2,578
Views
45
CrossRef citations to date
0
Altmetric
Research Papers

Short-time near-the-money skew in rough fractional volatility models

, , , &
Pages 779-798 | Received 22 Mar 2017, Accepted 07 Sep 2018, Published online: 13 Nov 2018

Abstract

We consider rough stochastic volatility models where the driving noise of volatility has fractional scaling, in the ‘rough’ regime of Hurst parameter H<1/2. This regime recently attracted a lot of attention both from the statistical and option pricing point of view. With focus on the latter, we sharpen the large deviation results of Forde-Zhang [Asymptotics for rough stochastic volatility models. SIAM J. Financ. Math., 2017, 8(1), 114–145] in a way that allows us to zoom-in around the money while maintaining full analytical tractability. More precisely, this amounts to proving higher order moderate deviation estimates, only recently introduced in the option pricing context. This in turn allows us to push the applicability range of known at-the-money skew approximation formulae from CLT type log-moneyness deviations of order t1/2 (works of Alòs, León & Vives and Fukasawa) to the wider moderate deviations regime.

2010 Mathematics Subject Classification:

1. Introduction

Since the groundbreaking work of Gatheral et al. (Citation2014), the past two years have brought about a gradual shift in volatility modeling, leading away from classical diffusive stochastic volatility models towards so-called rough volatility models. The term was coined in Gatheral et al. (Citation2014) and Bayer et al. (Citation2016), and it essentially describes a family of (continuous-path) stochastic volatility models where the driving noise of the volatility process has Hölder regularity lower than Brownian motion, typically achieved by modeling the fundamental noise innovations of the volatility process as a fractional Brownian motion with Hurst exponent (and hence Hölder regularity) H<1/2. Here, we would also like to mention pioneering work on asymptotics for rough volatility models in Alòs et al. (Citation2007) and Fukasawa (Citation2011). A major appeal of such rough volatility models lies in the fact that they effectively capture several stylized facts of financial markets both from a statistical (Gatheral et alCitation2014; Bennedsen et alCitation2016) and an option-pricing point of view (Bayer et alCitation2016). In particular, with regards to the latter point of view, a widely observed empirical phenomenon in equity markets is the ‘steepness of the smile on the short end’ describing the fact that as time to maturity becomes small the empirical implied volatility skew follows a power law with negative exponent, and thus becomes arbitrarily large near zero. While standard stochastic volatility models with continuous paths struggle to capture this phenomenon, predicting instead a constant at-the-money implied volatility behavior on the short end (Gatheral Citation2011), models in the fractional stochastic volatility family (and more specifically so-called rough volatility models) constitute a class, well-tailored to fit empirical implied volatilities for short dated options.

Typically, the popularity of asset pricing models hinges on the availability of efficient numerical pricing methods. In the case of diffusions, these include Monte Carlo estimators, PDE discretization schemes, asymptotic expansions and transform methods. With fractional Brownian motion being the prime example of a process beyond the semimartingale framework, most currently prevalent option pricing methods – particularly the ones assuming semimartingality or Markovianity – may not easily carry over to the rough setting. In fact, the memory property (aka non-Markovianity) of fractional Brownian motion rules out PDE methods, heat kernel methods and all related methods involving a Feynman-Kac-type Ansatz. Previous work has thus focused on finding efficient Monte Carlo simulation schemes (Bayer et alCitation2016; Bennedsen et alCitation2017; Bayer et alCitation2017) or – in the special case of the Rough Heston model – on an explicit formula for the characteristic function of the log-price (see El Euch and Rosenbaum Citation2016), thus in this particular model making pricing amenable to Fourier based methods. In our work, we rely on small-maturity approximations of option prices. This is a well-studied topic for which we mention (with no claim to completeness) a number of works, either based on large deviations or central limit type scaling regime, that inspired this work: Alòs et al. (Citation2007), Fukasawa (Citation2011), Deuschel et al. (Citation2014a), Deuschel et al. (Citation2014b) and Fukasawa (Citation2017), also Medvedev and Scaillet (Citation2003Citation2007), Osajima (Citation2007Citation2015), Guennoun et al. (Citation2014), Mijatović and Tankov (Citation2016) and especially Forde and Zhang (Citation2017). Rather recently, Friz et al. (Citation2018) introduced another regime called moderately-out-of-the-money (MOTM), which, in a sense, effectively navigates between the two regimes mentioned above, by rescaling the strike with respect to the time to maturity. This approach has various advantages. On the one hand, it reflects the market reality that as time to maturity approaches zero, strikes with acceptable bid-ask spreads tend to move closer to the money (see Friz et alCitation2018 for more details). On the other hand, it allows us to zoom in on the term structure of implied volatility around the money at a high resolution scale. To be more specific, our paper adds to the existing literature in two ways. First, we obtain a generalization of the Osajima energy expansion (Osajima Citation2015) to a non-Markovian case, and using the new expansion, we extend the analysis of Friz et al. (Citation2018) to the case, where the volatility is driven by a rough (H<1/2) fractional Brownian motion. Indeed, Laplace approximation methods on Wiener space in the spirit of Azencott (Citation1982Citation1985), Ben Arous (Citation1988) and Bismut (Citation1984) can be adapted to the present context, so that our analysis builds upon this framework in a fractional setting. Unlike many other works in this field, we do not rely on density expansions. Finally, using a version of the ‘rough Bergomi model’ (Bayer et alCitation2016), we demonstrate numerically that our implied volatility asymptotics capture very well the geometry of the term structure of implied volatility over a wide array of maturities, extending up to a year.

