quad x\in\mathbb{R}.

Theorem 13 below describes the limiting behaviour of certain European derivatives under an affine stochastic volatility process (X,V)(X,V). Before we state it, we collect the following facts.

Remarks.

(i) If (X,V)(X,V) is a non-degenerate affine stochastic volatility process that satisfies assumptions of Lemma 9, then the limiting cumulant generating functions hh and h~\widetilde{h} (defined in (23) and (35) respectively) are strictly convex with strictly positive second derivatives in the interior of their respective effective domains. Remark (B) after Definition 5 implies that their convex duals hh^{*} and h~\widetilde{h}^{*} are strictly convex and differentiable with respective unique global minima attained at

(38) x=h(0)\displaystyle x^{*}=h^{\prime}(0) and x~=h~(0)=h(1).\displaystyle\widetilde{x}^{*}=\widetilde{h}^{\prime}(0)=h^{\prime}(1).

Lemma 9 also implies the following inequalities:

(39) x\displaystyle x^{*} <0<\displaystyle<\quad 0\quad< x~.\displaystyle\widetilde{x}^{*}.

(ii) If (X,V)(X,V) is a degenerate affine stochastic volatility process that satisfies assumptions of Proposition 12, then hh^{\prime} is strictly increasing on 𝒟h\mathcal{D}_{h}^{\circ} and its image is equal to \mathbb{R}. The unique global minima of the Fenchel-Legendre transforms hh^{*} and h~\widetilde{h}^{*} are (by Remark (B) after Definition 5) explicitly given by

(40a) x\displaystyle x^{*} =h(0)=12σ2{0}(eξ11ξ1)ν(dξ1),\displaystyle=h^{\prime}(0)=-\frac{1}{2}\sigma^{2}-\int_{\mathbb{R}\setminus\{0\}}\left(\mathrm{e}^{\xi_{1}}-1-\xi_{1}\right)\,\nu(\mathrm{d}\xi_{1}),
(40b) x~\displaystyle\widetilde{x}^{*} =h(1)=12σ2+{0}(eξ1(ξ11)+1)ν(dξ1).\displaystyle=h^{\prime}(1)=\frac{1}{2}\sigma^{2}+\int_{\mathbb{R}\setminus\{0\}}\left(\mathrm{e}^{\xi_{1}}(\xi_{1}-1)+1\right)\,\nu(\mathrm{d}\xi_{1}).

Formulae (40) show that the inequalities in (39) hold also in the degenerate case.
(iii) In the case of the Black-Scholes model (i.e. ν=0\nu=0), the assumptions of Proposition 12 are satisfied. The effective domains of hBSh_{\mathrm{BS}} and h~BS\widetilde{h}_{\mathrm{BS}} are equal to \mathbb{R} and the following formulae hold

(41) hBS(u)=12σ2(u2u)\displaystyle h_{\mathrm{BS}}(u)=\frac{1}{2}\sigma^{2}(u^{2}-u) and h~BS(u)=12σ2(u2+u)foru,\displaystyle\widetilde{h}_{\mathrm{BS}}(u)=\frac{1}{2}\sigma^{2}(u^{2}+u)\quad\text{for}\quad u\in\mathbb{R},
(42) hBS(x;σ2)=12σ2(x+σ22)2\displaystyle h^{*}_{\mathrm{BS}}\left(x;\sigma^{2}\right)=\frac{1}{2\sigma^{2}}\left(x+\frac{\sigma^{2}}{2}\right)^{2} and h~BS(x;σ2)=12σ2(xσ22)2forx.\displaystyle\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma^{2}\right)=\frac{1}{2\sigma^{2}}\left(x-\frac{\sigma^{2}}{2}\right)^{2}\quad\text{for}\quad x\in\mathbb{R}.

Therefore we have x=σ2/2x^{*}=-\sigma^{2}/2 and x~=σ2/2\widetilde{x}^{*}=\sigma^{2}/2.

Theorem 13.

Let (X,V)(X,V) be a non-degenerate (resp. degenerate) affine stochastic volatility process that satisfies the assumptions of Theorem 10 or Corollaries 11 (i), 11 (ii) (resp. Proposition 12). Then the family of random variables (Xt/t)t1(X_{t}/t)_{t\geq 1} satisfies the LDPs under the measures 𝖯\mathsf{P} and 𝖯~\widetilde{\mathsf{P}} with the respective good rate functions hh^{*} and h~\widetilde{h}^{*}, where hh is given in (23) (resp. (4.2)) and h~\widetilde{h} in (35). Fix xx\in\mathbb{R}, let x,x~x^{*},\widetilde{x}^{*} be as in (38) (resp. (40)) and denote S=eXS=\mathrm{e}^{X} and y+:=max{0,y}y^{+}:=\max\{0,y\} for yy\in\mathbb{R}.

  1. (i)

    The asymptotic behaviour of a put option with strike S0extS_{0}\mathrm{e}^{xt} is given by the following formula

    limtt1log𝖤[(S0extSt)+]\displaystyle\lim_{t\to\infty}t^{-1}\log\mathsf{E}\left[\left(S_{0}\mathrm{e}^{xt}-S_{t}\right)^{+}\right] =\displaystyle= {xh(x)if xx,xif x>x.\displaystyle\left\{\begin{array}[]{ll}x-h^{*}\left(x\right)&\text{if }x\leq x^{*},\\ x&\text{if }x>x^{*}.\end{array}\right.
  2. (ii)

    The asymptotic behaviour of a call option, struck at S0etxS_{0}\mathrm{e}^{tx}, is given by the formula

    limtt1log𝖤[(StS0ext)+]\displaystyle\lim_{t\to\infty}t^{-1}\log\mathsf{E}\left[\left(S_{t}-S_{0}\mathrm{e}^{xt}\right)^{+}\right] =\displaystyle= {h~(x)if xx~,0if x<x~.\displaystyle\left\{\begin{array}[]{ll}-\widetilde{h}^{*}\left(x\right)&\text{if }x\geq\widetilde{x}^{*},\\ 0&\text{if }x<\widetilde{x}^{*}.\end{array}\right.
  3. (iii)

    The asymptotic behaviour of a covered call option with payoff St(StS0etx)+S_{t}-(S_{t}-S_{0}\mathrm{e}^{tx})^{+} is given by

    limtt1log(S0𝖤[(StS0ext)+])\displaystyle\lim_{t\to\infty}t^{-1}\log\left(S_{0}-\mathsf{E}\left[\left(S_{t}-S_{0}\mathrm{e}^{xt}\right)^{+}\right]\right) =\displaystyle= {0if x>x~,xh(x)if x[x,x~],xif x<x.\displaystyle\left\{\begin{array}[]{ll}0&\text{if }x>\widetilde{x}^{*},\\ x-h^{*}\left(x\right)&\text{if }x\in\left[x^{*},\widetilde{x}^{*}\right],\\ x&\text{if }x<x^{*}.\end{array}\right.

