1,688
Views
3
CrossRef citations to date
0
Altmetric
Research Papers

Short-dated smile under rough volatility: asymptotics and numerics

, &
Pages 463-480 | Received 27 Oct 2020, Accepted 15 Oct 2021, Published online: 07 Dec 2021

Abstract

In Friz et al. [Precise asymptotics for robust stochastic volatility models. Ann. Appl. Probab, 2021, 31(2), 896–940], we introduce a new methodology to analyze large classes of (classical and rough) stochastic volatility models, with special regard to short-time and small-noise formulae for option prices, using the framework [Bayer et al., A regularity structure for rough volatility. Math. Finance, 2020, 30(3), 782–832]. We investigate here the fine structure of this expansion in large deviations and moderate deviations regimes, together with consequences for implied volatility. We discuss computational aspects relevant for the practical application of these formulas. We specialize such expansions to prototypical rough volatility examples and discuss numerical evidence.

1. Introduction

In Friz et al. (Citation2021), precise short-time asymptotics were established for call and put option prices under stochastic volatility, under a set of abstract conditions satisfied by most classical and rough volatility (RoughVol) models. These results are refinements of large deviation statements, providing the higher-order, algebraic term in an asymptotic expression, known as Laplace expansion. For RoughVol models, short-dated large deviation pricing is due to Forde and Zhang (Citation2017), as is the induced implied volatility expansion (FZ expansion), which can be seen as a “rough” BBF (Berestycki–Busca–Florent Berestycki et al. Citation2004) formula. Our precise asymptotics provide a mechanism to compute refined implied volatility expansions, for log-strike kt=xt1/2H, of the form (1) σBS2(t,kt)=Σ2(x)+t2Ha(x)+o(t2H) as t0,(1) where the zero-order Σ(x) term corresponds to the rough BBF formula in Forde and Zhang (Citation2017). The next-order term is seen of order t2H and hence increasingly important for small Hurst parameter H, the basic premise of RoughVol modeling. Inclusion of this term hinges on an accurate evaluation of a. In this paper, we assume that the volatility process is of the form σ(Wˆt,t2H), where Wˆ is the Riemann–Liouville fractional Brownian motion (fBM) given by the self-similar Gaussian Volterra process in (EquationA3). It has Hurst exponent H(0,1/2] and it is ρ-correlated with the Brownian driving the asset.

The functions Σ(x) and a(x) do not have explicit expressions and we discuss how to compute them numerically. Following Forde and Zhang (Citation2017), Σ(x) can be computed using the Ritz method. Moreover, we propose a method for computing a(x) based on a Karhunen–Loeve (KL) decomposition of the Brownian motions. (This entails a numerical approximation to an infinite-dimensional Carleman–Fredholm determinant.)

We also derive near-the-money (meaning, as x0) expansions of Σ(x) and of the term structure a(x) which can alternatively be used for numerics (and have the advantage of being explicit functions of model parameters). From these asymptotics, we derive consequences for at-the-money (ATM) implied skew and curvature. We also refine some moderate deviation asymptotics for call prices and implied volatilities, cf. (Friz et al. Citation2017, Bayer et al. Citation2019, Gulisashvili Citation2020, Jacquier and Pannier Citation2020).

Being able to evaluate Σ(x) and a(x) allows us to test the accuracy of the short-time asymptotics in practice. We do so with a numerical case study of the rough Bergomi (rBergomi) model. To exploit our general framework, we look at a volatility given by (2) σ(Wˆt,t2H)=σ0exp(η2Wˆtθη24t2H),(2) so that for θ=1 we get the rBergomi model considered in Bayer et al. (Citation2016) and Bennedsen et al. (Citation2017) with constant forward variance, for θ=0 the rBergomi version in Bayer et al. (Citation2019) and Forde and Zhang (Citation2017). Note, however, that (Equation2) is a genuine rBergomi model for any value of θ, as discussed in Remark 4.1. We compare our approximation to the FZ expansion from Forde and Zhang (Citation2017) and to the Edgeworth asymptotics in El Euch et al. (Citation2019). We consider how smiles vary as θ varies in (Equation2) and as expiry t increases. We discuss and test the volatility term structure and its slope ATM, and observe how the term a(x)t2H improves the asymptotics as H decreases. We observe the same feature when we implement the moderate deviation asymptotics for implied volatility, where for H small the inclusion of the term structure correction a(0)t2H significantly improves on the numerical results presented in Bayer et al. (Citation2019).

Proofs rely on stochastic Taylor expansions, rate function representations in Forde and Zhang (Citation2017) and Bayer et al. (Citation2019) and on the local analysis on the Wiener space introduced in Friz et al. (Citation2021) and Bayer et al. (Citation2020). The classical Gao–Lee results (Gao and Lee Citation2014) are used to go from option prices to implied volatility asymptotics both in large and moderate deviation regimes.

Rough Volatility. It has been shown in recent years that RoughVol models provide great fits to observed volatility surfaces (Bayer et al. Citation2016) capturing fundamental stylized facts of implied volatility in a parsimonious way. Specifically, this class of models can reproduce the steep short end of the smile, displaying exploding implied skew (Alòs et al. Citation2007, Fukasawa Citation2011Citation2017), and they are the only models consistent with the power law of the skew (Bayer et al. Citation2016, Lee Citation2005) not admitting arbitrage (Fukasawa Citation2021). RoughVol is also supported by statistical and time series analysis (Gatheral et al. Citation2018, Fukasawa et al. Citation2019, Bennedsen et al. Citation2021) and by market microstructure considerations (El Euch et al. Citation2018). Many authors have even argued that H0, such as to be consistent with a skew explosion close to t1/2 (Bayer et al. Citation2016Citation2021). One main aspect of RoughVol is non-Markovianity. This is a serious complication when it comes to pricing, as Monte Carlo methods become more expensive and PDE methods are not available. For this reason, efficient simulation schemes have been proposed (Bayer et al. Citation2020, Bennedsen et al. Citation2017, McCrickerd and Pakkanen Citation2018). Fourier-based methods are available for the rough Heston model (El Euch and Rosenbaum Citation2019). Deep and machine learning approaches have also recently been discussed in Bayer et al. (Citation2019) and Goudenège et al. (Citation2020). Small maturity approximations are used in this context to obtain starting points for calibration procedures, which are then based on numerical evaluations.

Asymptotic option pricing. Classical motivation for (semi-closed form) asymptotic pricing includes fast calibration and a quantitative understanding of the impact of model parameters on relevant quantities such as implied skew and curvature/convexity along the moneyness dimension or slope along the term-structure dimension. Explicit expressions for such quantities (that follow in this setting from our expansion) and their shape characteristics are also used to choose the most appropriate model to be fitted to data (Ait-Sahalia et al. Citation2020), leave alone being the origin of some widely used parametrisations of the volatility surface. An interesting, if recent, addition to this list comes from a machine learning perspective: the form of an expansion such as (Equation1) may be viewed as expert knowledge, which significantly narrows the learning task to finer information such as the error in that expansions; it is equally conceivable to learn a=a(x) and other components in the expansion.

Under Markovian stochastic volatility, expansion (Equation2) is analogous, e.g. to the result derived in Forde et al. (Citation2012) for the Heston model. There, the term structure is a(x)t (due to the diffusive scaling of the volatility), whereas here the correction term is a(x)t2H (due to the rough scaling of the volatility). Similar expansions are derived also in Osajima (Citation2015), for more general Markovian models, and (formally) in Medvedev and Scaillet (Citation2003) and Medvedev and Scaillet (Citation2007) for Markov stochastic volatility models with jumps.

In recent years several authors have studied the short-time behavior of RoughVol models. Theoretical results on short-time skew and curvature are given in Fukasawa (Citation2017) and Alòs and León (Citation2017). A second-order short-time expansion is given in El Euch et al. (Citation2019) for general (rough) stochastic volatility models. In Jacquier et al. (Citation2018), the pathwise large deviation behavior under rBergomi dynamics is studied. Pathwise large and moderate deviation principles for (possibly rough) Gaussian stochastic volatility models are established in Gulisashvili (Citation2020) and Gulisashvili (Citation2020), together with asymptotic results at the central limit (Edgeworth) regime. For the rough Heston model, the recent work (Forde et al. Citation2020) provides call expansions of the same type as ours, involving the energy function and the first-order algebraic term, at the same large deviations regime kt=xt1/2H. (The rigid infinite-dimensional affine structure which underlies (Forde et al. Citation2020) is not available for rBergomi type models as considered in this work.) As already mentioned, our work builds on the large deviations principle proved in Forde and Zhang (Citation2017) for models with volatility σ(Wˆt), and on Bayer et al. (Citation2019), where the at-the-money behavior of the Forde–Zhang rate function is used to prove moderate deviation principles and implied volatility asymptotics for the same type of models. The theoretical foundations of the present paper are given in Friz et al. (Citation2021).

In Section 2, we explain our RoughVol setting. In Section 3, we state and comment our results. In Section 4 we discuss and implement our results in the case of the rBergomi model. In Section 5, we show how Σ and a can be computed using Ritz method and KL decomposition. We collect all the proofs in Section 6.

