1 Introduction
One of the well-established benchmarks for evaluating option pricing models is comparing the model-generated Black–Scholes implied volatility surface $(\tau ,\kappa )\hspace{-0.1667em}\mapsto \hspace{-0.1667em}\widehat{\sigma }(\tau ,\kappa )$ with the empirically observed one $(\tau ,\kappa )\mapsto {\widehat{\sigma }_{\text{emp}}}(\tau ,\kappa )$. In this context, τ represents the time to maturity and $\kappa :=\log \frac{K}{{e^{r\tau }}{S_{0}}}$ is the log-moneyness with K denoting the strike, ${S_{0}}$ the current price of an underlying asset and r being the instantaneous interest rate. In particular, for any fixed τ, the values of ${\widehat{\sigma }_{\text{emp}}}(\tau ,\kappa )$ plotted against κ are known to produce convex “smiley” patterns with negative slopes at-the-money (i.e. when $\kappa \approx 0$). Furthermore, as reported in, e.g., [8, 16, 20] or [12, Subsection 2.2], the smile at-the-money becomes progressively steeper as $\tau \to 0$ with a rule-of-thumb behavior
The phenomenon (1) is known as the power law of the at-the-money implied volatility skew, and if one wants to replicate it, one may look for a model with
However, it turns out that the property (2) is not easy to obtain: for example, as discussed in [1, Section 7.1] or [23, Remark 11.3.21], classical Brownian diffusion stochastic volatility models fail to produce implied volatilities with power law (2). In the literature, (2) is usually replicated by introducing a volatility process with a very low Hölder regularity within the rough volatility framework popularized by Gatheral, Jaisson and Rosenbaum in their landmark paper [20]. The efficiency of this approach can be explained as follows.
(1)
\[ \bigg|\frac{{\widehat{\sigma }_{\text{emp}}}(\tau ,\kappa )-{\widehat{\sigma }_{\text{emp}}}(\tau ,{\kappa ^{\prime }})}{\kappa -{\kappa ^{\prime }}}\bigg|\propto {\tau ^{-\frac{1}{2}+H}},\hspace{1em}\kappa ,{\kappa ^{\prime }}\approx 0,\hspace{2.5pt}H\in \bigg(0,\frac{1}{2}\bigg).\](2)
\[ {\bigg|\frac{\partial \widehat{\sigma }}{\partial \kappa }(\tau ,\kappa )\bigg|_{\kappa =0}}=O\big({\tau ^{-\frac{1}{2}+H}}\big),\hspace{1em}\tau \to 0.\]-
• On the one hand, a theoretical result of Fukasawa [17] suggests that the volatility process cannot be Hölder continuous of a high order in continuous nonarbitrage models exhibiting the property (2). In other words, the roughness of volatility is, in some sense, a necessary condition to reproduce (2) (at least in the fully continuous setting).
-
• On the other hand, as proved in the seminal 2007 paper [1] by Alòs, León and Vives, the short-term explosion (2) of the implied volatility skew can be deduced from the explosion of the Malliavin derivative of volatility. In particular, the latter characteristic is exhibited by fractional Brownian motion with $H\lt 1/2$, a common driver in the rough volatility literature.
However, despite the ability to reproduce the power law (2), rough volatility models are not perfect. In particular,
For more details on rough volatility, we refer the reader to the recent review [12, Subsection 3.3.2] or the regularly updated literature list on the subject [28].
-
– furthermore, there is no guaranteed procedure of transition between physical and pricing measures: it is not always clear whether the volatility process $\sigma =\{\sigma (t),\hspace{2.5pt}t\in [0,T]\}$ hits zero and therefore the integral ${\textstyle\int _{0}^{t}}\frac{1}{{\sigma ^{2}}(s)}ds$ that is typically present in martingale densities (see, e.g., [5]) may be poorly defined;
-
– just like many classical Brownian stochastic volatility models (see, e.g., [2]), they may suffer from moment explosions in price, which results in complications with the pricing of some assets, quadratic hedging, and numerical methods.
Recently, a series of papers [9–11] introduced the Sandwiched Volterra Volatility (SVV) model which accounts for all the problems mentioned above. More precisely, the volatility process $Y=\{Y(t),\hspace{2.5pt}t\in [0,T]\}$ is assumed to follow the stochastic differential equation
driven by a general Hölder continuous Gaussian Volterra process
The special part of the equation above is the drift b. It is assumed that there are two continuous functions $0\lt \varphi \lt \psi $ such that for some $\varepsilon \gt 0$
\[\begin{array}{r@{\hskip0pt}l@{\hskip0pt}r}\displaystyle b(t,y)& \displaystyle \ge \frac{C}{{(y-\varphi (t))^{\gamma }}},\hspace{2em}& \displaystyle y\in \big(\varphi (t),\varphi (t)+\varepsilon \big),\\ {} \displaystyle b(t,y)& \displaystyle \le -\frac{C}{{(\psi (t)-y)^{\gamma }}},\hspace{2em}& \displaystyle y\in \big(\psi (t)-\varepsilon ,\psi (t)\big).\end{array}\]
Such an explosive nature of the drift resembling the one in SDEs for Bessel processes (see, e.g., [27, Chapter XI]) or singular SDEs of [21] ensures that, with probability 1,
which immediately solves the moment explosion problem (see, e.g., [9, Theorem 2.6]) and allows for a transparent transition between physical and pricing measures [9, Subsection 2.2]. In addition, the flexibility in the choice of the kernel $\mathcal{K}$ should allow to replicate both the long memory and the power law behavior (2).The main goal of this paper is to give the theoretical justification to the latter claim: we prove that, with the correct choice of the Volterra kernel $\mathcal{K}$, the SVV model indeed reproduces (2). In order to do that, we employ the fundamental result [1, Theorem 6.3] by Alòs, León and Vives mentioned above and check that the Malliavin derivative $DY(t)$ indeed exhibits explosive behavior. The difficulty of this approach is as follows. While the first-order Malliavin differentiability of $Y(t)$ is established in [9, Section 3] with
However, justifying the computations in (3) is far from straightforward. For example, the functions $y\mapsto {b^{\prime }_{y}}(t,y)$ and $y\mapsto {b^{\prime\prime }_{yy}}(t,y)$ demonstrate explosive behavior as $y\to \varphi (t)+$ and $y\to \psi (t)-$ for any $t\in [0,T]$. This makes it impossible to use the classical Malliavin chain rules such as [25, Proposition 1.2.3] requiring boundedness of the derivative or [25, Proposition 1.2.4] demanding the Lipschitz condition. In order to overcome this issue, we have to use some special properties of the volatility process established in [11] and tailor a version of the Malliavin chain rule specifically for our needs.