Furthermore the convergence in (i)-(iii) is uniform in xx on compact subsets of \mathbb{R}.

Remarks.

(I) The formulae in (i), (ii) and (iii) of Theorem 13 are continuous in xx since the value of the Fenchel-Legendre transforms hh^{*} (resp. h~\widetilde{h}^{*}) at xx^{*} (resp. x~\widetilde{x}^{*}) is equal to zero. Note further that the formulae in Theorem 13 are independent of the starting value (X0,V0)(X_{0},V_{0}) of the model.
(II) The reason for studying the limiting behaviour of the put, call and covered call in Theorem 13 lies in the fact that these payoffs yield non-trivial limits on complementary subintervals of \mathbb{R}, thus obtaining a non-trivial limit for every xx\in\mathbb{R}. This limit will be compared in Section 6 with the corresponding limit in the Black-Scholes model, which will yield the formula for the limiting implied volatility smile under affine stochastic volatility models.

Proof.

Assume first that (X,V)(X,V) is a non-degenerate affine stochastic volatility process. The limiting cumulant generating function hh satisfies (18) and is essentially smooth either by Theorem 10 or by Corollaries 11 (i), 11 (ii). Hence its convex dual hh^{*} is non-negative (by (19) and the fact h(0)=0h(0)=0), has compact level sets (by (20) and 0𝒟h0\in\mathcal{D}_{h}^{\circ}) and is differentiable on 𝒟h=\mathcal{D}_{h^{*}}=\mathbb{R} with strictly increasing first derivative (by Lemma 9 and Remark (B) following Definition 5). Therefore by Theorem 6 the family (Xt/t)t1(X_{t}/t)_{t\geq 1} satisfies the LDP under 𝖯\mathsf{P} with the good rate function hh^{*}. Since 1𝒟h1\in\mathcal{D}_{h}^{\circ}, by (36) the function h~\widetilde{h} satisfies the condition in (18). Therefore all of the assumptions of Theorem 6 hold under 𝖯~\widetilde{\mathsf{P}} and hence (Xt/t)t1(X_{t}/t)_{t\geq 1} satisfies the LDP with the good rate function h~\widetilde{h}^{*}. Furthermore h~\widetilde{h}^{*} enjoys the same regularity on 𝒟h~=\mathcal{D}_{\widetilde{h}^{*}}=\mathbb{R} as the rate function hh^{*}. The LDPs in the degenerate case follow from the same argument with Theorem 10, Corollaries 11 (i), 11 (ii) and Lemma 9 replaced by Proposition 12.

We now prove the formulae in Theorem 13. Without loss of generality we may assume that S0=1S_{0}=1, i.e. X0=0X_{0}=0. The following inequality holds for all t1t\geq 1 and ε>0\varepsilon>0:

etx(1eε)I{Xt/t<xε}(exteXt)+etxI{Xt/t<x}.\mathrm{e}^{tx}\left(1-\mathrm{e}^{-\varepsilon}\right)I_{\{X_{t}/t<x-\varepsilon\}}\leq\left(\mathrm{e}^{xt}-\mathrm{e}^{X_{t}}\right)^{+}\leq\mathrm{e}^{tx}I_{\{X_{t}/t<x\}}.

Hence by taking expectations, logarithms, multiplying by 1/t1/t and applying the LDP for (Xt/t)t1(X_{t}/t)_{t\geq 1} under 𝖯\mathsf{P} we obtain the following inequalities

xinfy<xεh(y)lim inft1tlog𝖤[(exteXt)+]lim supt1tlog𝖤[(exteXt)+]xinfyxh(y).x-\inf_{y<x-\varepsilon}h^{*}(y)\leq\liminf_{t\to\infty}\frac{1}{t}\log\mathsf{E}\left[\left(\mathrm{e}^{xt}-\mathrm{e}^{X_{t}}\right)^{+}\right]\leq\limsup_{t\to\infty}\frac{1}{t}\log\mathsf{E}\left[\left(\mathrm{e}^{xt}-\mathrm{e}^{X_{t}}\right)^{+}\right]\leq x-\inf_{y\leq x}h^{*}(y).

Since hh^{*} is continuous on \mathbb{R}, strictly decreasing for xxx\leq x^{*} and takes value 0 at xx^{*}, the formula in Theorem 13 (i) holds.

We now consider the call option case. The following inequality holds for all t1t\geq 1 and ε>0\varepsilon>0:

eXt(1eε)I{Xt/t>x+ε}(eXtext)+eXtI{Xt/t>x}.\mathrm{e}^{X_{t}}\left(1-\mathrm{e}^{-\varepsilon}\right)I_{\{X_{t}/t>x+\varepsilon\}}\leq\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\leq\mathrm{e}^{X_{t}}I_{\{X_{t}/t>x\}}.

Again by taking expectations, changing measure to 𝖯~\widetilde{\mathsf{P}}, applying logarithms, multiplying by 1/t1/t and applying the LDP for (Xt/t)t1(X_{t}/t)_{t\geq 1} under 𝖯~\widetilde{\mathsf{P}} we obtain the following inequalities

infy>x+εh~(y)lim inft1tlog𝖤[(eXtext)+]lim supt1tlog𝖤[(eXtext)+]infyxh~(y).-\inf_{y>x+\varepsilon}\widetilde{h}^{*}(y)\leq\liminf_{t\to\infty}\frac{1}{t}\log\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\leq\limsup_{t\to\infty}\frac{1}{t}\log\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\leq-\inf_{y\geq x}\widetilde{h}^{*}(y).

Note that x~\widetilde{x}^{*} is a global minimum for h~\widetilde{h}^{*} at which value 0 is attained. The continuity of h~\widetilde{h}^{*} implies the formula in Theorem 13 (ii).