The paper is organized as follows: In Section 2 we set the scene, describing the class of models included in our framework ((Equation1) and (Equation2)) and recalling some known results ((Equation4) and (Equation7)), which are the starting point of our analysis. Most importantly, we argue that for small-time considerations it would suffice to restrict our attention to a class of stochastic volatility models of the form (Equation3) with a volatility process driven by a Gaussian Volterra process such as in (Equation2). We formulate general assumptions on the Volterra kernel (Assumptions 2.1 and 2.5) and on the function σ in (Equation3) (Assumption 2.4) under which our results are valid. In Section 3 we gather our main results, concerning a higher order expansion of the energy (Theorem 3.1), and a general expansion formula for the corresponding call prices. We derive the classical Black-Scholes expansion for the call price, using the latter result mentioned above. In addition, in Section 3 we formulate moderate deviation expansions, which allow us to derive the corresponding asymptotic formulae for implied volatilities and implied volatility skews. Finally, Section 4 displays our simulation results. Sections 57 are devoted to proofs of the energy expansion, the price expansion and the moderate deviations expansion, respectively. In the appendix, we have collected some auxiliary lemmas, which are used in different sections.

2. Exposition and assumptions

We consider a rough stochastic volatility model, normalized to r=0 and S0=1, of the form suggested by Forde and Zhang (Citation2017) (1) dStSt=σ(Bˆt)dρ¯Wt+ρBt.(1) Here (W,B) are two independent standard Brownian motions, ρ(1,1) a correlation parameter, and ρ¯2=1ρ2. Then ρ¯W+ρB is another standard Brownian motion which has constant correlation ρ with the factor B, which drives the stochastic volatility σstocht,ω:=σ(Bˆtω)σ(Bˆ). Here σ(.) is some real-valued function, typically smooth but not bounded, and we will denote by σ0:=σ(0) the spot volatility, with Bˆ a Gaussian (Volterra) process of the form (2) Bˆt=0tKt,sdBs,t0,(2) for some kernel K, which shall be further specified in Assumptions 2.1 and 2.5 below. The log-price Xt=log(St) satisfies (3) dXt=12σ2(Bˆt)dt+σ(Bˆt)dρ¯W+ρB,X0=0.(3) Recall that by Brownian scaling, for fixed t>0, (Bts,Wts)s0=lawε(Bs,Ws)s0,where εε(t)t1/2. As a direct consequence, classical short-time SDE problems can be analyzed as small-noise problems on a unit time horizon. For our analysis, it will also be crucial to impose such a scaling property on the Gaussian process Bˆ (more precisely, on the kernel K in (Equation2)) driving the volatility process in our model:

Assumption 2.1

Small time self-similarity

There exists a number t0 with 0<t01 and a function tεˆ=εˆ(t), 0tt0, such that (Bˆts:0st0)=law(εˆBˆs:0st0).

In fact, we will always have εˆεˆ(t)tH=ε2H, which covers the examples of interest, in particular standard fractional Brownian motion Bˆ=BH or Riemann-Liouville fBM with explicit kernel K(t,s)=2HtsH1/2. (This is very natural, even from a general perspective of self-similar processes, see Lamperti Citation1962.)

We insist that no (global) self-similarity of Bˆ is required, as only Bˆ|[0,t] for arbitrarily small t matters.

Remark 2.2

It should be possible to replace the fractional Brownian motion by a certain fractional Ornstein-Uhlenbeck process in the results obtained in this paper. Intuitively, this replacement creates a negligible perturbation (for t1) of the fBm environment. A similar situation was in fact encountered in Cass and Friz (Citation2010), where fractional scaling at times near zero was important. To quantify the perturbation, the authors of Cass and Friz (Citation2010) introduced an easy to verify coupling condition (see Corollary 2 in Cass and Friz Citation2010). It should be possible to employ a version of this condition in the present paper to justify the replacement mentioned above. We will however not pursue this point further here.