2. Preliminaries on rough volatility

We consider the following RoughVol model, with H(0,1/2], normalized to rate r = 0 and S0=1 (3) dStSt=σ(Wˆt,t2H)d(ρWt+ρ¯W¯t),(3) where W,W¯ are independent Brownian motions (BM) and ρ(1,1), ρ2+ρ¯2=1. We also write W~=ρW+ρ¯W¯. Moreover, Wˆ=(Wˆt)t0 is a Gaussian Volterra process of the form (4) Wˆt=(KW˙)t=0tK(t,s)dWs,(4) for a kernel K(t,s) such that Wˆ is self-similar with exponent H(0,1/2], meaning (5) Law(Wˆϵ2t:tt¯)=ϵ2HLaw(Wˆt:tt¯),for some t¯>0.(5) The BM W drives the stochastic “rough” volatility, meaning (with abusive notation) that σ(t,ω)=σ(Wˆt,t2H), where σ(x,y) is a smooth deterministic real-valued function. We denote σ(x,y)=xσ(x,y), σ(x,y)=xxσ(x,y), σ˙(x,y)=yσ(x,y). We also denote σ0=σ(0,0) >0 the spot volatility and (6) σ0=σ(0,0),σ0=σ(0,0),σ˙0=σ˙(0,0),(6) the derivatives of the volatility function at the initial condition. We consider a dependence in t2H in σ(), because this is the scaling of the variance of the fBm at time t. For this reason, this is the scaling of the time-dependent term in the rBergomi model, and also the scaling such that we observe a dependence in σ˙0 in our precise asymptotics. We apply the abstract results proved in Friz et al. (Citation2021) for K(t,s)=const×(ts)H1/2. However, we expect these approximations to hold in greater generality: the same type of expansions should hold for other kernels such that Wˆ in (Equation4) satisfies (Equation5). Self-similarity is equivalent to the fact that K can be written in the following form (7) K(t,s)=(ts)H1/2fK(s/t),(7) for a suitable function fK (see Jost Citation2007, Lemma 2.4), so that all such kernels can be seen as a perturbation of (ts)H1/2. Two classical processes of this form are the Mandelbrot–Van Ness and the Riemann–Liouville fBMs (see Appendix). Without loss of generality, we also assume K(t,s)=0 for t<s.

A similar setting has been considered in Forde and Zhang (Citation2017) and Bayer et al. (Citation2019). The main difference in the structure of the model is that here we allow for a direct dependence on time in σ(t,ω)=σ(Wˆt,t2H), whereas in Forde and Zhang (Citation2017) and Bayer et al. (Citation2019) the volatility function depends only on the fBM, so σ(t,ω)=σ(Wˆt). As mentioned in the introduction, assuming that the volatility is a deterministic function only of the fBM rules out the rBergomi model σ(Wˆt,t2H)=σ0exp(ηWˆt/2η2t2H/4), see Bayer et al. (Citation2016) and Bennedsen et al. (Citation2017), from the analysis, so a modified version of rBergomi is considered in Bayer et al. (Citation2019). We discuss in detail both versions of this model in Section 4. With a volatility function σ(Wˆt,t2H), one can write the dynamics of the log-price X=logS as (8) Xt=0tσ(Wˆs,s2H)d(ρ¯W¯+ρW)s120tσ2(Wˆs,s2H)ds.(8) In this case, a LDP holds, writing ϵˆ=ϵ2H, for (9) X¯1ϵ=01σ(ϵˆWˆt,ϵˆ2t2H)ϵˆd(ρ¯W¯+ρW)t12ϵϵˆ01σ2(ϵˆWˆt,ϵˆ2t2H)dt,(9) with speed ϵˆ2 and rate function (10) Λ(x):=inf{12h,h¯H12:01σ(hˆ,0)d(ρ¯h¯+ρh)=x}12hx,h¯xH12,(10) where hˆt=(Kh˙)t and H1 is the Cameron–Martin norm. The existence of a minimizer above is obtained from a standard compactness argument. Through the space-time scaling t=ϵ2 and the fact that, in law, X¯1ϵ=ϵˆϵXϵ2, this small-noise LDP translates to a short-time LDP. This result was proved for σ(Wˆt,t2H)=σ(Wˆt) in Forde and Zhang (Citation2017) and then extended to possible dependence in t2H in Friz et al. (Citation2021, Section 7.3). In general, when looking only at large (or moderate) deviations, the t2H-dependence in σ() does not affect the analysis, and the large (or moderate) deviations behavior is the same one would get with volatility σ(Wˆt,0). In Friz et al. (Citation2021), we consider a general asymptotic setting, obtaining for generic stochastic volatility models (including RoughVol ones) precise asymptotics that refine such large deviations asymptotics. For such refinement, this t2H-dependence actually affects the asymptotics. In the present paper, we provide computationally relevant results that allow for the practical usage of such refined pricing asymptotics and discuss their consequences on the Black–Scholes implied volatility.

3. Results

We consider call and put prices under model (Equation8), i.e. c(t,k)=E[(expXtexpk)+],p(t,k)=E[(expkexpXt)+], where k is the log-strike (or log-moneyness). In Friz et al. (Citation2021, Theorem 1.1) we obtain precise small-noise price expansions for generic (classical and rough) volatility dynamics. As in the classical Brownian case, such small-noise results can be translated into short-time results writing t=ϵ2. In this paper, we focus on the short-time setting. We write ∼ for asymptotic equivalence, ftgt if ft/gt1 as t0, and “≈” for “is close to” in informal terms. We also write σx2=2Λ(x)/Λ(x)2.

Assumption 3.1

Throughout the paper, we assume K in (Equation4) is of the form K(t,s)=const×(ts)H1/2.

In short-time, Friz et al. (Citation2021, Theorem 1.1) reads as follows:

Theorem 3.2

Let H(0,1/2] and kt=xt1/2H. Assume that a LDP holds for c, p above, and the existence of 1+ moments for expXt. Then , for x>0 small enough, the rate function Λ=Λ(x) is continuously differentiable at x and c(t,kt)exp(Λ(x)t2H)t1/2+2HA(x)(Λ(x))2σx2π as t0, for some function A(x) with A(x)1 as x0. Similarly, for x<0, close enough to 0, we have p(t,kt)exp(Λ(x)t2H)t1/2+2HA(x)(Λ(x))2σx2π as t0, for some function A(x) with A(x)1 as x0. Moreover, such A can be expressed as (11) A(x)={E[exp(Λ(x)Δ2x)],if H<1/2,exE[exp(Λ(x)Δ2x)],if H=1/2,(11) where Δ2x is a certain quadratic Wiener functional (specified in Friz et al. (Citation2021, Equation (7.4)), see also (Equation36) below).

Remark 3.3

The fact that x>0 above has to be taken small enough is in order for the minimizer (hx,h¯x) in (Equation10) to be unique and non-degenerate. The latter means, in a nutshell, that the Hessian of I(h,h¯):=12h,h¯H12 is strictly positive when restricted to those (h,h¯) such that 01σ(hˆ,0)d(ρ¯h¯+ρh)=x, and is equivalent to the finiteness of A(x) defined above.

We write Kf(t)=0tK(t,s)f(s)ds, K2f(t)=0tK2(t,s)f(s)ds and , for the inner product in L2[0,1]. We also denote K¯ the adjoint of K in L2[0,1] so that K¯1(u)=u1K(t,u)dt. Fully explicit expressions are computable in the case of the Riemann–Liouville fBM (Appendix) and in particular in the case of standard BM (this is the classical case of Markovian stochastic volatility). We denote CK,ρ=K21,1232(K¯1)2,1+ρ2(172K1,1232(K1)2,13K1,K¯1),C¯K,ρ=K21,1232ρ2(K1)2,1.

Lemma 3.4

Fine structure of A

For H(0,1/2], the following expansion holds for A(x) as x0: (12) A(x)=1xρσ0K1,1σ02+x2((σ0)2σ04CK,ρ+σ0σ03C¯K,ρ+σ˙0(2H+1)σ03)+(x2+x28)1{H=1/2}+O(x3).(12)

As a consequence of Theorem 3.2 the following expansion holds for the Black–Scholes implied volatility (by a standard application of Gao and Lee (Citation2014), detailed in Friz et al. (Citation2021, Appendix D)).

Corollary 3.5

Asymptotic smile and term structure at the large deviations regime

Writing kt=xt1/2H, we have the following expansion, for xR{0} such that Theorem 3.2 holds: (13) σBS2(t,kt)=Σ2(x)+t2Ha(x)+o(t2H) as t0,(13) where (14) Σ(x)=|x|2Λ(x)(14) and (15) a(x)={x22Λ(x)2log(2A(x)Λ(x)Λ(x)x)if H<1/2,x22Λ(x)2log(2A(x)Λ(x)Λ(x)xexp(x/2))if H=1/2.(15)

Remark 3.6

In general, from a LDP for call prices follows the celebrated BBF formula for implied volatility (Berestycki–Busca–Florent Berestycki et al. Citation2004, see also Pham Pham Citation2010 for a derivation). Under RoughVol pricing with σ(ω,t)=σ(Wˆt), this has been extended in Forde and Zhang (Citation2017) to (16) σBS2(t,kt)x22Λ(x),(16) holding for fixed x, in short-time, with kt=xt1/2H. Thanks to the A-term in (Equation12), we can extend this approximation, adding the term structure t2Ha(x). Note that the expansions hold for H(0,1/2], but for H=1/2 their functional form is different, as some additional terms appear in A(x) and in the term structure of the Black–Scholes implied volatility a(x).

We denote now (17) DK,ρ=K21,1(K¯1)2,1+ρ2(3K1,12(K1)2,12K1,K¯1),D¯K,ρ=K21,1ρ2(K1)2,1.(17) The short-time implied volatility coefficients in the previous statement can be expanded as follows near-the-money.