In the case of the covered call, the following simple inequalities hold for all xx\in\mathbb{R}:

(46) extI{Xt/tx}\displaystyle\mathrm{e}^{xt}I_{\{X_{t}/t\geq x\}} eXt(eXtetx)+=\displaystyle\leq\quad\mathrm{e}^{X_{t}}-\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{tx}\right)^{+}\quad= eXtI{Xt/t<x}+extI{Xt/tx},\displaystyle\mathrm{e}^{X_{t}}I_{\{X_{t}/t<x\}}+\mathrm{e}^{xt}I_{\{X_{t}/t\geq x\}},
(47) eXtI{Xt/tx}\displaystyle\mathrm{e}^{X_{t}}I_{\{X_{t}/t\leq x\}} eXt(eXtetx)+\displaystyle\leq\quad\mathrm{e}^{X_{t}}-\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{tx}\right)^{+}\quad\leq eXt,\displaystyle\mathrm{e}^{X_{t}},
(48) extI{Xt/tx}\displaystyle\mathrm{e}^{xt}I_{\{X_{t}/t\geq x\}} eXt(eXtetx)+\displaystyle\leq\quad\mathrm{e}^{X_{t}}-\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{tx}\right)^{+}\quad\leq ext.\displaystyle\mathrm{e}^{xt}.

Inequality (46) and the LDP under measures 𝖯\mathsf{P} and 𝖯~\widetilde{\mathsf{P}} imply the inequalities

ext𝖯[Xt/tx]\displaystyle\mathrm{e}^{xt}\mathsf{P}\left[X_{t}/t\geq x\right] \displaystyle\leq 1𝖤[(eXtext)+]=𝖯~[Xt/t<x]+ext𝖯[Xt/tx]\displaystyle 1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\quad=\quad\widetilde{\mathsf{P}}\left[X_{t}/t<x\right]+\mathrm{e}^{xt}\mathsf{P}\left[X_{t}/t\geq x\right]
\displaystyle\leq exp(tinfyxh~(y)+εt)+extexp(tinfyxh(y)+εt)\displaystyle\exp\left(-t\inf_{y\leq x}\widetilde{h}^{*}(y)+\varepsilon t\right)+\mathrm{e}^{xt}\exp\left(-t\inf_{y\geq x}h^{*}(y)+\varepsilon t\right)

for any xx\in\mathbb{R}, ε>0\varepsilon>0 and tt large enough. Assume now x[x,x~]x\in\left[x^{*},\widetilde{x}^{*}\right] and note that in this case we have infyxh(y)=h(x)\inf_{y\geq x}h^{*}(y)=h^{*}(x) and infyxh~(y)=h~(x)\inf_{y\leq x}\widetilde{h}^{*}(y)=\widetilde{h}^{*}(x). By (37) we obtain

x+1tlog𝖯[Xt/tx]1tlog(1𝖤[(eXtext)+])xh(x)+ε+1tlog2x+\frac{1}{t}\log\mathsf{P}\left[X_{t}/t\geq x\right]\leq\frac{1}{t}\log\left(1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\right)\leq x-h^{*}(x)+\varepsilon+\frac{1}{t}\log 2

for any ε\varepsilon and all large tt. Therefore we find the inequalities

xh(x)\displaystyle x-h^{*}(x) \displaystyle\leq lim inft1tlog(1𝖤[(eXtext)+])\displaystyle\liminf_{t\to\infty}\frac{1}{t}\log\left(1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\right)
\displaystyle\leq lim supt1tlog(1𝖤[(eXtext)+])xh(x)+ε\displaystyle\limsup_{t\to\infty}\frac{1}{t}\log\left(1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\right)\leq x-h^{*}(x)+\varepsilon

for all ε>0\varepsilon>0. This proves the formula in Theorem 13 (iii) for x[x,x~]x\in[x^{*},\widetilde{x}^{*}].

Assume that x>x~x>\widetilde{x}^{*} and take expectations, change measure to 𝖯~\widetilde{\mathsf{P}}, apply logarithms and multiply by 1/t1/t the inequalities in (47) to obtain the following:

1tlog𝖯~[Xt/tx]1tlog(1𝖤[(eXtext)+])0.\frac{1}{t}\log\widetilde{\mathsf{P}}\left[X_{t}/t\leq x\right]\leq\frac{1}{t}\log\left(1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\right)\leq 0.

Since infy<xh~(y)=0\inf_{y<x}\widetilde{h}^{*}(y)=0 the LDP for (Xt/t)t1(X_{t}/t)_{t\geq 1} under 𝖯~\widetilde{\mathsf{P}} implies the formula in Theorem 13 (iii) that corresponds to x>x~x>\widetilde{x}^{*}.

Finally let x<xx<x^{*}. Inequalities in (48) imply the following

x+1tlog𝖯[Xt/tx]1tlog(1𝖤[(eXtext)+])x.x+\frac{1}{t}\log\mathsf{P}\left[X_{t}/t\geq x\right]\leq\frac{1}{t}\log\left(1-\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]\right)\leq x.

An application of the LDP for (Xt/t)t1(X_{t}/t)_{t\geq 1} under 𝖯\mathsf{P} completes the proof of part (iii).

We now show that the limits in the theorem are uniform in xx on compact sets in \mathbb{R}. Since the argument is similar in all the cases, we concentrate on Theorem 13 (i). Let (x0,y0)(x_{0},y_{0}) be a finite interval in \mathbb{R} and define for any xx\in\mathbb{R} and t1t\geq 1

V(t,x)=t1log𝖤[(exteXt)+]v(x),V(t,x)=t^{-1}\log\mathsf{E}\left[\left(\mathrm{e}^{xt}-\mathrm{e}^{X_{t}}\right)^{+}\right]-v(x),

where v(x)v(x) denotes the continuous limit that appears in Theorem 13 (i). It follows that

V(t,x0)+v(x0)v(x)V(t,x)V(t,y0)+v(y0)v(x)for anyx(x0,y0).V(t,x_{0})+v(x_{0})-v(x)\leq V(t,x)\leq V(t,y_{0})+v(y_{0})-v(x)\qquad\text{for any}\quad x\in(x_{0},y_{0}).

We therefore find

(49) |V(t,x)|\displaystyle\lvert V(t,x)\rvert \displaystyle\leq max{|V(t,y0)|,|V(t,x0)|}+max{|v(x)v(x0)|,|v(x)v(y0)|}.\displaystyle\max\left\{\lvert V(t,y_{0})\rvert,\lvert V(t,x_{0})\rvert\right\}+\max\left\{\lvert v(x)-v(x_{0})\rvert,\lvert v(x)-v(y_{0})\rvert\right\}.