\[ {D_{s}}Y(t)=\mathcal{K}(t,s)+{\int _{s}^{t}}\mathcal{K}(u,s){b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}du,\]
[1, Theorem 6.3] actually demands the existence of the second-order Malliavin derivative. In principle, it is intuitively clear how this derivative should look like:
(3)
\[ \begin{aligned}{}{D_{r}}{D_{s}}Y(t)& ={D_{r}}{\int _{s}^{t}}\mathcal{K}(u,s){b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}du\\ {} & ={\int _{s}^{t}}\mathcal{K}(u,s){D_{r}}\Bigg[{b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\Bigg]du\\ {} & ={\int _{s}^{t}}\mathcal{K}(u,s)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{D_{r}}\big[{b^{\prime }_{y}}\big(u,Y(u)\big)\big]du\\ {} & \hspace{1em}+{\int _{s}^{t}}\mathcal{K}(u,s){b^{\prime }_{y}}\big(u,Y(u)\big){D_{r}}\Bigg[\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\Bigg]du\\ {} & ={\int _{s}^{t}}\mathcal{K}(u,s){b^{\prime\prime }_{yy}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{D_{r}}\big[Y(u)\big]du\\ {} & \hspace{1em}+{\int _{s}^{t}}\mathcal{K}(u,s){b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\times \\ {} & \hspace{2em}\hspace{2em}\times {\int _{u}^{t}}{b^{\prime\prime }_{yy}}\big(v,Y(v)\big){D_{r}}\big[Y(v)\big]dvdu.\end{aligned}\]The paper is organized as follows. In Section 2, we provide some necessary details about the sandwiched volatility process Y. In Section 3, we prove second-order Malliavin differentiability of $Y(t)$. Finally, in Section 4, we use [1, Theorem 6.3] to determine conditions on the kernel under which the SVV model reproduces (2). In Appendix A, we gather some necessary facts from Malliavin calculus, list some of the notation and, in addition, we prove a general Malliavin product rule to fit our purposes and that we were not able to find in the literature.
2 Preliminaries on sandwiched processes
In this section, we gather all the necessary details about the main object of our study: the class of sandwiched processes driven by Hölder-continuous Gaussian Volterra noises.
Fix some $T\in (0,\infty )$ and consider a kernel $\mathcal{K}:{[0,T]^{2}}\to \mathbb{R}$ satisfying the following assumptions.
Assumption 1.
Let $B=\{B(t),\hspace{2.5pt}t\in [0,T]\}$ be a standard Brownian motion. Assumption 1 allows to define a Gaussian Volterra process
and, moreover, Assumption 1(K2) together with [3, Theorem 1 and Corollary 4] implies that Z has a modification with Hölder continuous trajectories of any order $\lambda \in (0,H)$. In what follows, we always use this modification of Z: in other words, with probability 1, for any $\lambda \in (0,H)$ there exists a random variable $\Lambda =\Lambda (\lambda )\gt 0$ such that for all $0\le {t_{1}}\le {t_{2}}\le T$
Furthermore, as stated in [3, Theorem 1], the random variable Λ from (6) can be chosen such that
In what follows, we assume that (7) always holds.
(7)
\[ \mathbb{E}\big[{\Lambda ^{r}}\big]\lt \infty \hspace{1em}\text{for all}\hspace{2.5pt}r\in \mathbb{R}.\]Next, denote
Take $H\in (0,1)$ from Assumption 1(K2), consider two H-Hölder continuous functions φ, ψ: $[0,T]\to \mathbb{R}$ such that
and define a function b: $\mathcal{D}\to \mathbb{R}$ as
where the coefficients in (9) satisfy the following assumption.
(8)
\[ \begin{aligned}{}\mathcal{D}& :=\big\{(t,y)\in [0,T]\times \mathbb{R}\hspace{2.5pt}|\hspace{2.5pt}\varphi (t)\lt y\lt \psi (t)\big\},\\ {} \overline{\mathcal{D}}& :=\big\{(t,y)\in [0,T]\times \mathbb{R}\hspace{2.5pt}|\hspace{2.5pt}\varphi (t)\le y\le \psi (t)\big\}.\end{aligned}\](9)
\[ b(t,y):=\frac{{\theta _{1}}(t)}{{(y-\varphi (t))^{{\gamma _{1}}}}}-\frac{{\theta _{2}}(t)}{{(\psi (t)-y)^{{\gamma _{2}}}}}+a(t,y),\]Assumption 2.
The constants ${\gamma _{1}}$, ${\gamma _{2}}\gt 0$ and functions ${\theta _{1}}$, ${\theta _{2}}$, a are such that
-
(B1) ${\gamma _{1}}\gt \frac{1}{H}-1$, ${\gamma _{2}}\gt \frac{1}{H}-1$ with $H\in (0,1)$ being from Assumption 1(K2);
-
(B2) the functions ${\theta _{1}}$, ${\theta _{2}}$: $[0,T]\to \mathbb{R}$ are strictly positive and continuous;
-
(B3) the function a: $[0,T]\times \mathbb{R}\to \mathbb{R}$ is locally Lipschitz in y uniformly in t, i.e. for any $N\gt 0$ there exists a constant ${C_{N}}\gt 0$ that does not depend on t such that
-
(B4) a: $[0,T]\times \mathbb{R}\to \mathbb{R}$ is two times differentiable w.r.t. the spatial variable y with a, ${a^{\prime }_{y}}$, ${a^{\prime\prime }_{yy}}$ all being continuous on $[0,T]\times \mathbb{R}$.
Remark 2.
Note that ${b^{\prime }_{y}}$ is bounded from above on $\mathcal{D}$: indeed,
\[\begin{aligned}{}{b^{\prime }_{y}}(t,y)& =-\frac{{\gamma _{1}}{\theta _{1}}(t)}{{(y-\varphi (t))^{{\gamma _{1}}+1}}}-\frac{{\gamma _{2}}{\theta _{2}}(t)}{{(\psi (t)-y)^{{\gamma _{2}}+1}}}+{a^{\prime }_{y}}(t,y)\\ {} & \lt \underset{(t,y)\in \overline{\mathcal{D}}}{\max }{a^{\prime }_{y}}(t,y)\lt \infty .\end{aligned}\]
Finally, fix $\varphi (0)\lt {y_{0}}\lt \psi (0)$ and consider a stochastic differential equation of the form
By [11, Theorem 4.1], under Assumptions 1 and 2, the SDE (10) has a unique strong solution $Y=\{Y(t),\hspace{2.5pt}t\in [0,T]\}$. Moreover, with probability 1,
In what follows, we will need to analyze the behavior of the stochastic processes $|b(t,Y(t))|$, $|{b^{\prime }_{y}}(t,Y(t))|$ and $|{b^{\prime\prime }}(t,Y(t))|$, $t\in [0,T]$. In this regard, the property (11) alone is not sufficient: the process Y can, in principle, approach the bounds φ and ψ which results in an explosive growth of the processes mentioned above. Luckily, [11, Theorem 4.2] provides a refinement of (11) allowing for a more precise control of Y near φ and ψ. We give a slightly reformulated version of this result below.
Theorem 1.
Let Assumptions 1 and 2 hold and $\lambda \in (0,H)$, $\Lambda =\Lambda (\lambda )\gt 0$ be from (6). Then there exist deterministic constants ${C_{Y}}={C_{Y}}(\lambda )\gt 0$ and $\beta =\beta (\lambda )\gt 0$ such that
\[ \varphi (t)+\frac{{C_{Y}}}{{(1+\Lambda )^{\beta }}}\le Y(t)\le \psi (t)-\frac{{C_{Y}}}{{(1+\Lambda )^{\beta }}}\hspace{1em}\textit{for all}\hspace{2.5pt}t\in [0,T].\]
In particular, since Λ can be chosen to have moments of all orders, for all $r\ge 0$
We finalize this section by citing the first-order Malliavin differentiability result for the sandwiched process (10) proved in [9, Section 3].