Theorem 3.7

At-the-money expansion of the coefficients

For x0, the Σ coefficient has the following expansion: (18) Σ(x)=σ0+xΣ(0)+x2Σ(0)2+O(x3),(18) where Σ(0)=ρσ0K1,1σ0,Σ(0)2=(σ0)2σ03{3ρ2K1,12+ρ22(K1)2,1+12(K¯1)2,1+ρ2K1,K¯1}+σ0σ02ρ22(K1)2,1. The term structure coefficient, at the first order in x at 0, is (19) a(x)=a0+O(x),(19) with a0=(σ0)2DK,ρ+σ0σ0D¯K,ρ+σ0σ˙0H+1/2+ρσ0σ02K1,11{H=1/2}.

Remark 3.8

From definition (Equation14)–(Equation15) and from the fact that Λ is quadratic in x we see that (Equation19) implies a relation between A and Λ for x0.

Remark 3.9

Implied variance expansion (Equation13) reads as follows on implied volatility (20) σBS(t,kt)|x|2Λ(x)+t2Ha(x)|x|Λ(x)2.(20) In order to implement these expansions, one can use the methods discussed in Section 5, computing numerically the rate function Λ(x) and Σ(x) using FZ expansion, and then computing a(x) using KL. However, this last step can be computationally expensive, since a large number of basis functions are needed for the KL decomposition to be accurate, for H close to 0. As an alternative, one can use approximation (21) σBS(t,kt)Σ(x)+t2Ha02σ0=|x|2Λ(x)+t2Ha02σ0,(21) for implied volatility, which follows from implied variance expansion (Equation13) and (Equation19). If the rate function cannot be computed, we can use (Equation18) to expand the implied volatility as (22) σBS(t,kt)Σ(0)+Σ(0)x+Σ(0)2x2+t2Ha02σ0.(22) In particular, we get the following explicit expansion for the ATM term structure: (23) σBS(t,0)=σ0+t2Ha02σ0+o(t2H).(23)

Remark 3.10

The term structure of implied volatility

From the expansion of the ATM term structure (Equation23) we also see, in the short end, that σBS2(t,0) is increasing in t if a0>0 and decreasing if a0<0. This may be compared with a large body of literature concerning monotonicity properties of the term structure of implied volatility, see e.g. (Camara et al. Citation2011, Guo et al. Citation2014, Krylova et al. Citation2009, Vasquez Citation2017).

Corollary 3.11

Skew and curvature at the large deviation regime

Let kt=xt1/2H, for xR{0}. Then, if H<1/2, for t0 (24) σBS(t,kt)σBS(t,kt)2ktΣ(x)Σ(x)2xtH1/2.(24) (25) σBS(t,kt)+σBS(t,kt)2σBS(t,0)kt2Σ(x)+Σ(x)2Σ(0)x2t2H1.(25)

Remark 3.12

The quantities in the rhs of the equivalences converge as x0 to Σ(0),Σ(0) given in Theorem 3.7. The quantities in the lhs of the equivalences are finite difference approximations of ATM implied volatility skew kσBS(t,0) and curvature kkσBS(t,0). Such finite differences are relevant because only a finite number of prices are observable on real markets. They give skew and curvature at the large deviation regime, a result that complements (Fukasawa Citation2017, El Euch et al. Citation2019) (skew and curvature at central limit regime), Bayer et al. (Citation2019) (skew at moderate deviation regime), Forde et al. (Citation2020) (skew and curvature at large deviations regime for rough Heston), Alòs and León (Citation2017) (true skew and curvature).

From these formulas, we also infer the sign of implied skew and of implied curvature (convexity). Indeed, if σ0,σ00, it is clear that sgn(Σ(0))=sgn(ρ) and that (26) {Σ(0)=0 iff ρ2=(K¯1)2,16K1,12(K1)2,12K1,K¯1σ0σ0(σ0)2(K1)2,1,Σ(0)<0 iff ρ2>(K¯1)2,16K1,12(K1)2,12K1,K¯1σ0σ0(σ0)2(K1)2,1>0,Σ(0)>0 otherwise.(26)

Theorem 3.13

Moderate deviations

Assume that Λ is iN times continuously differentiable. Let H(0,1/2), β>0 and nN such that β(2Hn+1,2Hn]. Set kt=xt1/2H+β. Then c(t,kt)exp(i=2nΛ(i)(0)i!xitiβ2H)t1/2+2H2βσ03x22π. Moreover (27) σBS2(t,kt)=j=0n2(1)j2jσ02(j+1)(i=3nΛ(i)(0)i!xi2t(i2)β)j+o(t2H2β).(27)

Remark 3.14

An implied volatility expansion similar to (Equation27) was proved in Bayer et al. (Citation2019), in the case σ(t,ω)=σ(Wˆt), for β[2Hn+1,2Hn), with remainder of order max(t2H2βϵ,t(n1)β). The derivatives of the rate function were computed until Λ(0), here we also computed Λ(4)(0) (cf. Lemma 6.1). This allows us to use the second-order moderate deviation (instead of first order as in Bayer et al. Citation2019) σ(t,kt)=Σ(0)+Σ(0)xtβ+Σ(0)2x2t2β+o(t2H2β). Moreover, even if it does not show up in the asymptotics, the term structure can be incorporated as follows σ(t,kt)Σ(0)+Σ(0)xtβ+Σ(0)2x2t2β+a02σ0t2H, and this provides a sensible improvement in the implementation of such short-time result (cf. figure ).

4. A case study: the rough Bergomi model

4.1. The rough Bergomi model

Introduced in Bayer et al. (Citation2016), as a modification of the classical Bergomi model where the exponential ( Ornstein–Uhlenbeck) kernel is replaced by a power-law kernel, the rBergomi model provides great fits of empirical implied volatility surfaces with a very small number of parametres. In such model, the volatility is given by the “Wick” exponential of a Riemann–Liouville fBM (28) σ(t,ω)=σ0exp(η2Wˆtη24t2H).(28) In the most general framework (Bayer et al. Citation2016), the constant σ02 is replaced by the forward variance curve, which is a function of time observable on the market (so it plays the role of an initial condition, cf. also Remark 4.1). The specific volatility in (Equation28) did not fit in the framework of Forde and Zhang (Citation2017) and Bayer et al. (Citation2019), as in these papers the volatility is assumed to be σ(Wˆt). For this reason, in Bayer et al. (Citation2019), the following version of the rBergomi model is considered (29) σ(t,ω)=σ0exp(η2Wˆt).(29) In this work we consider (Equation2), a version of the rBergomi model with one additional parameter θR, that includes both the previous ones (for θ=0,1). The volatility function in (Equation3) is (30) σ(x,y)=σ0exp(η2xθη24y).(30) The interpretation of the parameters is the following: σ0 is the spot volatility and η represents the volatility of volatility. The parameters of the driving noise are the Hurst exponent H of Wˆ and the correlation parameter ρ between the BM W~ driving the asset and W in (Equation4). We can interpret the newly introduced θ parameter as a damping coefficient of the volatility.

Remark 4.1

Note that the forward variance curve model ξt(u)=E[σ2(Wˆu,u2H)|Ft] (and Ft filtration generated by W) induced by (Equation30) and (Equation2), is a genuine rBergomi model for any value of θ, with different values of θ corresponding to different specifications of the initial variance curve. More precisely, for fixed θ, ξ0(u)=E[σ2(Wˆu,u2H)]=σ02exp((1θ)η22u2H).

Coming now to short-time pricing, Lemma 6.1 holds for the general model in (Equation2), so that we are able to compare our asymptotics with large or moderate deviations results for the different versions of rBergomi in Bayer et al. (Citation2019), Forde and Zhang (Citation2017) and Jacquier et al. (Citation2018). However, in Corollary 3.5, Σ2(x) is not affected by the value of θ, but the term structure a(x) is.

From the volatility function (Equation30) we get σ0=σ0,σ0=σ0η2, σ0=σ0η24, σ˙0=θη24σ0, so all constants can be simplified. In particular condition (Equation26) for the convexity of the short-time smile (with σ0,η0) simplifies to {Σ(0)=0 iff ρ2=(K¯1)2,16K1,122(K1)2,12K1,K¯1,Σ(0)<0 iff ρ2>(K¯1)2,16K1,122(K1)2,12K1,K¯1>0,Σ(0)>0 otherwise. (note the dependence only on H, through K, and ρ). On calibrated parameters (for example in Bayer et al. Citation2016) we have that the condition for vanishing second derivative is almost satisfied. This means that the short-time ATM curvature is very close to 0, and indeed observed smiles are almost linear ATM.

All the constants in previous expansions depend on the kernel K. For the Riemann–Liouville kernel (EquationA4) the K-functionals involved are explicit, given in (EquationA5).

4.2. Implementation of rough Bergomi

Our goal in this section is to compare expansion (Equation20) with other known implied volatility expansions under RoughVol. We consider:

  • Implied volatility from Monte Carlo pricing, using the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2 (note that a slight modification of the implementation is necessary for θ1).

  • Our implied volatility expansion, where the term structure coefficient a(x) is computed using KL, so that we have (Equation20), or where a(x) is expanded at 0, so that we have (Equation21).