Since we have already proved that limt|V(t,x)|=0\lim_{t\to\infty}|V(t,x)|=0 for every xx and the limiting function v(x)v(x) is continuous, and hence uniformly continuous on every compact set, the inequality in (49) implies that the logarithms of the put option prices converge to v(x)v(x) uniformly in xx on compact sets in \mathbb{R}.

6. Asymptotic behaviour of the implied volatility

The value C(S0,K,t,σ2)C(S_{0},K,t,\sigma^{2}) of the European call option with strike KK and expiry tt in a Black-Scholes model (i.e. degenerate affine stochastic volatility model without jumps, see Section 4.2) is given by the Black-Scholes formula

(50) C(S0,K,t,σ2)\displaystyle C(S_{0},K,t,\sigma^{2}) =\displaystyle= S0N(d+)KN(d),whered±=log(S0/K)±σ2t/2σt\displaystyle S_{0}\,N(d_{+})-K\,N(d_{-}),\qquad\text{where}\quad d_{\pm}=\frac{\log(S_{0}/K)\pm\sigma^{2}t/2}{\sigma\sqrt{t}}

and N()N(\cdot) is the standard normal cumulative distribution function. Let S=eXS=\mathrm{e}^{X} be an affine stochastic volatility model from Definition 2 with the starting point S0=eX0S_{0}=\mathrm{e}^{X_{0}}. The implied volatility in the model S=eXS=\mathrm{e}^{X} for the strike K>0K>0 and maturity t>0t>0 is the unique positive number σ^(K,t)\widehat{\sigma}(K,t) that satisfies the following equation in the variable σ\sigma:

(51) C(S0,K,t,σ2)=𝖤[(StK)+].\displaystyle C\left(S_{0},K,t,\sigma^{2}\right)=\mathsf{E}\left[\left(S_{t}-K\right)^{+}\right].

Implied volatility is well-defined since the function σC(S0,K,t,σ2)\sigma\mapsto C\left(S_{0},K,t,\sigma^{2}\right) is strictly increasing for positive σ\sigma (i.e. vega of a call option Cσ(S0,K,t,σ2)=S0N(d+)t\frac{\partial C}{\partial\sigma}(S_{0},K,t,\sigma^{2})=S_{0}N^{\prime}(d_{+})\sqrt{t} is strictly positive) and the right-hand side of (51) lies in the image of the Black-Scholes formula by a no-arbitrage argument. Put-call parity, which holds since S=eXS=\mathrm{e}^{X} is a true martingale, implies the identity P(S0,K,t,σ^(K,t)2)=𝖤[(KSt)+]P\left(S_{0},K,t,\widehat{\sigma}(K,t)^{2}\right)=\mathsf{E}\left[\left(K-S_{t}\right)^{+}\right], where P(S0,K,t,σ2)P\left(S_{0},K,t,\sigma^{2}\right) denotes the price of the put option in the Black-Scholes model with volatility σ\sigma.

If the affine stochastic volatility process (X,V)(X,V) satisfies the assumptions of Theorem 13, then the implied volatility has the following limit

(52) limtσ^(K,t)=22h(0)=22h((h)1(0))\displaystyle\lim_{t\to\infty}\widehat{\sigma}(K,t)=2\sqrt{2h^{*}(0)}=2\sqrt{-2h\left((h^{\prime})^{-1}(0)\right)}

for any fixed strike K>0K>0, where hh^{*} is the rate function of the model (the second equality in (52) follows from (21) and (22)). Tehranchi [Teh09] proved that the first equality in (52) holds uniformly in KK on compact sets in 0\mathbb{R}_{\geqslant 0} for non-negative local martingales with cumulant generating functions that satisfy certain additional conditions. Note that the limit in (52) is independent of KK, which corresponds to the well-known flattening of the implied volatility smile at large maturities. The uniform limit (in KK) on compact subsets of 0\mathbb{R}_{\geqslant 0}, given in (52), also follows from Theorem 14 for affine stochastic volatility processes (both in non-degenerate and degenerate, i.e. Lévy, cases).

In order to obtain a non-trivial limit at infinity we define the implied volatility σt(x)\sigma_{t}(x) for the strike K=S0exp(tx)K=S_{0}\exp(tx), where xx\in\mathbb{R}, by

(53) σt(x)\displaystyle\sigma_{t}(x) =\displaystyle= σ^(S0exp(tx),t).\displaystyle\widehat{\sigma}\left(S_{0}\exp(tx),t\right).

We will show that if (X,V)(X,V) satisfies the assumptions of Theorem 13, then the limiting implied volatility takes the form

(54) σ(x)\displaystyle\sigma_{\infty}(x) =\displaystyle= 2[sgn(x~x)h~(x)+sgn(xx)h(x)]forx,\displaystyle\sqrt{2}\left[\operatorname{sgn}(\widetilde{x}^{*}-x)\sqrt{\widetilde{h}^{*}(x)}+\operatorname{sgn}(x-x^{*})\sqrt{h^{*}(x)}\right]\qquad\text{for}\quad x\in\mathbb{R},

where hh^{*} and h~\widetilde{h}^{*} are the Fenchel-Legendre transforms (see (19) for definition) of the limiting cumulant generating functions hh and h~\widetilde{h} of XX under 𝖯\mathsf{P} and 𝖯~\widetilde{\mathsf{P}} respectively and x=h(0)x^{*}=h^{\prime}(0), x~=h~(0)=h(1)\widetilde{x}^{*}=\widetilde{h}^{\prime}(0)=h^{\prime}(1). The function sgn(x)\operatorname{sgn}(x) is by definition equal to 11 if x0x\geq 0 and 1-1 otherwise.

Remarks.