Theorem 2.
Remark 4.
The result above actually holds for more general drifts than the one given in (9). The same is also, in principle, true for the results of the subsequent sections. Namely, it would be sufficient to assume that there exist deterministic constants $c\gt 0$, $r\gt 0$, $\gamma \gt \frac{1}{H}-1$ and $0\lt {y_{\ast }}\lt {\max _{t\in [0,T]}}|\psi (t)-\varphi (t)|$ such that
However, since (9) is the most natural choice satisfying these assumptions, we stick to this shape for notational convenience.
-
• b: $\mathcal{D}\to \mathbb{R}$ is continuous on $\mathcal{D}$ and has continuous partial derivatives ${b^{\prime }_{y}}$, ${b^{\prime\prime }_{yy}}$;
-
• for any $0\lt \varepsilon \lt \frac{1}{2}{\max _{t\in [0,T]}}|\psi (t)-\varphi (t)|$,
-
• b has an explosive growth to ∞ near φ and explosive decay to $-\infty $ near ψ of order $\gamma \gt \frac{1}{H}-1$, i.e.\[\begin{array}{r@{\hskip0pt}l@{\hskip0pt}r}\displaystyle b(t,y)& \displaystyle \ge \frac{c}{{(y-\varphi (t))^{\gamma }}},\hspace{2em}& \displaystyle y\in \big(\varphi (t),\varphi (t)+{y_{\ast }}\big),\\ {} \displaystyle b(t,y)& \displaystyle \le -\frac{c}{{(\psi (t)-y)^{\gamma }}},\hspace{2em}& \displaystyle y\in \big(\psi (t)-{y_{\ast }},\psi (t)\big);\end{array}\]
3 Second-order Malliavin differentiability
Let Assumptions 1 and 2 hold and $Y=\{Y(t),\hspace{2.5pt}t\in [0,T]\}$ be the sandwiched process defined by (10) with the drift (9).
Notation.
Here and in the sequel, C will denote any positive deterministic constant the exact value of which is not relevant. Note that C may change from line to line (or even within one line).
The main goal of this section is to establish the second-order Malliavin differentiability of the sandwiched process (10) and compute the corresponding derivative explicitly. As mentioned above, the main difficulty lies in controlling the behavior of $b(t,Y(t))$, ${b^{\prime }_{y}}(t,Y(t))$ and ${b^{\prime\prime }_{yy}}(t,Y(t))$ whenever $Y(t)$ approaces the bounds. Luckily, Theorem 1 gives all the necessary tools to do that as summarized in the following proposition.
Proposition 1.
Proof.
Fix $\lambda \in (0,H)$ and take the corresponding $\Lambda \gt 0$ from (6) and ${C_{Y}},\beta \gt 0$ being from Theorem 1. Then
\[\begin{aligned}{}\big|b\big(t,Y(t)\big)\big|& =\frac{|{\theta _{1}}(t)|}{{(Y(t)-\varphi (t))^{{\gamma _{1}}}}}+\frac{|{\theta _{2}}(t)|}{{(\psi (t)-Y(t))^{{\gamma _{2}}}}}+\big|a\big(t,Y(t)\big)\big|\\ {} & \le \frac{{\sup _{t\in [0,T]}}|{\theta _{1}}(t)|{(1+\Lambda )^{\beta {\gamma _{1}}}}}{{C_{Y}^{{\gamma _{1}}}}}\\ {} & \hspace{1em}+\frac{{\sup _{t\in [0,T]}}|{\theta _{2}}(t)|{(1+\Lambda )^{\beta {\gamma _{2}}}}}{{C_{Y}^{{\gamma _{2}}}}}\\ {} & \hspace{1em}+\underset{(t,y)\in \mathcal{D}}{\sup }\big|a(t,y)\big|\\ {} & :={\xi _{0}},\\ {} \big|{b^{\prime }_{y}}\big(t,Y(t)\big)\big|& =\frac{{\gamma _{1}}|{\theta _{1}}(t)|}{{(Y(t)-\varphi (t))^{{\gamma _{1}}+1}}}+\frac{{\gamma _{2}}|{\theta _{2}}(t)|}{{(\psi (t)-Y(t))^{{\gamma _{2}}+1}}}+\big|{a^{\prime }_{y}}\big(t,Y(t)\big)\big|\\ {} & \le \frac{{\gamma _{1}}{\sup _{t\in [0,T]}}|{\theta _{1}}(t)|{(1+\Lambda )^{\beta ({\gamma _{1}}+1)}}}{{C_{Y}^{{\gamma _{1}}+1}}}\\ {} & \hspace{1em}+\frac{{\gamma _{2}}{\sup _{t\in [0,T]}}|{\theta _{2}}(t)|{(1+\Lambda )^{\beta ({\gamma _{2}}+1)}}}{{C_{Y}^{{\gamma _{2}}+1}}}\\ {} & \hspace{1em}+\underset{(t,y)\in \mathcal{D}}{\sup }\big|{a^{\prime }_{y}}(t,y)\big|\\ {} & :={\xi _{1}},\\ {} \big|{b^{\prime\prime }_{yy}}\big(t,Y(t)\big)\big|& =\frac{{\gamma _{1}}({\gamma _{1}}+1)|{\theta _{1}}(t)|}{{(Y(t)-\varphi (t))^{{\gamma _{1}}+2}}}+\frac{{\gamma _{2}}({\gamma _{2}}+1)|{\theta _{2}}(t)|}{{(\psi (t)-Y(t))^{{\gamma _{2}}+2}}}+\big|{a^{\prime\prime }_{yy}}\big(t,Y(t)\big)\big|\\ {} & \le \frac{{\gamma _{1}}({\gamma _{1}}+1){\sup _{t\in [0,T]}}|{\theta _{1}}(t)|{(1+\Lambda )^{\beta ({\gamma _{1}}+2)}}}{{C_{Y}^{{\gamma _{1}}+2}}}\\ {} & \hspace{1em}+\frac{{\gamma _{2}}({\gamma _{2}}+1){\sup _{t\in [0,T]}}|{\theta _{2}}(t)|{(1+\Lambda )^{\beta ({\gamma _{2}}+2)}}}{{C_{Y}^{{\gamma _{2}}+2}}}\\ {} & \hspace{1em}+\underset{(t,y)\in \mathcal{D}}{\sup }\big|{a^{\prime\prime }_{yy}}(t,y)\big|\\ {} & :={\xi _{2}}.\end{aligned}\]
Note that ${\xi _{0}}$, ${\xi _{1}}$ and ${\xi _{2}}$ have moments of all orders by the properties of Λ, see (7), and hence, putting
we obtain the required result. □As noted in Theorem 2, $Y(t)\in {\mathbb{D}^{1,2}}$ for each $t\ge 0$. In fact, Proposition 1 together with the shape (12) of the derivative allows to establish a more general result.
Proof.