  • The FZ expansion (Equation16). In Forde and Zhang (Citation2017), Forde and Zhang show that this asymptotics holds for volatilities of type σ(t,ω)=σ(Wˆt), with no direct dependence on t, so this applies to (Equation30) for θ=0. However, as we have shown in Friz et al. (Citation2021, Section 7.3), the same large deviation behavior holds when θ0. Therefore, the FZ expansion gives the same asymptotic smile, independently of the choice of θ.

  • Expansion (Equation21), with ATM expansion of Σ as in (Equation22) (so, rate function is expanded as well). In case θ=1, one can check that this approximation is consistent with the expansion in El Euch et al. (Citation2019, Section 5), that we refer to as “EFGR expansion”. These two mathematical results ar different, since log-strikes are in our case (large deviation regime) kt=xt1/2H and in El Euch et al. (Citation2019) (central limit regime) kt=xt1/2. However, when plotting for finite k and t the approximate implied volatility, the two curves are the same.

We first use the numerical methods detailed in next Section 5 to compute Σ(x) and a(x). In figure , we display implied volatility smiles in the rBergomi model with θ=1, for varying t, where the rate function is computed using the Ritz method in Section 5.1 and the coefficient a(x) is computed using the KL decomposition from Section 5.3. For comparison, we also use approximation a(x)a0, and show (Equation21). We notice that both implementations perform well, and the use of KL decomposition gives a better approximation of the right wing. On several simulations, this improvement of KL over expansion a(x)a0 is more evident when taking θ=1, less when θ=0.

Figure 1. Implied volatility smile approximations for the rBergomi model with parameters θ=1,σ0=0.2,η=1.5,ρ=0.7,H=0.3, for expiry t=0.05,0.2. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps of length t/500. The rate function is computed using the Ritz method in Section 5.1 with N = 8 Haar basis functions, the coefficient a(x) is computed using the Karhunen–Loeve decomposition with N = 300 Haar basis functions (KL). We also compare with a(x) expanded at 0 (a(x)a(0)).

Figure 1. Implied volatility smile approximations for the rBergomi model with parameters θ=1,σ0=0.2,η=1.5,ρ=−0.7,H=0.3, for expiry t=0.05,0.2. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps of length t/500. The rate function is computed using the Ritz method in Section 5.1 with N = 8 Haar basis functions, the coefficient a(x) is computed using the Karhunen–Loeve decomposition with N = 300 Haar basis functions (KL). We also compare with a(x) expanded at 0 (a(x)≈a(0)).

Practically, implementation of the KL formula requires to approximate the infinite product (Equation38), and we observed that for smaller values of H the convergence of this product was much slower, requiring a prohibitively large number of basis functions, which is why we present these results for H = 0.3. We leave the numerically efficient implementation of the KL decomposition method for small values of H as a topic for future research. In what follows we will consider the approximation a(x)a(0), which is faster while still producing accurate smiles.

First, in figure , we show implied volatilities under model (Equation2), with realistic parameters (close to the calibrated parameter to the SPX volatility on February 4, 2010, see Bayer et al. Citation2016), varying θ from 0 to 1. We note how our approximation is general enough to be applicable for any θ, improving previous asymptotics in all cases. We also note a slight deterioration of the quality of the approximation in the right wing as θ1, that could be improved using KL to compute a(x).

Figure 2. Implied volatility smile approximation for the rBergomi model with parameters σ0=0.15,η=1.8,ρ=0.78,H=0.07, for expiry t = 0.05. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps. The rate function is computed using the Ritz method with N = 9 Fourier basis functions.

Figure 2. Implied volatility smile approximation for the rBergomi model with parameters σ0=0.15,η=1.8,ρ=−0.78,H=0.07, for expiry t = 0.05. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps. The rate function is computed using the Ritz method with N = 9 Fourier basis functions.

Then, instead of varying θ, we fix θ=0 and show in figure  the comparison with the same approximations as before, when the expiry t increases. We see how our expansion lifts the FZ expansion, improving the approximation of the Monte Carlo price. The difference between the two approximations is due to the term structure correction a0t2H. Clearly, the effect of this correction becomes more evident as t increases. On a number of numerical experiments, it is also clear that this correction becomes more and more important as H0, not surprisingly since t2H is larger, for small t, when H vanishes.

Figure 3. Implied volatility smile approximation for the rBergomi model with parameters θ=0,σ0=0.15,η=1.8,ρ=0.78,H=0.07, for expiry t = 0.01, 0.05, 0.2. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps of length t/500. The rate function is computed using the Ritz method with N = 9 Fourier basis functions.

Figure 3. Implied volatility smile approximation for the rBergomi model with parameters θ=0,σ0=0.15,η=1.8,ρ=−0.78,H=0.07, for expiry t = 0.01, 0.05, 0.2. The Monte Carlo price is computed via the hybrid scheme for rBergomi in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps of length t/500. The rate function is computed using the Ritz method with N = 9 Fourier basis functions.

Now we check how our approximations behave as time increases. To do so, in figure , we show the ATM term structure of implied volatility, comparing ATM implied volatilities computed using Monte Carlo simulations and expansion (Equation23), for rBergomi with θ=0 and θ=1. We do so for parameters as in figures and , with H = 0.3, and for a different choice of parameters with H = 0.1 and a smaller volatility of volatility η, as in Bayer et al. (Citation2019, Section 4). The value of η and H affect the quality of the approximation, which is less accurate for H very close to 0 and η>1. On the other hand, as we show in figure , for H = 0.3 and η>1 or H very close to 0 and η<1 the short-time approximation is very good. This is consistent with the considerations on the interplay of H and η in El Euch et al. (Citation2019, Page 505). We also see how the term structure is increasing in case θ=0 and decreasing in case θ=1. This is always the case: a0 in (Equation19) is always positive for θ=0, always negative for θ=1 (cf. Remark 3.10). Also note that if the coefficient σ0 were taken non-constant, the slope of the term structure would also be affected.

Figure 4. Term structure of volatility for the rBergomi model with parameters σ0=0.15,η=1.8,ρ=0.78,H=0.3 (above) and with parameters σ0=0.2557,η=0.2928,ρ=0.7571,H=0.1 (below). We plot ATM implied volatility as expiration time increases. We consider shorter expiries in the case of rougher trajectories (smaller Hurst parameter H; however, in this case we also take a smaller vol-of-vol parameter η). The Monte Carlo prices are computed via the hybrid scheme in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps.

Figure 4. Term structure of volatility for the rBergomi model with parameters σ0=0.15,η=1.8,ρ=−0.78,H=0.3 (above) and with parameters σ0=0.2557,η=0.2928,ρ=−0.7571,H=0.1 (below). We plot ATM implied volatility as expiration time increases. We consider shorter expiries in the case of rougher trajectories (smaller Hurst parameter H; however, in this case we also take a smaller vol-of-vol parameter η). The Monte Carlo prices are computed via the hybrid scheme in Bennedsen et al. (Citation2017) with κ=2, with 109 simulations and 500 time steps.

Figure 5. Moderate deviation with β=0.06 and x = 0.4 (time varying log-strike kt=xt1/2H+β) of implied volatility in rBergomi model with σ0=0.2557,η=0.2928,ρ=0.7571,H=0.1,θ=0. Simulation parameters: 108 simulation paths, 500 time steps. Time interval [0,0.1].

Figure 5. Moderate deviation with β=0.06 and x = 0.4 (time varying log-strike kt=xt1/2−H+β) of implied volatility in rBergomi model with σ0=0.2557,η=0.2928,ρ=−0.7571,H=0.1,θ=0. Simulation parameters: 108 simulation paths, 500 time steps. Time interval [0,0.1].

Finally, as in Remark 3.14, we consider moderate deviations. Figure is as in Bayer et al. (Citation2019, Figure ), the “very rough” case H = 0.1 (which was the most problematic case in Bayer et al. (Citation2019)). We are plotting, with kt=xt1/2H+β, where β=0.06, the Monte Carlo implied volatility and its approximation σ(t,kt)Σ(0)+Σ(0)xtβ+Σ(0)2x2t2β+a02σ0t2H, considering terms up to the first order moderate deviation tβ, then up to the second-order moderate deviation t2β, and finally considering also the term structure t2H. We see how the term structure term improves the moderate deviation pricing. This also explains why, in Bayer et al. (Citation2019), the moderate deviation pricing gets worse as H0, since the distance of such price from the real (Monte Carlo) one is of order t2H. We also see that using the second order moderate deviation actually does not improve much, and this follows from the fact that the curvature is almost 0 with such choice of parameters (cf. Remark 3.14). As for the term structure, the accuracy of the approximation formula based on moderate deviations gets worse as η increases, for fixed H.

Remark 4.2

As mentioned above, Monte Carlo pricing is implemented using the hybrid scheme, which introduces a bias in the volatility process, while this process could be simulated exactly. However, in extensive simulations we find that the exact simulation scheme is more unstable for very short maturities, even with a 109 trajectories and 500 time steps. This is most likely due to the singularity of the kernel at 0, which is what the hybrid scheme takes care of. On the other hand, with such a large number of paths and fine discretisation, for larger maturities the two schemes display no visible difference. Following these considerations, we used for our figures Monte Carlo prices simulated via the hybrid scheme in Bennedsen et al. (Citation2017).

Remark 4.3

In Forde and Zhang (Citation2017, Section 4.5) asymptotics for model (Equation3) with volatility driven by a Mandelbrot–Van Ness fBm (EquationA2) are implemented. Without being completely rigorous, we have applied our expansion also in this case. We computed the K-functional numerically, as in this case no explicit formulas are available. Also in this case the term a0t2H lifts the smile, which gets closer to the real (Monte Carlo) implied volatility, for small |x|, with respect to the sole FZ expansion.