Remark 2.3

Throughout this article, one can consider a classical (Markovian, diffusion) stochastic volatility setting by taking K1, or equivalently H1/2, by simply ignoring all hats ( ˆ ) in the sequel. In particular then, εˆ/ε1 in all subsequent formulae.

General facts on large deviations of Gaussian measures on Banach spaces (Deuschel and Stroock Citation1989) such as the path space C([0,1],R3) imply that a large deviation principle holds for the triple {εˆ(W,B,Bˆ):εˆ>0}, with speed εˆ2 and rate function (4) 12hH012+12fH012,f,hH01 and fˆ=Kf˙,+,otherwise,(4) where Kf˙(t):=0tKt,sf˙(s)ds for fH01, the space of absolutely continuous paths with L2 derivative (5) H01:=f:[0,1]RcontinuousfH012:=01f˙(s)2ds<, f(0)=0.(5) This enables us to derive a large deviations principle for X in (Equation3): the (local) small-time self-similarity property of Bˆ (Assumption 2.1) implies that Xt=lawX1ε where dXtε=σ(εˆBˆt)εdρ¯Wt+ρBt12ε2σ2(εˆBˆt)dt,X0ε=0. For what follows, it will be convenient to consider a rescaled version of (Equation3) dXˆtεdεˆεXtε=σ(εˆBˆt)εˆdρ¯Wt+ρBt12εεˆσ2(εˆBˆt)dt,Xˆ0ε=0. Under a linear growth condition on the function σ, Forde and Zhang (Citation2017) use the extended contraction principle to establish a large deviations principle for (Xˆ1ε) with speed εˆ2. More precisely, with (6) ϕ1h,f:=Φ1(h,f,fˆ)=01σ(fˆ)dρ¯h+ρf,(6) the rate function is given by (7) Ix=infh,fH011201h˙2dt+1201f˙2dt:ϕ1h,f=x=inffH0112xρσ(fˆ),f˙2ρ¯2σ2(fˆ),1+1201f˙2dt,(7) where , denotes the inner product on L2([0,1],dt). Several other proofs (under varying assumptions on σ) have appeared since (Jacquier et alCitation2017; Bayer et alCitation2017; Gulisashvili Citation2017).

As a matter of fact, this paper relies on moderate – rather than large – deviations, as emphasized in (iiic) below. To this end, let us make

Assumption 2.4

  1. (Positive spot vol) Assume σ:RR is smooth with σ0:=σ(0)>0.

  2. (Roughness) The Hurst parameter H satisfies H(0,1/2].

  3. (Martingality) The price process S=expX is a martingale.

  4. (Short-time moments) m< t>0:E(Stm)<.

While condition (iiia) hardly needs justification, we emphasize that conditions (iiia-b) are only used to the extent that they imply condition (iiic) given below (which thus may replace (iiia-b) as an alternative, if more technical, assumption). The reason we point this out explicitly is that all the conditions (iiia-c) are implicit (growth) conditions on the function σ(.). For instance, (iiia-b) was seen to hold under a linear growth assumption (Forde and Zhang Citation2017; Gulisashvili Citation2017), whereas the log-normal volatility case (think of σ(x)=ex) is complicated. Martingality, for instance, requires ρ0 and there is a critical moment m=m(ρ), even when ρ<0. See Sin (Citation1998), Jourdain (Citation2004) and Lions and Musiela (Citation2007) for the case H=1/2 and the forthcoming work (Friz and Gassiat Citation2018) for the general rough case H(0,1]. We view (iiic) simply as a more flexible condition that can hold in situations where (iiib) fails.

  1. (Call price upper moderate deviation bound) For every β(0,H), and every fixed x>0, and xˆε:=xε12H+2β, E[(eX1εexˆε)+]expx2+o(1)2σ02ε4H4β.

This condition is reminiscent of the ‘upper part’ of the large deviation estimate obtained in Forde and Zhang (Citation2017) (8) E[(eX1εexε12H)+]=expI(x)+o(1)ε4H.(8) If fact, if one formally applies this with x replaced by xε2β, followed by Taylor expanding the rate function, I(xε2β)12I(0)x2ε4β=12σ02x2ε4β, one readily arrives at the estimate (iiic). Unfortunately, o(1)=ox(1) in (Equation8), which is a serious obstacle in making this argument rigorous. Instead, we will give a direct argument (Lemma 7.1) to see how (iiia-b) implies (iiic).