Note that, by (11), $\mathbb{E}[|Y(t){|^{p}}]\lt \infty $ for any $p\gt 1$, so, by Lemma 1 from the Appendix, it is sufficient to prove that
By Assumption 1 and Remark 1,
therefore
which ends the proof. □
\[ \mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\lt \infty \]
for any $p\gt 1$. Note that, by Remark 2,
where
and, by Proposition 1, there exists a random variable ξ having all moments such that
Hence
(13)
\[ \begin{aligned}{}\big|{D_{s}}& Y(t)\big|\\ {} & \le \big|\mathcal{K}(t,s)\big|+{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{b^{\prime }_{y}}\big(u,Y(u)\big)\big|\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}du\\ {} & \le \big|\mathcal{K}(t,s)\big|+\xi \exp \{cT\}{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|du.\end{aligned}\](14)
\[ \begin{aligned}{}\mathbb{E}& \Bigg[{\Bigg({\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le C{\Bigg({\int _{0}^{T}}{\mathcal{K}^{2}}(t,s)ds\Bigg)^{\frac{p}{2}}}\\ {} & \hspace{1em}+C\mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\int _{0}^{t}}{\mathcal{K}^{2}}(u,s){\big({b^{\prime }_{y}}\big(u,Y(u)\big)\big)^{2}}\exp \Bigg\{2{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}duds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le C{\Bigg({\int _{0}^{T}}{\mathcal{K}^{2}}(t,s)ds\Bigg)^{\frac{p}{2}}}+C\mathbb{E}\big[{\xi ^{p}}\big]\exp \{pcT\}{\Bigg({\int _{0}^{T}}{\int _{0}^{t}}{\mathcal{K}^{2}}(u,s)duds\Bigg)^{\frac{p}{2}}}\\ {} & \lt \infty ,\end{aligned}\]Our next goal is to establish the Malliavin chain rule for the random variables ${b^{\prime }_{y}}(t,Y(t))$ and $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}$.
Proof.
1) We shall start from proving that ${b^{\prime }_{y}}(t,Y(t))\in {\mathbb{D}^{1,p}}$. Note that ${b^{\prime }_{y}}$ is not a bounded function itself and it does not have bounded derivatives – hence the classical chain rule from [25, Section 1.2] cannot be applied here in a straightforward manner. In order to overcome this issue, we will use the approach in the spirit of [26, Lemma A.1] or [9, Proposition 3.4]. For the reader’s convenience, we divide the proof into steps.
Step 0. First of all, observe that ${b^{\prime }}(t,Y(t))\in {L^{2}}(\Omega )$ as a direct consequence of Proposition 1. Also, for any $p\gt 1$,
\[ \mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\big({b^{\prime\prime }_{yy}}\big(t,Y(t)\big){D_{s}}Y(t)\big)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\lt \infty .\]
Indeed, again by Proposition 1 together with the proof of Proposition 2, we have
\[\begin{aligned}{}\mathbb{E}& \Bigg[{\Bigg({\int _{0}^{T}}{\big({b^{\prime\prime }_{yy}}\big(t,Y(t)\big){D_{s}}Y(t)\big)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le \mathbb{E}\Bigg[{\xi ^{p}}{\Bigg({\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le {\big(\mathbb{E}\big[{\xi ^{2p}}\big]\big)^{\frac{1}{2}}}{\Bigg(\mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\Bigg)^{p}}\Bigg]\Bigg)^{\frac{1}{2}}}\\ {} & \lt \infty .\end{aligned}\]
Therefore, by Lemma 1, it is sufficient to prove that ${b^{\prime }_{y}}(t,Y(t))\in {\mathbb{D}^{1,2}}$ with (15) being the corresponding Malliavin derivative.
Step 1. Let $\phi \in {C^{1}}(\mathbb{R})$ be a compactly supported function such that $\phi (x)=x$ whenever $|x|\le 1$ and $|\phi (x)|\le |x|$ for all $|x|\gt 1$. Fix $t\in [0,T]$ and, for $m\ge 1$, put
Observe that
Therefore, the function ${f_{m}}$ satisfies the conditions of the classical Malliavin chain rule [25, Proposition 1.2.3], so ${f_{m}}(Y(t))\in {\mathbb{D}^{1,2}}$ and, with probability 1 for a.a. $s\in [0,T]$,
\[ {f^{\prime }_{m}}(y)={b^{\prime\prime }_{yy}}(t,y){\phi ^{\prime }}\bigg(\frac{{b^{\prime }_{y}}(t,y)}{m}\bigg)\]
is bounded. Indeed, let $0\lt {\varepsilon _{m}}\lt \psi (t)-\varphi (t)$ be such that
\[ -\frac{{\gamma _{1}}{\theta _{1}}(t)}{{\varepsilon _{m}^{{\gamma _{1}}+1}}}+\underset{\varphi (t)\le x\le \psi (t)}{\max }{a^{\prime }_{y}}(t,x)\lt m\inf \operatorname{supp}\phi \]
and
\[ -\frac{{\gamma _{2}}{\theta _{2}}(t)}{{\varepsilon _{m}^{{\gamma _{2}}+1}}}+\underset{\varphi (t)\le x\le \psi (t)}{\max }{a^{\prime }_{y}}(t,x)\lt m\inf \operatorname{supp}\phi .\]
Then,
-
• if $y\in (\varphi (t),\varphi (t)+{\varepsilon _{m}})$, then\[\begin{aligned}{}{b^{\prime }_{y}}(t,y)& =-\frac{{\gamma _{1}}{\theta _{1}}(t)}{{(y-\varphi (t))^{{\gamma _{1}}+1}}}-\frac{{\gamma _{2}}{\theta _{2}}(t)}{{(\psi (t)-y)^{{\gamma _{2}}+1}}}+{a^{\prime }_{y}}(t,y)\\ {} & \le -\frac{{\gamma _{1}}{\theta _{1}}(t)}{{\varepsilon _{m}^{{\gamma _{1}}+1}}}+\underset{\varphi (t)\le x\le \psi (t)}{\max }{a^{\prime }_{y}}(t,x)\\ {} & \lt m\inf \operatorname{supp}\phi ,\end{aligned}\]so $\frac{{b^{\prime }_{y}}(t,y)}{m}\notin \operatorname{supp}\phi $, ${f_{m}}(y)=0$ and ${f^{\prime }_{m}}(y)=0$;
-
• if $y\in (\psi (t)-{\varepsilon _{m}},\psi (t))$, then, similarly,\[\begin{aligned}{}{b^{\prime }_{y}}(t,y)& =-\frac{{\gamma _{1}}{\theta _{1}}(t)}{{(y-\varphi (t))^{{\gamma _{1}}+1}}}-\frac{{\gamma _{2}}{\theta _{2}}(t)}{{(\psi (t)-y)^{{\gamma _{2}}+1}}}+{a^{\prime }_{y}}(t,y)\\ {} & \le -\frac{{\gamma _{2}}{\theta _{2}}(t)}{{\varepsilon _{m}^{{\gamma _{2}}+1}}}+\underset{\varphi (t)\le x\le \psi (t)}{\max }{a^{\prime }_{y}}(t,x)\\ {} & \lt m\inf \operatorname{supp}\phi ,\end{aligned}\]so $\frac{{b^{\prime }_{y}}(t,y)}{m}\notin \operatorname{supp}\phi $, ${f_{m}}(y)=0$ and ${f^{\prime }_{m}}(y)=0$;
-
• on the compact set $[\varphi (t)+{\varepsilon _{m}},\psi (t)-{\varepsilon _{m}}]$, both ${f_{m}}$ and its derivative ${f^{\prime }_{m}}$ are continuous and hence bounded.
\[ {D_{s}}{f_{m}}\big(Y(t)\big)={b^{\prime\prime }_{yy}}\big(t,Y(t)\big){\phi ^{\prime }}\bigg(\frac{{b^{\prime }_{y}}(t,Y(t))}{m}\bigg){D_{s}}Y(t).\]
Now it remains to prove that
in ${L^{2}}(\Omega )$ and
in ${L^{2}}(\Omega \times [0,T])$ as $m\to \infty $; then the result will follow immediately from the closedness of the Malliavin derivative operator D.