5. Computing the coefficients via projections

5.1. Computing Σ(x) using the Ritz method

In order to use (Equation20), the first challenge is the computation of the rate function. A numerical approximation to Λ can be obtained as described in Gelfand and Fomin (Citation2000, Section 40), using the Ritz method, as is done in Forde and Zhang (Citation2017). Natural choices for the orthonormal basis (ONB) {ei}i1 of H1 are the Fourier basis, (31) e˙1(s)=1,e˙2n(s)=2cos(2πns),e˙2n+1(s)=2sin(2πns), for nN{0},(31) or the Haar basis, (32) e˙1(s)=1,e˙2k+l(s)=2k/2(1[2l22k+1,2l12k+1](s)1[2l12k+1,2l2k+1](s)) for k0,1l2k.(32) We consider functions hH1 with h(0)=0 so that h˙(s)=n=1Nane˙n(s), for NN fixed. Then we minimize Λ(x)=(xρG(h))22ρ¯2F(h)+h˙,h˙2, with (33) F(h)=σ2(hˆ,0),1,G(h)=σ(hˆ,0),h˙,(33) over the Fourier coefficients (an)n. This representation of the energy function is also taken from Forde and Zhang (Citation2017) (see notation in Bayer et al. Citation2019, Proposition 5.1). The minimizing value for Λ(x) is therefore our approximation for the energy and the corresponding function hˆx is the approximate most likely path for the fBm WH associated with final condition x.

5.2. A stochastic Taylor development

The following stochastic Taylor expansion is sketched in Friz et al. (Citation2021, Section 7.2) for σ(ω,t)=σ(WtH). As discussed in Section 2 and Friz et al. (Citation2021, Section 7.3), our expansions can actually be carried out in the more general setting σ(ω,t)=σ(WtH,t2H). Under such volatility dynamics, the (rescaled) log-price process is as in (Equation9). As in Friz et al. (Citation2021, Section 7.2), we can shift the dynamics via ϵˆ(W,W¯)(ϵˆW+h,ϵˆW¯+h¯), and apply Girsanov theorem in order to center Brownian fluctuations in the minimizer. Then, a stochastic Taylor expansion gives 01σ(ϵˆWˆt+hˆt,ϵˆ2t2H)d[ϵˆW~+h~]tϵϵˆ201σ2(ϵˆWˆt+hˆt,ϵˆ2t2H)dtg0+ϵˆg1(ω)+ϵˆ2g2(ω)+r3(ω), where r3(ω) is smallFootnote1 , with (34) g1=01σ(hˆsx,0)Wˆsdh~sx+01σ(hˆsx,0)dW~s,(34) (cf. Friz et al. Citation2021, Section 7.2) and (35) g2={1201σ(hˆsx,0)Wˆs2dh~sx+01σ(hˆsx,0)WˆsdW~s+01σ˙(hˆsx,0)s2Hdh~sxif H<1/2,1201σ(hˆsx,0)Wˆs2dh~sx+01σ(hˆsx,0)WˆsdW~s+01σ˙(hˆsx,0)s2Hdh~sx1201σ2(hˆsx,0)dsif H=1/2.(35) The following formula for Δ2 follows as Friz et al. Citation2021, Equation 7.5 (36) Δ2={1201σ(hˆsx,0)Vˆs2dh~sx+01σ(hˆsx,0)VˆsdV~s,+01σ˙(hˆsx,0)s2Hdh~sxif H<1/2,1201σ(hˆsx,0)Vˆs2dh~sx+01σ(hˆsx,0)VˆsdV~s+01σ˙(hˆsx,0)s2Hdh~sx1201σ2(hˆsx,0)dsif H=1/2.(36) where we write vt=E[Wtg1]/E[g12],v¯t=E[W¯tg1]/E[g12],v~t=ρvt+ρ¯v¯t, vˆ=Kv˙ and V~t=W~tv~tg1,Vˆt=Wˆtg1vˆt.

5.3. Computing a(x) using Karhunen–Loeve decomposition

Assume we are given hx computed by the Ritz method. Note then that h¯x is obtained from hx via the following formula (37) h¯x˙(s)=xρG(hx)ρ¯F(hx)σ(hˆx(s)),(37) with G(h),F(h) as in (Equation33), as can be seen by optimizing over h¯ for fixed h in the definition (Equation10) of the rate function. Then we assume a Karhunen–Loeve (KL) decomposition of (W,W¯): W=iγiei,W¯=iγi¯ei, where {ei}i is the ONB in (Equation31) or (Equation32) and γi,γ¯i are i.i.d. standard Gaussians. This implies Wˆ=iγieˆi, with eiˆ(t)=(Ke˙i)t. This yields g1=igiγi+gi¯γi¯, where gi=01σ(hˆsx,0)eiˆ(s)dh~x(s)+ρ01σ(hˆsx,0)dei(s),gi¯=ρ¯01σ(hˆsx,0)dei(s). In particular σx2=igi2+gi¯2. Note then that v(t)=igiei(t)/σx2,v¯(t)=igi¯ei(t)/σx2,vˆ(t)=igieˆi(t)/σx2, We then can write all the terms in Δ2 as follows. We denote αij=01σ(hˆsx,0)eiˆ(s)ejˆ(s)dh~sx,βij=01σ(hˆsx,0)eiˆdej, δij=1i=j and g~i=ρgi+ρ¯g¯i. Now, expanding (Equation36) with some long but standard computations we get to Δ2=ij(γiγjδij)η0;ij+γiγj¯η1;ij+(γi¯γj¯δij)η2;ij+C=:Δ2(2)+C, where η0;ij=12αij1σx2gikgkαjk+12σx4gigj(klgkglαkl)+ρβij1σx2gikg~kβjkρσx2gikgkβkj+1σx4gigj(k,lgkg~lβkl)η1;ij=ρ¯βij1σx2g¯jkgkαik+12σx4gig¯j(k,lgkglαk,l)1σx2g¯jkg~kβikρ¯σx2gikgkβkjρσx2g¯jkgkβki+1σx4gig¯j(klgkg~lβkl)η2;ij=12σx4gi¯gj¯(k,lgkglαk,l)ρ¯σx2gi¯kgkβkj+1σx4gi¯gj¯(k,lgkg~lβk,l) and C=01σ˙(hˆsx,0)s2Hdh~sx+12iαii12σx2i,kgigkαik1σx2i,kgig~kβik. Recall that one has A(x)=eΛ(x)CEexp(Λ(x)Δ2(2)), where Λ(x)=sgn(x)2Λ(x)σx2, and since Δ2(2) is an element of the homogeneous Wiener chaos of order 2, the expectation above can be computed as the Carleman–Fredholm determinant det2(I2M)1/2, where M is the symmetric matrix M=Λ(x)(12(η0+η0t)η1(η1)t12(η2+η2t)). Namely one has (38) Eexp(Λ(x)Δ2(2))=Πk0(12λk)1/2eλk(38) where (λk)k are the eigenvalues of M (note that the fact that all λk<1/2 comes from the non-degeneracy assumption). This formula is a simple integral computation if M is diagonal, and the general case follows by diagonalization, cf e.g. Inahama (Citation2013, Remark 5.5) or Janson (Citation1997, p.78).

Of course, in practice we consider approximations WN, W¯N obtained by truncating the sums to only keep indices iN, where N is fixed, so that all the sums above are then replaced by finite sums. One also needs to compute numerically the integrals appearing in the definition of the coefficients g, α, β. We have found the Haar basis to be more convenient than the Fourier basis for this purpose since the eˆi's have explicit expressions in that case.

6. Proofs

6.1. Energy expansion

Lemma 6.1

Fourth order energy expansion

Consider a stochastic volatility model following dynamics (Equation3) and the associated energy function in (Equation10). Let Λ(x) be the energy function in (Equation10). Then Λ(x)=Λ(0)2x2+Λ(0)3!x3+Λ(4)(0)4!x4+O(x5) where (39) Λ(0)=1σ02,Λ(0)=6ρσ0σ04K1,1,(39) and Λ(4)(0)=12(σ0)2σ06{9ρ2K1,12ρ2(K1)2,1(K¯1)2,12ρ2K1,K¯1}12σ0σ05ρ2(K1)2,1.

Remark 6.2

In this lemma we expand the rate function Λ(x), which has been studied first in Forde and Zhang (Citation2017). The second- and third-order terms in (Equation39) have been computed in Bayer et al. (Citation2019, Theorem 3.4). In both these papers, the volatility function is supposed to be σ(WtH), but adding the dependence σ(WtH,t2H) does not change the large deviations behavior, meaning that the rate function is the same as the one of the model given by σ(WtH,0).

Proof.