In the sequel, we will use another mild assumption on the kernel.

Assumption 2.5

The kernel K has the following properties

  1. Bˆt=0tK(t,s)dBs has a continuous (in t) version on [0,1].

  2. t[0,1]:0tK(t,s)2ds<.

Note that the Riemann-Liouville kernel K(t,s)=2H(ts)γ, γ=H1/2 satisfies Assumption 2.5.

Remark 2.6

Assumption 2.5 implies that the Cameron-Martin space H of Bˆ is given by the image of H01 under K, i.e. H={Kf˙fH01}. See Lemma 5.3 and Remark 5.4 for more details. A reference and also a sufficient condition for Assumption 2.5 (i) can be found e.g. in Decreusefond (Citation2005, Section 3).

3. Main results

The following result can be seen as a non-Markovian extension of work by Osajima (Citation2015). The statement here is a combination of Theorem 5.10 and Proposition (5.14) below. Recall that σ0=σ(0) represents spot-volatility. We also set σ0σ(0).

Theorem 3.1

Energy expansion

The rate function (or energy) I in (Equation7) is smooth in a neighborhood of x=0 (at-the-money) and it is of the form Ix=1σ02x226ρσ0σ04010tK(t,s)dsdtx33!+O(x4).

The next result is an exact representation of call prices, valid in a non-Markovian generality, and amenable to moderate- and large-deviation analysis (Theorem 3.4 below).

Theorem 3.2

Pricing formula

For a fixed log-strike x0 and time to maturity t>0, set xˆ:=(ε/εˆ)x, where ε=t1/2 and εˆ=tH=ε2H, as before. Then we have (9) c(xˆ,t)=EexpXtexpxˆ+=eIx/εˆ2eε/εˆxJε,x,(9) where Jε,x:=Ee(Ix/εˆ2)UˆεexpεεˆUˆε1eIxR2ε1Uˆε0 and Uˆε is a random variable of the form (10) Uˆε=εˆg1+εˆ2R2ε(10) with g1 a centred Gaussian random variable, explicitly given in equation (Equation38) below, and R2ε is a (random) remainder term, in the sense of a stochastic Taylor expansion in εˆ, see Lemma 6.2 for more details.

Example 3.3

Black-Scholes model

We fix volatility σ()σ>0, and H=1/2 so that εˆ=ε and all ˆ can be omitted. Energy is given by I(x)=x2/2σ2 and Uε=εg1+ε2R2εεσW1ε2σ2/2 with R2ε=R2σ2/2 independent of ϵ. Moreover, (11) Jε,x=Ee(Ix/ε2)UεeUε1eIxR21Uε0=Ee(Ix/ε)g1eεg1ε2σ2/211{g1εσ2/2}=EeαW1eεσW1(εσ)2/211{W1εσ/2}=e(εσ)2/2Mα+εσMα(11) with α:=I(x)σ/ε=(1/σ)(x/ε), and, in terms of the standard Gaussian cdf Φ, Mβ:=EeβW11{W1εσ/2}=eβ2/2Φβεσ2. Using the expansion Φ(y)=(1/y2π)ey2/2(1y2+), as y one deduces, for fixed x>0, the asymptotic relation, as ε0, (12) Jε,xex/22πε3σ3x2.(12) We will be interested (cf. Theorem 3.4) in replacing x by x~=xε2β0 for β>0. This gives α~=(1/σ)(x/ε12β) and the above analysis, now based on α~, remains validFootnote1 for β in the ‘moderate’ regime β[0,1/2) and we obtain (13)  x>0,β[0,1/2):Jε,xε2β12πε34βσ3x2.(13) Let us point out, for the sake of completeness, that a similar expansion is not valid for β>1/2. To see this, first note that (Equation9) implies that J(ε,x)|x=0 is precisely the ATM call price with time t=ε2 from expiration. Well-known ATM asymptotics then imply that J(ε,x)|x=0(1/2π)εσ as ε0. These asymptotics are unchanged in case of o(t1/2)=o(ε) out-of-moneyness (‘almost-at-the-money’ in the terminology of Friz et alCitation2018), which readily implies  x>0,β>1/2:Jε,xε2β12πεσ=const×ε At last, we have the borderline case β=1/2, or x~=xε. From e.g. Muhle-Karbe and Nutz (Citation2011, Theorem 3.1), we see that c(xε,ε2)a(x;σ)ε with positive constant a(x;σ). A look at (Equation9) then reveals  x>0:Jε,xεa(x;σ)εex2/2σ2=const×ε. For the call price expansion in the large / moderate deviations regime, β[0,1/2), the polynomial in ϵ-behavior of (Equation13) implies that the J-term in the pricing formula will be negligible on the moderate / large deviation scale, in the sense for any θ>0, we have εθlogJ(ε,xε2β)0 as ε0. Consequently, with kt=ktβ, for t=ε2, k>0, β[0,1/2), we get the ‘moderate’ Black-Scholes call price expansion, logcBS(kt,t)=1t12βk22σ21+o1as t0.