Step 2: ${f_{m}}(Y(t))\to {b^{\prime }}(t,Y(t))$ in ${L^{2}}(\Omega )$ as $m\to \infty $. By the definitions of ${f_{m}}$ and ϕ, ${f_{m}}(Y(t))\to {b^{\prime }}(t,Y(t))$ a.s. as $m\to \infty $. Moreover, with probability 1, $|{f_{m}}(Y(t))|\le |{b^{\prime }_{y}}(t,Y(t))|\in {L^{2}}(\Omega )$ and hence the required convergence follows from the dominated convergence theorem.
Step 3: $D{f_{m}}(Y(t))\to {b^{\prime\prime }_{yy}}(t,Y(t))DY(t)$ in ${L^{2}}(\Omega \times [0,T])$ as $m\to \infty $. By the definitions of ${f_{m}}$ and ϕ, with probability 1,
\[\begin{aligned}{}& {\bigg({b^{\prime\prime }_{yy}}\big(t,Y(t)\big){\phi ^{\prime }}\bigg(\frac{{b^{\prime }_{y}}(t,Y(t))}{m}\bigg)\bigg)^{2}}{\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\\ {} & \hspace{1em}\to {\big({b^{\prime\prime }_{yy}}\big(t,Y(t)\big)\big)^{2}}{\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\end{aligned}\]
as $m\to \infty $. Moreover, since ϕ has compact support, ${\max _{y\in \mathbb{R}}}{({\phi ^{\prime }}(y))^{2}}\lt \infty $, so we can write
\[\begin{aligned}{}{\int _{0}^{T}}{\big({D_{s}}{f_{m}}\big(Y(t)\big)\big)^{2}}ds& ={\bigg({b^{\prime\prime }_{yy}}\big(t,Y(t)\big){\phi ^{\prime }}\bigg(\frac{{b^{\prime }_{y}}(t,Y(t))}{m}\bigg)\bigg)^{2}}{\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\\ {} & \le \underset{y\in \mathbb{R}}{\max }{\big({\phi ^{\prime }}(y)\big)^{2}}{\big({b^{\prime\prime }_{yy}}\big(t,Y(t)\big)\big)^{2}}{\int _{0}^{T}}{\big({D_{s}}Y(t)\big)^{2}}ds\in {L^{2}}(\Omega ).\end{aligned}\]
Therefore, by the dominated convergence theorem,
which proves the first claim of the Proposition.2) Let us proceed with the second claim and verify that
with (16) being the corresponding Malliavin derivative. Note that, since ${b^{\prime }_{y}}$ is bounded from above, $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}$ is also bounded from above and hence is an element of ${L^{p}}(\Omega )$ for any $p\gt 1$. Moreover, by Proposition 1, boundedness of $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}$ and (13), we can write
\[\begin{aligned}{}\mathbb{E}& \Bigg[{\Bigg({\int _{0}^{T}}{\Bigg(\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{\int _{u}^{t}}{b^{\prime\prime }_{yy}}\big(v,Y(v)\big){D_{s}}Y(v)dv\Bigg)^{2}}ds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le C\mathbb{E}\Bigg[{\xi ^{p}}{\Bigg({\int _{0}^{T}}{\int _{u}^{t}}{\big({D_{s}}Y(v)\big)^{2}}dvds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \le C\mathbb{E}\Bigg[{\xi ^{p}}{\Bigg({\int _{0}^{T}}{\int _{u}^{t}}{\mathcal{K}^{2}}(v,s)dvds\Bigg)^{\frac{p}{2}}}\Bigg]\\ {} & \hspace{1em}+C\exp \{pcT\}\mathbb{E}\big[{\xi ^{2p}}\big]{\Bigg({\int _{0}^{T}}{\int _{u}^{t}}{\int _{s}^{v}}{\mathcal{K}^{2}}(u,s)dudvds\Bigg)^{\frac{p}{2}}}\\ {} & \lt \infty ,\end{aligned}\]
and hence it is sufficient to prove that $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}\in {\mathbb{D}^{1,2}}$.Since the Malliavin derivative operator D is closed and the expression ${\textstyle\int _{u}^{t}}{b^{\prime\prime }_{yy}}(v,Y(v)){D_{s}}Y(v)dv$ is well-defined by Proposition 1, Step 1 of the current proof and Hille’s theorem [22, Theorem 1.2.4] guarantee that
and
\[ {D_{s}}{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)Y(v)dv={\int _{u}^{t}}{b^{\prime\prime }_{yy}}\big(v,Y(v)\big){D_{s}}Y(v)dv.\]
Finally, the function $x\mapsto {e^{x}}$ satisfies the conditions of the chain rule from [9, Proposition 3.4] and hence $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}\in {\mathbb{D}^{1,2}}$ and (16) holds. □Corollary 1.
For any $0\le s\lt t\le T$ and $p\gt 1$,
\[ {b^{\prime }_{y}}\big(s,Y(s)\big)\exp \Bigg\{{\int _{s}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\in {\mathbb{D}^{1,p}}\]
and
(17)
\[ \begin{aligned}{}{D_{u}}& \Bigg[{b^{\prime }_{y}}\big(s,Y(s)\big)\exp \Bigg\{{\int _{s}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\Bigg]\\ {} & ={b^{\prime\prime }_{yy}}\big(s,Y(s)\big)\exp \Bigg\{{\int _{s}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{D_{u}}Y(s)\\ {} & \hspace{1em}+{b^{\prime }_{y}}\big(s,Y(s)\big)\exp \Bigg\{{\int _{s}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{\int _{s}^{t}}{b^{\prime\prime }_{yy}}\big(v,Y(v)\big){D_{u}}Y(v)dv.\end{aligned}\]Proof.
For fixed $0\le s\lt t\le T$, denote
\[ {X_{1}}:={b^{\prime }_{y}}\big(s,Y(s)\big),\hspace{1em}{X_{2}}:=\exp \Bigg\{{\int _{s}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}.\]
By Proposition 3 and Lemma 2 from the Appendix, it is sufficient to check that for all $p\ge 2$
All conditions (i)–(iii) can be checked in a straightforward manner using Proposition 1 and the arguments similar to the proof of Proposition 2. □We are now ready to formulate the main result of this section.
Theorem 3.
For any $t\in [0,T]$ and $p\ge 2$,
Proof.