We have the following development for the minimizer hx in (Equation10), for x0: (40) htx=αtx+βtx22+γtx36+O(x4),(40) with αt=ρσ0t,βt=2σ0σ03[ρ2K1,1[0,t]+K1[0,t],13ρ2tK1,1], where α,β have been also computed in Bayer et al. (Citation2019). We make here the ansatz that the expansion goes on one more order with γ, that we do not actually need to compute. The existence of such γ follows from the smoothness of σ(,) (cf. Friz et al. Citation2021 and Bayer et al. Citation2019, Section 5.2). We can compute, using K(K1),1=K1,K¯1 and K(K¯1),1=(K¯1)2,1, Kβ˙=2σ0σ03[ρ2K(K1)+K(K¯1)3ρ2K1,1K1],Kβ˙,1=2σ0σ03[ρ2K1,K¯1+(K¯1)2,13ρ2K1,12],K1,β˙=2σ0σ03[ρ2(K1)2,1+K1,K¯13ρ2K1,12]. We also have (41) σ(hˆsx,0)=σ0+xσ0σ0ρK1(s)+(σ0σ02ρ2(K1)2(s)+σ0Kβ˙(s))x22+O(x3).(41) We use now (Equation33) and compute (42) F(hx)=σ02+x2ρσ0K1,1+x2{((σ0σ0)2+σ0σ0)ρ2(K1)2,1+σ0σ0Kβ˙,1((σ0σ0)2+σ0σ0)}+O(x3),G(hx)=ρx+x2(σ0σ02ρ2K1,1+σ02β1)+x3(σ06γ1+σ02σ03ρ3(K1)2,1+ρ(σ0)2σ04[(ρ2+1)K1,K¯1+ρ2(K1)2,1+(K¯1)2,16ρ2K1,12])+O(x3),(42) from which we get xρG(hx)=(1ρ2)xx2ρσ0σ02(1ρ2)K1,1x3ρ(σ06γ1+σ02σ03ρ3(K1)2,1+ρ(σ0)2σ04[(ρ2+1)K1,K¯1+ρ2(K1)2,1+(K¯1)2,16ρ2K1,12]σ06)+O(x3),1F(hx)=1σ02x2ρσ0σ04K1,1x2{σ0σ05ρ2(K1)2,1+(σ0σ03)2(ρ2(K1)2,1+2(K¯1)2,1+2ρ2K1,K¯110ρ2K1,12)σ0σ05}+O(x3),(xρG(hx))21ρ2=(1ρ2)x2x32ρσ0σ02(1ρ2)K1,1+x4[ρ2(σ0)2σ04(1+11ρ2)K1,12σ0σ03ρ4(K1)2,1ρσ03γ12ρ2(σ0)2σ04×[(ρ2+1)K1,K¯1+ρ2(K1)2,1+(K¯1)2,1])(σ0)2σ04]+O(x5),(xρG(hx))22(1ρ2)F(hx)=()x2+()x3x4ρ6σ0γ1+x4{ρ22σ0σ05(K1)2,12ρ2(σ0)2σ06K1,K¯1(σ0)2σ06ρ4+ρ22(K1)2,1(σ0)2σ06(K¯1)2,1+(σ0)2σ0632ρ2(5ρ2)K1,12}+O(x5). We also have, from (Equation40) h˙x,h˙x=+x4(ρ3σ0γ1+(σ0σ03)2[ρ4(K1)2,1+(K¯1)2,1+2ρ2K1,K¯1+3ρ4K1,126ρ2K1,1K¯1,1](σ0σ03)2)+O(x5). Now we write, from Bayer et al. (Citation2019, Proposition 5.1), Λ(x)=(xρG(hx))22ρ¯2F(hx)+h˙x,h˙x2 and use the expansions above for the two summands. The fourth-order expansion of Λ(x) follows.

6.2. Proof of Lemma 3.4

Let us take x0.

STEP 1: We first need to expand h¯tx in (Equation10), for small x (an expansion of hx was computed in Bayer et al. Citation2019). We write (43) Φ1(W,W¯)=X1(43) for the Itô map associated with the RoughVol model (Equation8). Computing the Frechet derivative of Φ1 with respect to the second component at h=(h,h¯) in the direction f we get (cf. (Equation34)) (44) DΦ1(h),(0,f)=D2Φ1(h),f=ddδΦ1(h,h¯+δf)=ρ¯01σ(hˆ,0)df.(44) From the first order optimality condition (Friz et al. Citation2021, Appendix B), we get that for hx minimizer and any f in the Cameron–Martin space H1, htx,ftH1=Λ(x)DΦ1,f. Let f be the second component of f. Using (Equation44) we get 01h¯˙txf˙tdt=h¯tx,ftH1=Λ(x)D2Φ1,f=ρ¯Λ(x)01σ(hˆtx,0)f˙tdt. Now, from (Equation39) we derive that, for x0, (45) Λ(x)=xσ023x2ρσ0σ04K1,1+O(x3).(45) We get h¯tx=xρ¯σ0t+O(x2). We also have (46) htx=xρσ0t+O(x2),h~tx=ρhtx+ρ¯h¯tx=xtσ0+O(x2),hˆtx=(Kh˙x)t=xρσ0K1(t),(46) and σ(hˆx,0)=σ0+xρσ0σ0K1+O(x2),σ(hˆx,0)=σ0+xρσ0σ0K1+O(x2), σ˙(hˆx,0)=σ˙0+O(x). STEP 2: We recall here, from Friz et al. (Citation2021), the definition of some quantities needed to compute A(x). Let g1 be as in (Equation34) and let us write σx2=Var(g1) for its variance. We recall, again from Friz et al. (Citation2021, Equation (6.3)), σx2=2Λ(x)/Λ(x)2, from which we get (47) σx2=σ02+4ρσ0K1,1x+O(x2).(47) From (Equation34) we define and compute (48) vt=E[Wtg1]E[g12]=1σx2(ρ0tσ(hˆsx,0)ds+01σ(hˆsx,0)K1[0,t](s)dh~sx),v¯t=E[W¯tg1]E[g12]=1σx2ρ¯0tσ(hˆsx,0)ds.(48) (Note that v,v¯ are in the Cameron–Martin space). From (Equation11) we have that A(x) in Theorem 3.2 is A(x)=E[exp(Λ(x)Δ2)], where Δ2 is given in (Equation36).

STEP 3: We can expand now such quantity, for x0 and we get (49) A(x)=1+xΛ(0)E[Δ20]+x2(Λ(0)E[Δ20]+Λ(0)2E[(Δ20)2]2+Λ(0)E[x|x=0Δ2]Λ(0)E[Δ20]+Λ(0)2E[(Δ20)2]2)+O(x3),(49) where Δ20 denotes Δ2|x=0. The statement of the theorem follows from the computation of the quantities in (Equation49).

STEP 4: We compute v˙t=1σx2(ρσ(hˆtx,0)+t1σ(hˆsx,0)K(s,t)dh~sx), and we obtain, also using (Equation48), (50) σx2vt=ρσ0t+xσ0σ0(ρ2K1,1[0,t]+K1[0,t],1)+O(x2),σx2v~t=σ0t+xσ0σ0ρ(K1,1[0,t]+K1[0,t],1)+O(x2),σx2vˆt=ρσ0K1(t)+xσ0σ0(ρ2K(K1)(t)+K(K¯1)(t))+O(x2),(50) where we have used K1[0,t],1=0tK¯1(u)du. We have σx4vˆtdv~t=ρσ02K1(t)dt+xσ0(ρ2K(K1)(t)+K(K¯1)(t)+ρ2K1(t)(K1(t)+K¯1(t)))dt. Putting together the previous expressions and using K(K1),1=K1,K¯1 and K(K¯1),1=(K¯1)2,1 we get (51) σx401vˆtdv~t=ρσ02K1,1+xσ0(ρ2(K1)2,1+(K¯1)2,1+2ρ2K1,K¯1)+O(x2),σx401K1(t)vˆtdv~t=ρσ02(K1)2,1+O(x),σx401vˆt2dt=ρ2σ02(K1)2,1+O(x).(51) This implies, together with (Equation47), (52) x(σx201vˆtdv~t)=σ0σ02(2ρ2K1,K¯1+ρ2(K1)2,1+(K¯1)2,14ρ2K1,12).(52) We can now compute E01(VˆsdV~s)=E01WˆsdW~s+E[g12]01vˆsdv~sE[g1(01Wˆsdv~s+vˆsdW~s)], where E[g101Wˆsdv~s]=010sK(s,u)dE[g1Wu]dv~s=σx201vˆsdv~s,E[g101vˆsdW~s]=01vˆsdE[g1W~s]=σx201vˆsdv~s, so that E01(VˆsdV~s)=σx201vˆsdv~s. We also compute 01E[Vˆs2]ds=K21,1σx201vˆs2ds,01K1(s)E[VˆsdV~s]=σx201K1(s)vˆsdv~s, and all these quantities can be expanded in x using (Equation51). Now we use (Equation36) to write, in the case H<1/2 (53) EΔ20=σ0E01Vˆs0dV~s0=ρσ0K1,1.(53) Moreover, using (Equation46), x|x=001σ˙(hˆsx,0)s2Hdh~sx=σ˙0(2H+1)σ0. Now, also using (Equation36) and (Equation51) we get xEΔ2|x=0=σ02σ001E[(Vˆs0)2]ds+ρσ0σ001K1(s)E[Vˆs0dV~s0]+σ001xE[VˆsdV~s]|x=0+σ˙0(2H+1)σ0=σ0σ0(K21,1232ρ2(K1)2,1)(σ0σ0)2(2ρ2K1,K¯1+ρ2(K1)2,1+(K¯1)2,14ρ2K1,12)+σ˙0(2H+1)σ0. STEP 5: We need now to compute E[(Δ20)2]=(σ0)2E(01Vˆs0dV~s0)2, where (using definitions and (Equation50)) Vˆs0=WˆsρK1(s)W~1 and dV~s0=dW~sW~1ds. We can rewrite 01Vˆs0dsW~1=ρ¯01K¯1(u)W~udBu+01(0uρ(K¯1(u)2K1,1)W~u+0uK¯1(s)dWs)dW~u. and, differentiating the product 01K1(u)dW~u,W~1 01Vˆs0dV~s0=ρ¯01K¯1(u)W~udBu+01(Vˆu00uK¯1(s)dWs+ρ(2K1,1K¯1(u))W~u0u)dW~u=ρ¯01K¯1(u)W~udBuρ01K1(u)du+01(Wˆu0uK¯1(s)dWs+ρ(2K1,1K¯1(u)K1(u))W~uρ0uK1(s)dW~s)dW~u with W~ independent of B. Therefore, by Itô isometry, E[(01Vˆs0dV~s0)2]=ρ¯201K¯1(u)2udu+ρ2(01K1(u)du)2+E01(Wˆu0uK¯1(s)dWs+ρ(2K1,1K¯1(u)K1(u))W~uρ0uK1(s)dW~s)2du. We can apply again Itô isometry to compute the last expectations, and E[(01Vˆs0dV~s0)2]=ρ2010u(K(u,s)K¯1(s)+2K1,1K¯1(u)K1(u)K1(s))2dsdu+ρ2(01K1(u)du)2+ρ¯2010u(K(u,s)K¯1(s))2dsdu+ρ¯201K¯1(u)2udu. At this point it is a (long) calculus excercise (noting K¯1,1=K1,1) to show that (54) E[(01Vˆs0dV~s0)2]=ρ2(3K1,12(K1)2,12K1,K¯1)+K21,1(K¯1)2,1(54)