(i) Under the assumptions of Theorem 13, the points xx^{*}, x~\widetilde{x}^{*} are the locations of the unique global minima of the good rate functions hh^{*} and h~\widetilde{h}^{*} respectively and by (39) satisfy x<0<x~x^{*}<0<\widetilde{x}^{*}. Note that h~(x)h(x)\widetilde{h}^{*}(x)\leq h^{*}(x) for x0x\geq 0 and h~(x)h(x)\widetilde{h}^{*}(x)\geq h^{*}(x) for x0x\leq 0 and hence the following strict inequality σ(x)>0\sigma_{\infty}(x)>0 holds for all xx\in\mathbb{R}.
(ii) The function σ:(0,)\sigma_{\infty}:\mathbb{R}\to(0,\infty) in (54) is chosen so that it satisfies the following identities:

(55) hBS(x;σ(x)2)=h(x)andh~BS(x;σ(x)2)=h~(x),x,\displaystyle h^{*}_{\mathrm{BS}}\left(x;\sigma_{\infty}(x)^{2}\right)=h^{*}(x)\qquad\text{and}\qquad\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma_{\infty}(x)^{2}\right)=\widetilde{h}^{*}(x),\qquad x\in\mathbb{R},

where the polynomials hBS(x;σ2)h^{*}_{\mathrm{BS}}\left(x;\sigma^{2}\right) and h~BS(x;σ2)\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma^{2}\right) are given in (42). Quantities of importance in the proof of Theorem 14 will be the following partial derivatives

(56) hBSσ2(x;σ2)=h~BSσ2(x;σ2)=18σ4(σ2+2x)(σ22x).\displaystyle\frac{\partial h^{*}_{\mathrm{BS}}}{\partial\sigma^{2}}\left(x;\sigma^{2}\right)=\frac{\partial\widetilde{h}^{*}_{\mathrm{BS}}}{\partial\sigma^{2}}\left(x;\sigma^{2}\right)=\frac{1}{8\sigma^{4}}\left(\sigma^{2}+2x\right)\left(\sigma^{2}-2x\right).

(iii) In formula (54), the function σ(x)\sigma_{\infty}(x) is given as a linear combination of h(x)\sqrt{h^{*}(x)} and h~(x)\sqrt{\widetilde{h}^{*}(x)}. The coefficients in this linear combination are not uniquely determined by identities (55) (there are four possibilities). However definition (54) is the only choice that implies the following important properties

(57) σ(x)2\displaystyle\sigma_{\infty}(x)^{2} <\displaystyle< 2|x|,forx[x,x~]andσ(x)2=2x,σ(x~)2=2x~,\displaystyle 2|x|,\qquad\text{for}\quad x\in\mathbb{R}\setminus[x^{*},\widetilde{x}^{*}]\quad\text{and}\quad\sigma_{\infty}(x^{*})^{2}=2x^{*},\quad\sigma_{\infty}(\widetilde{x}^{*})^{2}=2\widetilde{x}^{*},
(58) σ(x)2\displaystyle\sigma_{\infty}(x)^{2} >\displaystyle> 2|x|,forx(x,x~),\displaystyle 2|x|,\qquad\text{for}\quad x\in(x^{*},\widetilde{x}^{*}),

which will be crucial in the proof of Theorem 14. Note that (57) and (58) trivially hold in the Black-Scholes model: x=σ2/2x^{*}=-\sigma^{2}/2, x~=σ2/2\widetilde{x}^{*}=\sigma^{2}/2 and the limiting smile is constant and equal to σ\sigma. The inequality in (57) on the interval (,x)(-\infty,x^{*}) follows from the identity σ(x)2+2x=4[h(x)h(x)2xh(x)]\sigma_{\infty}(x)^{2}+2x=4\left[h^{*}(x)-\sqrt{h^{*}(x)^{2}-xh^{*}(x)}\right] and the fact that for x<xx<x^{*} we have h(x)>0h^{*}(x)>0. Likewise the identity σ(x)2+2x=4[h(x)+h(x)2xh(x)]\sigma_{\infty}(x)^{2}+2x=4\left[h^{*}(x)+\sqrt{h^{*}(x)^{2}-xh^{*}(x)}\right] for x(x,0)x\in(x^{*},0) yields half of (58). The other half of (57) and (58) follow from analogous identities involving h~\widetilde{h}^{*}.
(iv) The arbitrary choice for sgn(0)=1\operatorname{sgn}(0)=1 in (54) is of no consequence since h(x)=h~(x~)=0h^{*}(x^{*})=\widetilde{h}^{*}(\widetilde{x}^{*})=0.

Theorem 14.

Let (X,V)(X,V) be an affine stochastic volatility process that satisfies the assumptions of Theorem 13. Then we have

limtσt(x)=σ(x)for anyx,\lim_{t\nearrow\infty}\sigma_{t}(x)=\sigma_{\infty}(x)\quad\text{for any}\quad x\in\mathbb{R},

where σt(x)\sigma_{t}(x), defined in (53), is the implied volatility in the model S=eXS=\mathrm{e}^{X} for the strike K=S0extK=S_{0}\mathrm{e}^{xt}, and σ(x)\sigma_{\infty}(x) is given in (54). Furthermore for any compact subset CC in {x,x~}\mathbb{R}\setminus\{x^{*},\widetilde{x}^{*}\}, where x,x~x^{*},\widetilde{x}^{*} are defined in (38), we have

supxC|σt(x)σ(x)|0ast.\sup_{x\in C}\>\Big{\lvert}\sigma_{t}(x)-\sigma_{\infty}(x)\Big{\rvert}\longrightarrow 0\quad\text{as}\quad t\nearrow\infty.
Remark.

Theorem 14 implies formula (52) obtained in [Teh09]: define x=log(K/S0)/tx=\log(K/S_{0})/t and apply the uniform convergence of σt(x)\sigma_{t}(x) on a compact neighbourhood of zero.

Proof.

We can assume without loss of generality that S0=1S_{0}=1. Assume further that x0>x~x_{0}>\widetilde{x}^{*} and pick ε>0\varepsilon>0. The goal is to find δ>0\delta>0 such that the following inequality holds for all large tt:

(59) |σt(x)2σ(x)2|<ε,wherex(x0δ,x0+δ).\displaystyle\lvert\sigma_{t}(x)^{2}-\sigma_{\infty}(x)^{2}\rvert<\varepsilon,\quad\text{where}\quad x\in(x_{0}-\delta,x_{0}+\delta).

Inequality (57) implies that there exists ε(0,ε)\varepsilon^{\prime}\in(0,\varepsilon) such that σ(x0)2+ε<2x0\sigma_{\infty}(x_{0})^{2}+\varepsilon^{\prime}<2x_{0} and 0<σ(x0)2ε0<\sigma_{\infty}(x_{0})^{2}-\varepsilon^{\prime}. By (56) we conclude that σ2h~BS(x;σ2)\sigma^{2}\mapsto\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma^{2}\right) is strictly decreasing on the interval (0,2x)(0,2x) and hence obtain the following inequalities:

h~BS(x0;σ(x0)2ε)>h~BS(x0;σ(x0)2)>h~BS(x0;σ(x0)2+ε).\widetilde{h}^{*}_{\mathrm{BS}}\left(x_{0};\sigma_{\infty}(x_{0})^{2}-\varepsilon^{\prime}\right)\>>\>\widetilde{h}^{*}_{\mathrm{BS}}\left(x_{0};\sigma_{\infty}(x_{0})^{2}\right)\>>\>\widetilde{h}^{*}_{\mathrm{BS}}\left(x_{0};\sigma_{\infty}(x_{0})^{2}+\varepsilon^{\prime}\right).