Our goal is to prove that $Y(t)\in {\mathbb{D}^{2,p}}$ and
Item (i) above follows immediately from Corollary 1. As for item (ii), observe that, by Proposition 1, (13) as well as the boundedness of $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}$, we have
\[\begin{aligned}{}{D_{r}}{D_{s}}Y(t)& ={\int _{s}^{t}}\mathcal{K}(u,s){D_{r}}\Bigg[{b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\Bigg]du\\ {} & ={\int _{s}^{t}}\mathcal{K}(u,s){D_{r}}\big[{F_{1}}(t,u)\big]du,\end{aligned}\]
since, in such case, (18) follows immediately from Corollary 1. Recall that
Clearly, for any $0\le r,s\lt t\le T$,
so, by closedness of D and Hille’s theorem [22, Theorem 1.2.4], it is enough to show that
-
(i) for a.a. $0\le s\le u\lt t\le T$, $\mathcal{K}(u,s){F_{1}}(t,u)\in {\mathbb{D}^{1,p}}$ and
-
(ii) for a.a. $0\le s\lt t\le T$,\[\begin{aligned}{}{\int _{0}^{T}}& {\Bigg(\mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\big({D_{r}}\big[\mathcal{K}(u,s){F_{1}}(t,u)\big]\big)^{2}}dr\Bigg)^{\frac{p}{2}}}\Bigg]\Bigg)^{\frac{1}{p}}}du\\ {} & ={\int _{0}^{T}}\mathcal{K}(u,s){\Bigg(\mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\big({D_{r}}\big[{F_{1}}(t,u)\big]\big)^{2}}dr\Bigg)^{\frac{p}{2}}}\Bigg]\Bigg)^{\frac{1}{p}}}du\\ {} & \lt \infty .\end{aligned}\]
\[\begin{aligned}{}{\big({D_{r}}\big[{F_{1}}(t,u)\big]\big)^{2}}& \le C\Bigg({\big({b^{\prime\prime }_{yy}}\big(u,Y(u)\big)\big)^{2}}\exp \Bigg\{2{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}{\big({D_{r}}Y(u)\big)^{2}}\\ {} & \hspace{1em}+{\big({b^{\prime }_{y}}\big(u,Y(u)\big)\big)^{2}}\exp \Bigg\{2{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\times \\ {} & \hspace{2em}\hspace{2em}\times {\int _{u}^{t}}{\big({b^{\prime\prime }_{yy}}\big(v,Y(v)\big){D_{r}}Y(v)\big)^{2}}dv\Bigg)\\ {} & \le C\Bigg({\xi ^{2}}{\big({D_{r}}Y(u)\big)^{2}}+{\xi ^{4}}{\int _{u}^{t}}{\big({D_{r}}Y(v)\big)^{2}}dv\Bigg)\\ {} & \le C{\xi ^{2}}\Bigg({\mathcal{K}^{2}}(u,r)+{\int _{r}^{u}}{\mathcal{K}^{2}}(z,r)dz\Bigg)\\ {} & \hspace{1em}+C{\xi ^{4}}\Bigg({\int _{u}^{t}}{\mathcal{K}^{2}}(v,r)dv+{\int _{u}^{t}}{\int _{r}^{v}}{\mathcal{K}^{2}}(z,r)dzdv\Bigg).\end{aligned}\]
Hence, for any $p\ge 2$, Remark 1 implies
\[\begin{aligned}{}{\int _{0}^{T}}{\big({D_{r}}\big[{F_{1}}(t,u)\big]\big)^{2}}dr& \le C{\xi ^{2}}\Bigg({\int _{0}^{T}}{\mathcal{K}^{2}}(u,r)dr+{\int _{0}^{T}}{\int _{r}^{u}}{\mathcal{K}^{2}}(z,r)dzdr\Bigg)\\ {} & \hspace{1em}+C{\xi ^{4}}{\int _{0}^{T}}{\int _{u}^{t}}{\mathcal{K}^{2}}(v,r)dvdr\\ {} & \hspace{1em}+C{\xi ^{4}}{\int _{0}^{T}}{\int _{u}^{t}}{\int _{r}^{v}}{\mathcal{K}^{2}}(z,r)dzdvdr\\ {} & \le C\big({\xi ^{2}}+{\xi ^{4}}\big),\end{aligned}\]
so, since ξ has moments of all orders, (ii) holds, which finalizes the proof. □Finally, denote ${\mathbb{L}^{2,p}}:={L^{p}}([0,T];{\mathbb{D}^{2,p}})$. We complete the section with the following result.
Proof.
By the definition of the $\| \cdot {\| _{2,p}}$-norm in (32) from Appendix A, it is sufficient to check that
and
By (11), (19) holds automatically. Next, (20) can be easily deduced from (14). Finally, using Proposition (1) and the boundedness of $\exp \{{\textstyle\int _{u}^{t}}{b^{\prime }_{y}}(v,Y(v))dv\}$, it is easy to prove a bound similar to (14) for
(21)
\[ {\int _{0}^{T}}\mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\int _{0}^{T}}{\big({D_{r}}{D_{s}}Y(t)\big)^{2}}dsdr\Bigg)^{\frac{p}{2}}}\Bigg]dt\lt \infty .\]
\[ \mathbb{E}\Bigg[{\Bigg({\int _{0}^{T}}{\int _{0}^{T}}{\big({D_{r}}{D_{s}}Y(t)\big)^{2}}dsdr\Bigg)^{\frac{p}{2}}}\Bigg],\]
which implies (21). By this, the proof is complete. □4 Power law in SVV model
Having the second-order Malliavin differentiability in place, we now possess all the necessary tools to analyze the behavior of implied volatility skew of a model with the sandwiched process (10) as stochastic volatility. Namely, we consider a (risk-free) market model with the price process $S=\{S(t),\hspace{2.5pt}t\in [0,T]\}$ of the form
where ${B_{1}},{B_{2}}$ are two independent Brownian motions, $X=\{X(t),t\in [0,T]\}$ denotes the (risk-free) log-price of an asset starting from some level ${x_{0}}\in \mathbb{R}$, r is a constant instantaneous interest rate, and $\rho \in (-1,1)$ is a correlation coefficient that accounts for the leverage effect. As previously, the drift b and the Volterra kernel $\mathcal{K}$ satisfy Assumptions 1 and 2.
(22)
\[ \begin{aligned}{}S(t)& ={e^{X(t)}},\\ {} X(t)& ={x_{0}}+rt-\frac{1}{2}{\int _{0}^{t}}{Y^{2}}(s)ds+{\int _{0}^{t}}Y(s)\big(\rho d{B_{1}}(s)+\sqrt{1-{\rho ^{2}}}d{B_{2}}(s)\big),\\ {} Y(t)& ={y_{0}}+{\int _{0}^{t}}b\big(s,Y(s)\big)ds+{\int _{0}^{t}}\mathcal{K}(t,s)d{B_{1}}(s),\end{aligned}\]The goal of this section is to establish conditions under which (22) reproduces the power law (2) of the short-term at-the-money implied volatility. Namely, we have the following result.
Theorem 4.