STEP 6: Substituting in (Equation49) we get A(x)=1xρσ0σ02K1,1+x2{(σ0)2σ04(3ρ2K1,12+12E[(01Vˆs0dV~s0)2])+σ0σ03(K21,1232ρ2(K1)2,1)(σ0)2σ04(2ρ2K1,K¯1+ρ2(K1)2,1+(K¯1)2,14ρ2K1,12)+σ˙0(2H+1)σ03}+O(x3)=1xρσ0σ02K1,1+x2{(σ0)2σ04(K21,1232(K¯1)2,1+ρ2(172K1,1232(K1)2,13K1,K¯1))+σ0σ03(K21,1232ρ2(K1)2,1)+σ˙0(2H+1)σ03}+O(x3)

and we get Theorem Equation12.

STEP 7: When H=1/2, Δ2 in (Equation36) has an additional summand. Let us write Δ~2=1201σ(hˆsx,0)Vˆs2dh~sx+01σ(hˆsx,0)VˆsdV~s+01σ˙(hˆsx,0)dh~sx, so that Δ~2 has the same expression as Δ2 in the rough case H<1/2. For H=1/2 we can write (55) Δ2=Δ~21201σ2(hˆsx,0)ds,(55) so that (56) Δ20=Δ~20σ022(56) and, using (Equation41) (57) x|x=0Δ2=x|x=0Δ~2σ0ρK1,1(57) Now, A(x) in Theorem 3.2 is A(x)=exE[exp(Λ(x)Δ2)] with Δ2 as above. Expanding in x we find (58) A(x)=1+x(Λ(0)E[Δ20]+1)+x2(Λ(0)E[Δ20]+E[(Λ(0)Δ20+1)2]2+Λ(0)E[x|x=0Δ2]Λ(0)E[Δ20]+E[(Λ(0)Δ20+1)2]2)+O(x3)=1+xΛ(0)E[Δ~20]+x2(Λ(0)E[Δ~20]+Λ(0)2E[(Δ~20)2]2+Λ(0)E[x|x=0Δ~2]Λ(0)E[Δ~20]+Λ(0)2E[(Δ~20)2]2)+x2+x28+O(x3)(58) (we have used (Equation39) and (Equation53)).

6.3. Proof of Theorem 3.7

A Taylor expansion gives Σ(x)=x2Λ(x)=1Λ(0)(1Λ(0)6Λ(0)x+Λ(0)2Λ(0)Λ(4)(0)24Λ(0)2x2)+O(x3). The explicit expressions for the three terms now follow from Lemma 6.1. Let us compute a(x). The rate function is quadratic and Λ(0)=Λ(0)=0. Then, using Taylor developments of Λ and x11+x we get 2Λ(x)xΛ(x)=1x6Λ(0)Λ(0)+x212{(Λ(0)Λ(0))2Λ(4)(0)Λ(0)}+O(x3)=1+Λ(0)(xΣ(0)+x2Σ(0))+O(x3) From Lemma 6.1, 2Λ(x)xΛ(x)=1+xρσ0K1,1σ02+x2Λ(0)Σ(0)+O(x3) and, with A(x) given in Lemma Equation12, we have when H<1/2 2A(x)Λ(x)xΛ(x)=1+x2a¯0+O(x3) with a¯0=(σ0)2σ04CK,ρ+σ0σ03C¯K,ρ+σ˙0(2H+1)σ03+Λ(0)v(0)ρ2(σ0)2K1,12σ04=(σ0)2σ04DK,ρ2+σ0σ03D¯K,ρ2+σ˙0(2H+1)σ03, with DK,ρ,D¯K,ρ, defined in (Equation17). Now, as a consequence of Lemma 6.1, we have x22Λ(x)22Λ(0)2x2=2σ04x2 and the expansion of a(x) follows. When H=1/2, 2A(x)Λ(x)xΛ(x)exp(x/2)=1+x2a¯0+O(x3) with a¯0=(σ0)2σ04DK,ρ2+σ0σ03D¯K,ρ2+σ˙0(2H+1)σ03+ρσ02σ02K1,1. We conclude as in the case H<1/2.

6.4. Proof of Theorem 3.13

The call asymptotics is a corollary of Theorem 3.2, taking into consideration that Λ(xt)=i=2nΛ(i)(0)i!xitiβ+O(t(n+1)β) and that O(t(n+1)β2H)0 under β(2Hn+1,2Hn]. Recall Λ(0)=σ02 and the first statement follows.

Let us write α=2H2β, δ=1/2+2H2β, γ=1/2H+β and M(t,x)=i=3nΛ(i)(0)i!xitiβ2H. We intend to apply (Gao and Lee Citation2014, Corollary 7.1, Equation (7.2)), where G(k,u) denotes 2(u+ku) and V denotes tσBS. To do so, we notice that (59) G2(k,u)=k22u+o(k2u2)(59) when k0 and u. In the notation of Gao and Lee (Citation2014), we have Lt=logc(t,kt), and G will be computed for u=Lt32logLt+log(kt4π), so let us compute Lt32logLt+log(kt4π)=x22σ02tα+M(t,x)logσ03x22πδlogt32logLt+log(kt4π)+o(1) and take care of the logarithmic terms in t. For t0, δlogt32logLt+log(kt4π)=(δ+32α+γ)logt32log(x22σ02)+log(x4π)+o(1)=32log(x22σ02)+log(x4π)+o(1) So (60) Lt32logLt+log(x4π)=x22σ02tα+M(t,x)+o(1)(60) Equations (Equation59) and (Equation60) tell us that (61) 1tG2(kt,Lt32logLt+log(kt4π))=σ0211+2σ02M(t,x)x2tα+o(tα)+o(tα)(61) The proof now boils down to writing the development of this factor using the Taylor developement of 11+u, with u=2σ02M(t,x)tα/x2+o(tα). We have, for jN, uj=(2σ02M(t,x)x2tα)j+o(tα) using M(t,x)p1tjα=o(tα) for jp1. Also notice un1=O((M(t,x)tα)n1)=o(tα) because β(2Hn+1,2Hn]. We have (62) 11+u=j=0n2(1)juj+O(un1)=j=0n2(1)j(2σ02M(t,x)x2tα)j+o(tα)(62) So from (Equation61) and (Equation62) (63) 1tG2(kt,Lt32logLt+log(kt4π))=j=0n2(1)j2jσ02(j+1)(M(t,x)x2tα)j+o(tα)(63)

We apply now (Gao and Lee Citation2014, Corollary 7.1, Equation (7.2)): |1tG2(kt,Lt32logLt+log(kt4π))σBS2(kt)|=o(kt2tLt2)=o(tα) and obtain expansion (Equation27).

Acknowledgments

We are grateful to C. Bayer and M. Fukasawa for discussion and to F. Bourgey and M. Pakkanen for the Python and R code for simulating the rough Bergomi model. We thank an anonymous reviewer for several remarks that helped us to improve the paper.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been corrected with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

PKF and PP gratefully acknowledge financial support from European Research Council (ERC) Grant CoG-683164 and German science foundation (DFG) via the cluster of excellence MATH+, project AA4-2. PG acknowledges financial support from the French ANR via the project ANR-16-CE40- 0020-01.

Notes

1 The precise control of this remainder is detailed in Friz et al. (Citation2021) and requires the sophisticated mathematical framework of regularity structures, that we do not intend to introduce in this paper. The interested reader is referred to Bayer et al. (Citation2020) and Friz et al. (Citation2021).