Since all the functions are continuous and identitiy (55) holds, there exists a δ>0\delta>0 such that x0δ>x~x_{0}-\delta>\widetilde{x}^{*} and the strict inequalities hold

(60) h~BS(x;σ(x)2ε)<h~(x)<h~BS(x;σ(x)2+ε)\displaystyle-\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma_{\infty}(x)^{2}-\varepsilon^{\prime}\right)\><\>-\widetilde{h}^{*}(x)\><\>-\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma_{\infty}(x)^{2}+\varepsilon^{\prime}\right)

for all x(x0δ,x0+δ)x\in(x_{0}-\delta,x_{0}+\delta). Theorem 13 implies that the call option converges uniformly on the interval (x0δ,x0+δ)(x_{0}-\delta,x_{0}+\delta) to h~(x)=limtt1log𝖤[(eXtext)+]-\widetilde{h}^{*}(x)=\lim_{t\to\infty}t^{-1}\log\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]. In particular in the Black-Scholes model we get h~BS(x;σ(x)2±ε)=limtt1logC(1,ext,t,σ(x)2±ε)-\widetilde{h}^{*}_{\mathrm{BS}}\left(x;\sigma_{\infty}(x)^{2}\pm\varepsilon^{\prime}\right)=\lim_{t\to\infty}t^{-1}\log C(1,\mathrm{e}^{xt},t,\sigma_{\infty}(x)^{2}\pm\varepsilon^{\prime}) and the convergence is uniform in xx on (x0δ,x0+δ)(x_{0}-\delta,x_{0}+\delta). Since σt(x)\sigma_{t}(x) satisfies 𝖤[(eXtext)+]=C(1,ext,t,σt(x)2)\mathsf{E}\left[\left(\mathrm{e}^{X_{t}}-\mathrm{e}^{xt}\right)^{+}\right]=C(1,\mathrm{e}^{xt},t,\sigma_{t}(x)^{2}) by definition, the inequalities in (60) imply that

C(1,ext,t,σ(x)2ε)<C(1,ext,t,σt(x)2)<C(1,ext,t,σ(x)2+ε)C(1,\mathrm{e}^{xt},t,\sigma_{\infty}(x)^{2}-\varepsilon^{\prime})\><\>C(1,\mathrm{e}^{xt},t,\sigma_{t}(x)^{2})\><\>C(1,\mathrm{e}^{xt},t,\sigma_{\infty}(x)^{2}+\varepsilon^{\prime})

for all x(x0δ,x0+δ)x\in(x_{0}-\delta,x_{0}+\delta) and all large tt. Since the Black-Scholes formula is strictly increasing in σ2\sigma^{2} (i.e. vega is strictly positive), these inequalities imply (59). This proves uniform convergence on any compact subset CC of (x~,)(\widetilde{x}^{*},\infty). The proof for a compact set C(,x~){x}C\subset(-\infty,\widetilde{x}^{*})\setminus\{x^{*}\} is analogous.

We now consider convergence at the point x~\widetilde{x}^{*}. Pick any ε>0\varepsilon>0 such that σ(x~)2=2x~>ε\sigma_{\infty}(\widetilde{x}^{*})^{2}=2\widetilde{x}^{*}>\varepsilon. Identity (42) implies that

h~BS(x~;σ(x~)2ε)>h~BS(x~;σ(x~)2)=h~(x~)=0<h~BS(x~;σ(x~)2+ε).\widetilde{h}^{*}_{\mathrm{BS}}\left(\widetilde{x}^{*};\sigma_{\infty}(\widetilde{x}^{*})^{2}-\varepsilon\right)>\widetilde{h}^{*}_{\mathrm{BS}}\left(\widetilde{x}^{*};\sigma_{\infty}(\widetilde{x}^{*})^{2}\right)=\widetilde{h}^{*}(\widetilde{x}^{*})=0<\widetilde{h}^{*}_{\mathrm{BS}}\left(\widetilde{x}^{*};\sigma_{\infty}(\widetilde{x}^{*})^{2}+\varepsilon\right).

The first inequality and the argument above imply that σ(x~)2ε<σt(x~)2\sigma_{\infty}(\widetilde{x}^{*})^{2}-\varepsilon<\sigma_{t}(\widetilde{x}^{*})^{2} for all large tt. Since σ(x~)2+ε>2x~\sigma_{\infty}(\widetilde{x}^{*})^{2}+\varepsilon>2\widetilde{x}^{*}, the second inequality, Theorem 13 yields

1C(1,ex~t,t,σt(x~)2)> 1C(1,ex~t,t,σ(x~)2+ε)1-C(1,\mathrm{e}^{\widetilde{x}^{*}t},t,\sigma_{t}(\widetilde{x}^{*})^{2})\>>\>1-C(1,\mathrm{e}^{\widetilde{x}^{*}t},t,\sigma_{\infty}(\widetilde{x}^{*})^{2}+\varepsilon)

for all large tt. This implies σt(x~)2<σ(x~)2+ε\sigma_{t}(\widetilde{x}^{*})^{2}<\sigma_{\infty}(\widetilde{x}^{*})^{2}+\varepsilon and hence proves the theorem for x~\widetilde{x}^{*}. The point xx^{*} can be dealt with analogously. ∎

The following corollary is a simple consequence of our results.

Corollary 15.

Let (X,V)(X,V) be a non-degenerate affine stochastic volatility process that satisfies the assumptions of Theorem 13. Then there exists a Lévy process YY, which satisfies assumptions of Theorem 14 as a degenerate affine stochastic volatility process, such that the limiting smiles of the models eX\mathrm{e}^{X} and eY\mathrm{e}^{Y} are identical.

Proof.

Let hh be the limiting cumulant generating function for (X,V)(X,V). Theorem 8 implies that hh is a cumulant generating function of an infinitely divisible random variable. By Theorem 10, the characteristic triplet of hh satisfies conditions (33) and (34). Therefore, if we define a Lévy process YY with this characteristic triplet, Theorem 14 and formula (54) imply that models XX and YY have identical limiting volatility smiles. ∎

Remarks.