Let Assumptions 1 and 2 hold with $H\in (\frac{1}{6},\frac{1}{2})$. Assume that the Volterra kernel $\mathcal{K}$ is such that, for any $0\le s\lt t\le T$,
for some constant $C\gt 0$, and
for some finite constant ${K_{Y}}$. Then, with probability 1, the SVV implied volatility $\widehat{\sigma }$ exhibits the property
(23)
\[ \frac{1}{{\tau ^{\frac{3}{2}+H}}}{\int _{0}^{\tau }}{\int _{s}^{\tau }}\mathcal{K}(t,s)dtds\to {K_{Y}},\hspace{1em}\tau \to 0+,\]
\[ \underset{\tau \to 0}{\lim }{\tau ^{\frac{1}{2}-H}}\frac{\partial \widehat{\sigma }}{\partial \kappa }(\tau ,\kappa ){\bigg|_{\kappa =0}}=\frac{\rho }{{y_{0}}}{K_{Y}}.\]
In particular, if $\rho {K_{Y}}\ne 0$, the SVV model (22) reproduces the power law (2) of the at-the-money implied volatility skew.
Remark 6.
The behavior of empirically observed implied volatilities (see, e.g., [12]) shows that realistic market models should produce $\widehat{\sigma }$ with
In the SVV setting (22), Theorem 4 guarantees that (24) holds for all small enough τ provided that $\rho {K_{Y}}\lt 0$.
(24)
\[ \frac{\partial \widehat{\sigma }}{\partial \kappa }(\tau ,\kappa ){\bigg|_{\kappa =0}}\lt 0.\]To prove Theorem 4, we will apply the fundamental result [1, Theorem 6.3] which connects the shape of the skew with the Malliavin derivative of the volatility.
Remark 8.
In the recent literature (see, e.g., [4, 8, 12, 20]), it is typical to characterize the implied volatility skew in terms of $\frac{\partial \widehat{\sigma }}{\partial \kappa }$ with $\kappa =\log \frac{K}{{e^{r\tau +{x_{0}}}}}$ being the log-moneyness. In [1], a slightly different parametrization ${\widehat{\sigma }_{\text{log-price}}}(\tau ,{x_{0}})$ is considered with
\[ {\widehat{\sigma }_{\text{log-price}}}(\tau ,x)=\widehat{\sigma }\bigg(\tau ,\log \frac{K}{{e^{r\tau }}}-x\bigg).\]
With this parametrization,
\[ \frac{\partial {\widehat{\sigma }_{\text{log-price}}}(\tau ,x)}{\partial x}=-\frac{\partial \widehat{\sigma }(\tau ,\log \frac{K}{{e^{r\tau }}}-x)}{\partial \kappa }\]
and the power law (2) is equivalent to
Theorem 5.
Consider a risk-free log-price
where ${B_{1}}$, ${B_{2}}$ are two independent Brownian motions, ${x_{0}}\in \mathbb{R}$ is a deterministic initial value, r is an instantaneous interest rate, $\rho \in (-1,1)$ is a correlation coefficient and $\sigma =\{\sigma (t),\hspace{2.5pt}t\in [0,T]\}$ is a square-integrable stochastic process with right-continuous trajectories adapted to the filtration $\mathcal{F}=\{{\mathcal{F}_{t}},\hspace{2.5pt}t\in [0,T]\}$ generated by ${B_{1}}$.
(25)
\[ X(t)={x_{0}}+rt-\frac{1}{2}{\int _{0}^{t}}{\sigma ^{2}}(s)ds+{\int _{0}^{t}}\sigma (s)\big(\rho d{B_{1}}(s)+\sqrt{1-{\rho ^{2}}}d{B_{2}}(s)\big),\]
Assume that
Finally, assume that there exists a constant ${K_{\sigma }}\gt 0$ such that, with probability 1,
Then, with probability 1,
-
(H1) $\sigma \in {\mathbb{L}^{2,4}}$ with respect to ${B_{1}}$;
-
(H2) there exists a constant ${\varphi _{\ast }}\gt 0$ such that, with probability 1, $\sigma (t)\gt {\varphi _{\ast }}$ for all $t\in [0,T]$;
-
(H4) σ has a.s. right-continuous trajectories;
-
(H5) ${\sup _{r,s,t\in [0,\tau ]}}\mathbb{E}[{(\sigma (s)\sigma (t)-{\sigma ^{2}}(r))^{2}}]\to 0$ when $\tau \to 0+$.
Remark 9.
Observe that the SVV model (22) automatically satisfies a number of assumptions of Theorem 5:
Therefore, it remains to check (H3), (H5), and (28). Naturally, given the shape of the Malliavin derivative (12), both (H3) and (28) require additional assumptions on the kernel, so let us start with (H5).
-
• assumption (H2) with ${\varphi ^{\ast }}:={\min _{t\in [0,T]}}\varphi (t)\gt 0$;
-
• assumption (H4) since Y is continuous a.s.;
-
• assumption (H1) by the results of Section 3 above.
Proof.
By [10, Lemma 3.6], there exists a positive random variable $\Upsilon ={\Upsilon _{T}}$ such that, for all ${t_{1}},{t_{2}}\in [0,T]$,
and, for any $r\gt 0$,
Therefore, given that ${\max _{t\in [0,T]}}Y(t)\lt {\max _{t\in [0,T]}}\psi (t)$ by (11),
\[\begin{aligned}{}\mathbb{E}& \big[{\big(Y(s)Y(t)-{Y^{2}}(r)\big)^{2}}\big]\\ {} & =\mathbb{E}\big[{\big(Y(s)\big(Y(t)-Y(r)\big)+Y(r)\big(Y(s)-Y(r)\big)\big)^{2}}\big]\\ {} & \le 2\mathbb{E}\big[({Y^{2}}(s){\big(Y(t)-Y(r)\big)^{2}}\big]+2\mathbb{E}\big[{Y^{2}}(r){\big(Y(s)-Y(r)\big)^{2}}\big]\\ {} & \le 2|t-r{|^{2\lambda }}\underset{s\in [0,T]}{\max }{\psi ^{2}}(s)\mathbb{E}\big[{\Upsilon ^{2}}\big]+2|s-r{|^{2\lambda }}\underset{s\in [0,T]}{\max }{\psi ^{2}}(s)\mathbb{E}\big[{\Upsilon ^{2}}\big]\end{aligned}\]
and hence, with probability 1,
\[\begin{aligned}{}\underset{r,s,t\in [0,\tau ]}{\sup }\mathbb{E}\big[{\big(Y(s)Y(t)-{Y^{2}}(r)\big)^{2}}\big]& \le 4{\tau ^{2\lambda }}\underset{s\in [0,T]}{\max }{\psi ^{2}}(s)\mathbb{E}\big[{\Upsilon ^{2}}\big]\to 0\end{aligned}\]
as $\tau \to 0+$. □Our next step is to handle (28).
Proof.