References

  • Ait-Sahalia, Y., Li, C. and Li, C.X., Implied stochastic volatility models. Rev. Financ. Stud., 2020, 34, 394–450.
  • Alòs, E. and León, J., On the curvature of the smile in stochastic volatility models. SIAM J. Financ. Math., 2017, 8(1), 373–399.
  • Alòs, E., León, J. and Vives, J., On the short-time behavior of the implied volatility for jump-diffusion models with stochastic volatility. Finance Stoch., 2007, 11(4), 571–589.
  • Bayer, C., Friz, P. and Gatheral, J., Pricing under rough volatility. Q. Finance, 2016, 16(6), 887–904.
  • Bayer, C., Friz, P.K., Gassiat, P., Martin, J. and Stemper, B., A regularity structure for rough volatility. Math. Finance, 2020, 30(3), 782–832.
  • Bayer, C., Friz, P.K., Gulisashvili, A., Horvath, B. and Stemper, B., Short-time near-the-money skew in rough fractional volatility models. Quant. Finance, 2019, 19(5), 779–798.
  • Bayer, C., Hammouda, C.B. and Tempone, R., Hierarchical adaptive sparse grids and quasi-Monte Carlo for option pricing under the rough Bergomi model. Quant. Finance, 2020, 20(9), 1457–1473.
  • Bayer, C., Harang, F.A. and Pigato, P., Log-modulated rough stochastic volatility models. SIAM J. Financ. Math., 2021, 12(3), 1257–1284.
  • Bayer, C., Horvath, B., Muguruza, A., Stemper, B. and Tomas, M., On deep calibration of (rough) stochastic volatility models. arXiv preprint arXiv:1908.08806, 2019,
  • Bennedsen, M., Lunde, A. and Pakkanen, M.S., Hybrid scheme for Brownian semistationary processes. Finance Stoch., 2017, 21(4), 931–965.
  • Bennedsen, M., Lunde, A. and Pakkanen, M.S., Decoupling the short- and long-term behavior of stochastic volatility. J. Financ. Econ., 2021, nbaa049.
  • Berestycki, H., Busca, J. and Florent, I., Computing the implied volatility in stochastic volatility models. Commun. Pure. Appl. Math., 2004, 57(10), 1352–1373.
  • Camara, A., Krehbiel, T. and Lib, W., Expected returns, risk premia, and volatility surfaces implicit in option market prices. J. Banking Finance, 2011, 35(1), 215–230.
  • El Euch, O., Fukasawa, M., Gatheral, J. and Rosenbaum, M., Short-term at-the-money asymptotics under stochastic volatility models. SIAM J. Financ. Math., 2019, 10(2), 491–511.
  • El Euch, O., Fukasawa, M. and Rosenbaum, M., The microstructural foundations of leverage effect and rough volatility. Finance Stoch., 2018, 22(2), 241–280.
  • El Euch, O. and Rosenbaum, M., The characteristic function of rough heston models. Math. Finance, 2019, 29(1), 3–38.
  • Forde, M., Gerhold, S. and Smith, B., Small-time, large-time and H→0 asymptotics for the rough heston model. Math. Finance, 2020.
  • Forde, M., Jacquier, A. and Lee, R., The small-time smile and term structure of implied volatility under the Heston model. SIAM J. Financ. Math., 2012, 3(1), 690–708.
  • Forde, M. and Zhang, H., Asymptotics for rough stochastic volatility models. SIAM J. Financ. Math., 2017, 8(1), 114–145.
  • Friz, P.K., Gassiat, P. and Pigato, P., Precise asymptotics: Robust stochastic volatility models. Ann. Appl. Probab., 2021, 31(2), 896–940.
  • Friz, P.K., Gerhold, S. and Pinter, A., Option pricing in the moderate deviations regime. Math. Finance, 2017, 28(3), 962–988.
  • Fukasawa, M., Asymptotic analysis for stochastic volatility: Martingale expansion. Finance Stoch., 2011, 15(4), 635–654.
  • Fukasawa, M., Short-time at-the-money skew and rough fractional volatility. Quant. Finance, 2017, 17(2), 189–198.
  • Fukasawa, M., Volatility has to be rough. Quant. Finance, 2021, 21(1), 1–8.
  • Fukasawa, M., Takabatake, T. and Westphal, R., Is volatility rough? arXiv preprint arXiv:1905.04852, 2019.
  • Gao, K. and Lee, R., Asymptotics of implied volatility to arbitrary order. Finance Stoch., 2014, 18(2), 349–392.
  • Gatheral, J., Jaisson, T. and Rosenbaum, M., Volatility is rough. Quant. Finance, 2018, 18, 1–17.
  • Gelfand, I.M. and Fomin, S.V., Calculus of Variations, 2000 (Dover Publications: Englewood Cliffs, NJ).
  • Goudenège, L., Molent, A. and Zanette, A., Machine learning for pricing American options in high-dimensional Markovian and non-Markovian models. Quant. Finance, 2020, 20(4), 573–591.
  • Gulisashvili, A., Gaussian stochastic volatility models: Scaling regimes, large deviations, and moment explosions. Stoch. Proc. Appl., 2020, 130(6), 3648–3686.
  • Gulisashvili, A., Time-inhomogeneous Gaussian stochastic volatility models: Large deviations and super roughness. arXiv preprint arXiv:2002.05143, 2020.
  • Guo, B., Han, Q. and Zhao, B., The Nelson-Siegel model of the term structure of option implied volatility and volatility components. J. Futures Mark., 2014, 34(8), 788–806.
  • Inahama, Y., Laplace approximation for rough differential equation driven by fractional Brownian motion. Ann. Probab., 2013, 41(1), 170–205.
  • Jacquier, A., Pakkanen, M.S. and Stone, H., Pathwise large deviations for the rough Bergomi model. J. Appl. Probab., 2018, 55(4), 1078–1092.
  • Jacquier, A. and Pannier, A., Large and moderate deviations for stochastic Volterra systems. arXiv preprint arXiv:2004.10571, 2020.
  • Janson, S., Gaussian Hilbert Spaces, Vol. 129, 1997 (Cambridge University Press).
  • Jost, C., A note on ergodic transformations of self-similar Volterra Gaussian processes. Electron. Commun. Probab., 2007, 12, 259–266.
  • Krylova, E., Nikkinen, J. and Vähämaa, S., Cross-dynamics of volatility term structures implied by foreign exchange options. J. Econ. Bus., Sept. 2009, 61(5), 355–375.
  • Lee, R.W., Implied volatility: Statics, dynamics, and probabilistic interpretation. In Recent Advances in Applied Probability, pp. 241–268, 2005 (Springer: New York).
  • Mandelbrot, B. and Van Ness, J.W., Fractional Brownian motions, fractional noises and applications. SIAM Rev., 1968, 10, 422–437.
  • McCrickerd, R. and Pakkanen, M.S., Turbocharging Monte Carlo pricing for the rough bergomi model. Quant. Finance, 2018, 18(11), 1877–1886.
  • Medvedev, A. and Scaillet, O., A simple calibration procedure of stochastic volatility models with jumps by short term asymptotics. Research Paper No. 93, September 2003, FAME – International Center for Financial Asset Management and Engineering, 2003. Available at SSRN 477441, 2003.
  • Medvedev, A. and Scaillet, O., Approximation and calibration of short-term implied volatilities under jump-diffusion stochastic volatility. Rev. Financ. Studies, 2007, 20(2), 427–459.
  • Nualart, D., The Malliavin Calculus and Related Topics, Vol. 1995, 2006 (Springer).
  • Osajima, Y., General asymptotics of Wiener functionals and application to implied volatilities. In Large Deviations and Asymptotic Methods in Finance, pp. 137–173, 2015 (Springer).
  • Pham, H., Large deviations in finance. Third SMAI European Summer School in Financial Mathematics, 2010.
  • Vasquez, A., Equity volatility term structures and the cross section of option returns. J. Financ. Quant. Anal., 2017, 52(06), 2727–2754.

Appendix: Fractional Brownian motion

The fBM is a “rough” continuous-time Gaussian process in that, depending on a parameter H(0,1), its trajectories are locally Hölder continuous of any order strictly less than H. Unlike classical BM, the increments of fBm are not independent if H1/2. The fBM was introduced for the first time by Mandelbrot and Van Ness in Mandelbrot and Van Ness (Citation1968) as the following stochastic integral, for t0: ZtH=cH[t(ts)H1/2dZs0(s)H1/2dZs], where Z is a BM and cH=(0[(1+s)1/2Hs1/2H]2ds+12H)1/2. Such process is Gaussian with covariance (A1) E[ZtHZsH]=12(|t|2H+|s|2H|ts|2H).(A1) It can also be represented as a Volterra integral on the interval [0,t]: (A2) ZtH=0tKH(s,t)dBs,(A2) with KH as in Nualart (Citation2006) or Forde and Zhang (Citation2017, Section 3.1)). One can consider the following variant of fBM, known as Riemann–Liouville process (Mandelbrot and Van Ness Citation1968), introduced in 1953 by Lévy. This process is also represented as Volterra integral as (A3) BˆtH=0tK(t,s)dBs,(A3) with a simpler kernel (A4) K(t,s)=2H(ts)H1/2,forH(0,1).(A4) It is still self-similar, but stationarity of increments does not hold. Moreover, the covariance structure is more complicated than (EquationA1). It can be expressed using hypergeometric functions (see Bayer et al. Citation2019, Lemma 4.1). The K-functionals that we find in our expansion can be computed in this case as (A5) K1,1=2H(H+1/2)(H+3/2)K21,1=12H+1(K1)2,1=(K¯1)2,1=H(H+1)(H+1/2)2K1,K¯1=2H(H+1/2)2β(H+3/2,H+3/2)(A5) where β is the beta function. In the case, K1 the fBM driving the volatility is actually a BM and we are back to the classical setting of a diffusive Markovian volatility. In this case, our expansions can be compared e.g. to Medvedev and Scaillet (Citation2003Citation2007).