(i) In other words Corollary 15 states that in the limit, non-degenerate affine stochastic volatility models cannot generate the behaviour of implied volatility, which is different from that generated by the processes with constant volatility and stationary, independent increments.
(ii) Corollary 15 suggests the following natural open question: can any limiting smile of an exponential Lévy model be obtained as a limit of implied volatility smiles of a non-degenerate affine stochastic volatility process? It is not immediately clear how to approach this problem because the characterisation of the limiting cumulant generating function hh of a model (X,V)(X,V), given in Theorems 8 and 10, does not give an explicit form of Lévy-Khintchine triplet of hh.

6.1. Examples of limiting smiles

We now apply the analysis to the examples of affine stochastic volatility models described in Section 2.2. In each of the cases the limiting cumulant generating function hh is available in closed form. If the assumptions of Theorem 10 or Corollaries 11 (i), 11 (ii) are satisfied, then the convex dual hh^{*} is a good rate function and hence the formula in (54) defines the limiting smile as maturity tends to infinity.

6.1.1. Heston model

The characteristics F,RF,R are given in (11) and χ(u)=uζρλ\chi(u)=u\zeta\rho-\lambda (see (12)). Assumption A5 is satisfied if and only if χ(1)<0\chi(1)<0, which is equivalent to λ>ζρ\lambda>\zeta\rho. Since λθ0\lambda\theta\neq 0 it follows that wF(0,w)w\mapsto F(0,w) is not identically 0. Since the assumptions of Corollary 11 (ii) are satisfied, Theorem 14 implies that the limiting smile is given by the formula in (54), where

(61) h(u)\displaystyle h(u) =\displaystyle= λθζ2(χ(u)+Δ(u))andΔ(u)=χ(u)2ζ2(u2u).\displaystyle-\frac{\lambda\theta}{\zeta^{2}}\left(\chi(u)+\sqrt{\Delta(u)}\right)\qquad\text{and}\quad\Delta(u)=\chi(u)^{2}-\zeta^{2}(u^{2}-u).

This implies the main result in [FJ11][FJM11]. A first order asymptotic expansion for the large maturity smile in the Heston model was obtained in [FJM10] using saddle point methods.

6.1.2. Heston model with state-independent jumps

The functions F,RF,R are given in (13) and χ(u)=uζρλ\chi(u)=u\zeta\rho-\lambda. As in Section 6.1.1, λ>ζρ\lambda>\zeta\rho implies that (X,V)(X,V) defined Section 2.2.2 is a non-degenerate affine stochastic volatility model that satisfies A5. As before assumption λθ0\lambda\theta\neq 0 implies that wF(0,w)w\mapsto F(0,w) is non-zero. κ~(u)\widetilde{\kappa}(u), defined in (14), is a cumulant generating function of the compensated pure-jump Lévy process JJ. Assume that there exists κ<0\kappa_{-}<0 such that |κ~(u)|<|\widetilde{\kappa}(u)|<\infty for u>κu>\kappa_{-}, |κ~(u)|=|\widetilde{\kappa}(u)|=\infty for u<κu<\kappa_{-} and (34) holds for u=κu_{-}=\kappa_{-} and u+=u_{+}=\infty (e.g. if the distribution of the absolute jump heights is exponential with parameter α>0\alpha>0, then κ=α\kappa_{-}=-\alpha). Under these assumptions on state-independent jumps, the function FF in (13a) is steep and {(0,0),(1,0)}𝒟F\{(0,0),(1,0)\}\subset\mathcal{D}_{F}^{\circ}. Hence Theorem 10 implies that the limiting cumulant generating function is of the form

h(u)=λθζ2(χ(u)+Δ(u))+κ~(u),h(u)=-\frac{\lambda\theta}{\zeta^{2}}\left(\chi(u)+\sqrt{\Delta(u)}\right)+\widetilde{\kappa}(u),

where Δ(u)\Delta(u) is as in (61), and Theorem 14 yields the limiting smile formula in (54). Note also that condition (34) on the jump measure is not necessary if Δ(u)<0\Delta(u)<0 for some u>κu>\kappa_{-}, since in this case FF in (13a) is automatically steep.

6.1.3. A model of Bates with state-dependent jumps

The functions F,RF,R are given in (15). Again we assume λ>ζρ\lambda>\zeta\rho and λθ0\lambda\theta\neq 0, which implies that (X,V)(X,V) defined in Section 2.2.3 is a non-degenerate affine stochastic volatility model that satisfies A5. It is clear from (15a) that the assumptions of Theorem 10 on FF are satisfied. Let κ~(u)\widetilde{\kappa}(u) be as in (15b) and assume that either |κ~(u)|<|\widetilde{\kappa}(u)|<\infty for all uu\in\mathbb{R} or there exists κ\kappa_{-}\in\mathbb{R} such that |κ~(u)|<|\widetilde{\kappa}(u)|<\infty for u>κu>\kappa_{-}, |κ~(u)|=|\widetilde{\kappa}(u)|=\infty for u<κu<\kappa_{-} and limn|κ~(un)|=\lim_{n\to\infty}|\widetilde{\kappa}(u_{n})|=\infty for any sequence (un)n(u_{n})_{n\in\mathbb{N}} with unκu_{n}\downarrow\kappa_{-}. Then 𝒟R\mathcal{D}_{R} is open in 2\mathbb{R}^{2} (see (15b)) and Theorem 10 implies that the limiting cumulant generating function takes the form

h(u)=λθζ2(χ(u)+Δ(u)),whereΔ(u)=χ(u)2ζ2(u2u+2κ~(u)).h(u)=-\frac{\lambda\theta}{\zeta^{2}}\left(\chi(u)+\sqrt{\Delta(u)}\right),\qquad\text{where}\quad\Delta(u)=\chi(u)^{2}-\zeta^{2}(u^{2}-u+2\widetilde{\kappa}(u)).