Recall that
\[ {F_{1}}(t,u):={b^{\prime }_{y}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\]
and that, by Proposition 1,
where $c:={\max _{(t,y)\in \overline{\mathcal{D}}}}{a^{\prime }_{y}}(t,y)$. Then we can write
\[\begin{aligned}{}\frac{1}{{\tau ^{\frac{3}{2}+H}}}& {\int _{0}^{\tau }}{\int _{s}^{\tau }}\mathbb{E}\big[{D_{s}}Y(t)\big]dtds\\ {} & =\frac{1}{{\tau ^{\frac{3}{2}+H}}}{\int _{0}^{\tau }}{\int _{s}^{\tau }}\mathcal{K}(t,s)dtds\\ {} & \hspace{1em}+\frac{1}{{\tau ^{\frac{3}{2}+H}}}{\int _{0}^{\tau }}{\int _{s}^{\tau }}{\int _{s}^{t}}\mathcal{K}(u,s)\mathbb{E}\big[{F_{1}}(t,u)\big]dudtds\\ {} & =\frac{1}{{\tau ^{\frac{3}{2}+H}}}{\int _{0}^{\tau }}{\int _{s}^{\tau }}\mathcal{K}(t,s)dtds\\ {} & \hspace{1em}+\frac{1}{{\tau ^{\frac{3}{2}+H}}}{\int _{0}^{\tau }}{\int _{s}^{\tau }}\mathcal{K}(u,s)\Bigg({\int _{u}^{\tau }}\mathbb{E}\big[{F_{1}}(t,u)\big]dt\Bigg)duds.\end{aligned}\]
The term $\frac{1}{{\tau ^{\frac{3}{2}+H}}}{\textstyle\int _{0}^{\tau }}{\textstyle\int _{s}^{\tau }}\mathcal{K}(t,s)dtds$ converges to ${K_{Y}}$ by (23). As for the second term, note that, with probability 1, for any $u\in [0,\tau ]$,
\[\begin{aligned}{}{\int _{u}^{\tau }}\big|\mathbb{E}\big[{F_{1}}(t,u)\big]\big|dt& \le C\mathbb{E}[\xi ]\tau \end{aligned}\]
and hence, given (23), with probability 1,
which ends the proof. □Finally, let us deal with (H3).
Proposition 6.
Proof.
Fix $0\lt r,s\lt t$. Then, taking into account (29), with probability 1,
which immediately implies (26). Next, by Proposition 1,
for any $v\in [0,T]$ and, for any $0\le u\le t\le T$,
(31)
\[ \begin{aligned}{}\big|{D_{s}}Y(t)\big|& \le \big|\mathcal{K}(t,s)\big|+{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{F_{1}}(t,u)\big|du\\ {} & \le C\Bigg(|t-s{|^{-\frac{1}{2}+H}}+\xi {\int _{s}^{t}}|u-s{|^{-\frac{1}{2}+H}}du\Bigg)\\ {} & \le C(1+T\xi )|t-s{|^{-\frac{1}{2}+H}}\\ {} & =:\zeta |t-s{|^{-\frac{1}{2}+H}},\end{aligned}\]
\[\begin{aligned}{}\big|{F_{2}}(t,u)\big|& =\Bigg|{b^{\prime\prime }_{yy}}\big(u,Y(u)\big)\exp \Bigg\{{\int _{u}^{t}}{b^{\prime }_{y}}\big(v,Y(v)\big)dv\Bigg\}\Bigg|\\ {} & \le {e^{cT}}\xi \end{aligned}\]
with $c:={\max _{(t,y)\in \overline{\mathcal{D}}}}{a^{\prime }_{y}}(t,y)$, so we can write
\[\begin{aligned}{}& \big|{D_{r}}{D_{s}}Y(t)\big|\\ {} & \hspace{1em}\le {\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{F_{1}}(t,u)\big|\Bigg({\int _{u}^{t}}\big|{b^{\prime\prime }_{yy}}\big(v,Y(v)\big)\big|\big|{D_{r}}Y(v)\big|dv\Bigg)du\\ {} & \hspace{2em}+{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{F_{2}}(t,u)\big|\big|{D_{r}}Y(u)\big|du\\ {} & \hspace{1em}\le C\Bigg({\xi ^{2}}{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\Bigg({\int _{u}^{t}}\big|{D_{r}}Y(v)\big|dv\Bigg)du+\xi {\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{D_{r}}Y(u)\big|du\Bigg)\\ {} & \hspace{1em}=C\Bigg({\xi ^{2}}{\int _{s}^{t}}\big|\mathcal{K}(u,s)\big|\Bigg({\int _{u\vee r}^{t}}\big|{D_{r}}Y(v)\big|dv\Bigg)du+\xi {\int _{r\vee s}^{t}}\big|\mathcal{K}(u,s)\big|\big|{D_{r}}Y(u)\big|du\Bigg).\end{aligned}\]
Taking into account (30) and (31),
\[\begin{aligned}{}\big|{D_{r}}{D_{s}}Y(t)\big|& \le C\Bigg({\xi ^{2}}\zeta {\int _{s}^{t}}|u-s{|^{-\frac{1}{2}+H}}\Bigg({\int _{u\vee r}^{t}}|v-r{|^{-\frac{1}{2}+H}}dv\Bigg)du\\ {} & \hspace{2em}+\xi \zeta {\int _{r\vee s}^{t}}|u-s{|^{-\frac{1}{2}+H}}|u-r{|^{-\frac{1}{2}+H}}du\Bigg)\\ {} & \le C\Bigg({\xi ^{2}}\zeta {\int _{s}^{t}}|u-s{|^{-\frac{1}{2}+H}}|t-r{|^{\frac{1}{2}+H}}du\\ {} & \hspace{2em}+\xi \zeta {\int _{r\vee s}^{t}}|u-s{|^{-\frac{1}{2}+H}}|u-r{|^{-\frac{1}{2}+H}}du\Bigg).\end{aligned}\]
Note that
\[\begin{aligned}{}{\int _{s}^{t}}|u-s{|^{-\frac{1}{2}+H}}|t-r{|^{\frac{1}{2}+H}}du& \le C|t-r{|^{\frac{1}{2}+H}}|t-s{|^{\frac{1}{2}+H}}\\ {} & \le C{\bigg(\frac{t-r}{t-s}\bigg)^{\frac{1}{2}-H}}.\end{aligned}\]
As for the integral ${\textstyle\int _{r\vee s}^{t}}|u-s{|^{-\frac{1}{2}+H}}|u-r{|^{-\frac{1}{2}+H}}du$, we have two cases:
In any case,
\[\begin{aligned}{}\big|{D_{r}}{D_{s}}Y(t)\big|& \le C\xi \zeta (\xi +1){\bigg(\frac{t-r}{t-s}\bigg)^{\frac{1}{2}-H}},\end{aligned}\]
where ξ and ζ are random variables having all moments, and hence (27) holds. □Having in mind all of the results above, we are ready to prove the main result of this section, namely Theorem 4.
Example 1.
Let $\frac{1}{6}\lt {H_{0}}\lt {H_{1}}\lt \cdots \lt {H_{n}}\lt 1$ be such that ${H_{0}}\lt \frac{1}{2}$ and ${\alpha _{k}}\gt 0$, $k=0,\dots ,n$. Then the kernel
\[ \mathcal{K}(t,s)=\Bigg({\sum \limits_{k=0}^{n}}{\alpha _{k}}{(t-s)^{{H_{k}}-\frac{1}{2}}}\Bigg){1_{s\lt t}}\]
satisfies the assumptions of Theorem 4, so the corresponding SVV model generates power law (2) with $H={H_{0}}$ provided that $\rho \lt 0$ in (22).