6.1.4. The Barndorff-Nielsen-Shephard model

The functions F,RF,R are given in (17). Since χ(u)=λ<0\chi(u)=-\lambda<0 and the jump measure is non-trivial (i.e. ν0\nu\neq 0), the process (X,V)(X,V) defined in Section 2.2.4 is a non-degenerate affine stochastic volatility process. Assume that κ(u)\kappa(u) defined in (16) is either finite for all uu\in\mathbb{R} or there exists κ+>0\kappa_{+}>0 with |κ(u)|<|\kappa(u)|<\infty for u<κ+u<\kappa_{+}, |κ(u)|=|\kappa(u)|=\infty for u>κ+u>\kappa_{+} and and (34) holds for u=κu_{-}=\kappa_{-} and u+=u_{+}=\infty (e.g. if the distribution of the absolute jump heights is exponential with parameter α>0\alpha>0, then κ=α\kappa_{-}=-\alpha). Then FF (see (17a)) satisfies the assumptions of Theorem 10 and the limiting cumulant generating function is of the form

h(u)=λκ(u22λ+u(ρ12λ))uλκ(ρ).h(u)=\lambda\kappa\left(\frac{u^{2}}{2\lambda}+u\left(\rho-\frac{1}{2\lambda}\right)\right)-u\lambda\kappa(\rho).

6.2. How close are the formula σ(x)\sigma_{\infty}(x) and the implied volatility σt(x)\sigma_{t}(x) for large maturity?

In this section we plot the difference |σ(x)σt(x)||\sigma_{\infty}(x)-\sigma_{t}(x)| for t{10,15}t\in\{10,15\} and x[0.1,0.1]x\in[-0.1,0.1] for the models with jumps from Section 6.1 (see Figure 1). In the case tt equals 10 years the error is approximately 4545 basis points (bp) with the strike KK ranging from 30%30\% to 200%200\% of the spot. At the maturity of 15 years the error is approximately 2020 bp and KK ranges between 20%20\% and 400%400\% of the spot.

In the cases of Heston with state-independent jumps and Bates with state-dependent jumps we took the following diffusion parameters

λ=1.15,ζ=0.2,θ=0.04,ρ=0.4\lambda=1.15,\quad\zeta=0.2,\quad\theta=0.04,\quad\rho=-0.4

and the Lévy measure ν(dξ1)=αeαξ1I{ξ1<0}dξ1\nu\left(\mathrm{d}\xi_{1}\right)=\alpha\mathrm{e}^{\alpha\xi_{1}}I_{\left\{\xi_{1}<0\right\}}\mathrm{d}\xi_{1} with α=0.6\alpha=0.6. The compensated cumulant generating function κ~(u)\widetilde{\kappa}(u) (see (14) for the definition of κ~(u)\widetilde{\kappa}(u) in Section 2.2.2 and note that it takes the same form in Section 2.2.3) is in this case given by

κ~(u)=u(u1)(u+α)(α+1)for all u(α,).\widetilde{\kappa}(u)=\frac{u\left(u-1\right)}{\left(u+\alpha\right)\left(\alpha+1\right)}\qquad\text{for all }u\in\left(-\alpha,\infty\right).

In the case of the BNS model we took a pure-jump subordinator JJ with Lévy measure ν(dξ2)=abebξ2I{ξ2>0}dξ2\nu\left(\mathrm{d}\xi_{2}\right)=ab\mathrm{e}^{-b\xi_{2}}I_{\left\{\xi_{2}>0\right\}}\mathrm{d}\xi_{2}. The cumulant generating function (16) is given by κ(u)=au/(bu)\kappa(u)=au/(b-u) for u<bu<b. We used the following values for the parameters

a=1.4338,b=11.6641,λ=0.5783,ρ=1.2606,a=1.4338,\quad b=11.6641,\quad\lambda=0.5783,\quad\rho=-1.2606,

which were taken from [Sch03, Section 7.3] where the model was calibrated to the options on the S&P 500.

Refer to caption
Refer to caption
Figure 1. This figure contains the plots of the function x|σ(x)σt(x)|x\mapsto|\sigma_{\infty}(x)-\sigma_{t}(x)| in the interval x[0.1,0.1]x\in[-0.1,0.1] for the models with jumps from Section 6.1 and maturities t{10,15}t\in\{10,15\}. The values of the model parameters used are given in Section 6.2.

References

  • [Bat00] D.S. Bates. Post-’87 crash fears in the S&P 500 futures option market. Journal of Econometrics, 94:181–238, 2000.
  • [BNS01] O.E. Barndorff-Nielsen and N. Shephard. Non-Gaussian Ornstein–Uhlenbeck based models and some of their uses in financial economics. Journal of the Royal Statistical Society B, 63:167–241, 2001.
  • [DFS03] D. Duffie, D. Filipovic, and W. Schachermayer. Affine processes and applications in finance. Annals of Applied Probability, 13(3):984–1053, 2003.
  • [DL06] D.A. Dawson and Z. Li. Skew convolution semigroups and affine Markov processes. Annals of Probability, 34(3):1103–1142, 2006.
  • [DZ98] A. Dembo and O. Zeitouni. Large deviations techniques and applications. Springer, Berlin, 2 edition, 1998.
  • [FJ11] M. Forde and A. Jacquier. The large-maturity smile for the Heston model. Forthcoming in Finance & Stochastics, 2011.
  • [FJM10] M. Forde, A. Jacquier, and A. Mijatović. Asymptotic formulae for implied volatility in the Heston model. Proceedings of the Royal Society A, 466(2124):3593–3620, 2010.
  • [FJM11] M. Forde, A. Jacquier, and A. Mijatović. A note on essential smoothness in the Heston model. Forthcoming in Finance & Stochastics, 2011.
  • [GL11] K. Gao and R. Lee. Asymptotics of implied volatility to arbitrary order. Preprint available at http://ssrn.com/abstract=1768383, 2011.
  • [Hes93] S.L. Heston. A closed-form solution for options with stochastic volatility with applications to bond and currency options. Review of Financial Studies, 6(2):327–343, 1993.
  • [KR11] M. Keller-Ressel. Moment explosions and long-term behavior of affine stochastic volatility models. Mathematical Finance, 1(21):73–98, 2011.
  • [Lew00] A. Lewis. Option valuation under stochastic volatility. Finance Press, 2000.
  • [Roc70] A. T. Rockafellar. Convex Analysis. Princeton University Press, New Jersey, 1970.
  • [Sat99] K.I. Sato. Lévy Processes and Infinitely Divisible Distributions. Cambridge University Press, 1999.
  • [Sch03] W. Schoutens. Lévy Processes in Finance: Pricing Financial Derivatives. Wiley, 2003.
  • [Teh09] M.R. Tehranchi. Asymptotics of implied volatility far from maturity. Journal of Applied Probability, 46(3):629–650, 2009.