While the above can be confirmed by elementary analysis of the Black–Scholes formula, the following theorem exhibits it as an instance of a general principle. See Friz et al. (Citation2018) for a general diffusion statement.

Theorem 3.4

Moderate Deviations

In the rough volatility regime H(0,1/2], consider log-strikes of the form kt=kt1/2H+βfor a constant  k0. (i) For β(0,H), and every θ>0, we have logc(kt,t)=I′′0t2H2βk22+O(t3β2H)+O(tθ)as t0. (ii) For β(0,23H), and every θ>0, we have logc(kt,t)=I′′0t2H2βk22+I′′′0t2H3βk36+O(t4β2H)+O(tθ)ast0. Moreover, I′′0=1σ02,I′′′0=6ρσ0σ04010tK(t,s)dsdt=6ρσ0σ04K1,1, where , is the inner product in L2([0,1]).

Remark 3.5

In principle, further terms (of order tiβ2H, i=4,5,) can be added to this expansion of log call prices, given that the energy has sufficient regularity, see Theorem 3.6. We also note that, for small enough β, the error term O(tθ) can be omitted. In any case, one can replace the additive error bounds by (cruder) ones, where the right-most term in the expansion is multiplied with (1+o(1)), as was done in Friz et al. (Citation2018).

Proof of Theorem 3.4

We apply Theorem 3.2 with xˆ=kt=kt1/2H+β, i.e. with x=ktβ=kε2β. In particular, we so get, with εˆ=tH and ε=t1/2, c(kt,t)=eIx/εˆ2eε/εˆxJε,kε2β. The technical Proposition 7.3 asserts that, for fixed k>0, the factor J is negligible in the sense that, for every θ>0, εθlogJ(ε,kε2β)0as ε0. The theorem now follows immediately from the Taylor expansion of I(x) around x=0 (see Theorem 3.1), plugging in x=ktβ. Indeed, replacing I(x) by the Taylor-jet seen in (i),(ii), leads exactly to an error term O(t3β2H), resp. O(t4β2H) .

Fix real numbers k>0, 0<H<12, 0<β<H, and an integer n2. For every t>0, set kt=kt1/2H+β, and denote φn,H,β,θ(t)=maxt2H2βθ, t(n1)β. Here, θ>0 can be arbitrarily small. It is clear that for all small t and θ small enough, φn,H,β,θ(t)=t2H2βθ2H2β(n1)β2Hn+1β, while φn,H,β,θ(t)=t(n1)β2H2β>(n1)ββ<2Hn+1. The following statement provides an asymptotic formula for the implied variance.

Theorem 3.6

Suppose 0<β<2H/n and θ>0 small enough. Then as t0 (and for k>0), (14) σimpl(kt,t)2=j=0n2(1)j2jI′′(0)j+1i=3nI(i)(0)i!ki2t(i2)βj+Oφn,H,β,θ(t).(14) The O-estimate in (Equation14) depends on n, H, β, θ, and k. It is uniform on compact subsets of [0,) with respect to the variable k.

Remark 3.7

Using the multinomial formula, we can represent the expression on the left-hand side of (Equation14) in terms of certain powers of t. However, the coefficients become rather complicated.

Remark 3.8

Let an integer n2 be fixed, and suppose we would like to use only the derivatives I(i)(0) for 2in in formula (Equation14) to approximate σimpl(kt,t)2. Then, the optimal range for β is the following: 2H/(n+1)β<2H/n. On the other hand, if β is outside of the interval [2H/(n+1),2H/n), more derivatives of the energy function at zero may be needed to get a good approximation of the implied variance in formula (Equation14).

We will next derive from Theorem 3.6 several asymptotic formulas for the implied volatility. In the next corollary, we take n=2.

Corollary 3.9

As t0, (15) σimpl(kt,t)=σ0+O(φ2,H,β,θ(t)).(15)

Corollary 3.9 follows from Theorem 3.6 with n=2, the equality (16) I′′(0)=σ02(16) given in Theorem 3.4, and the Taylor expansion 1+h=1+O(h) as h0.

In the next corollary, we consider the case where n=3.

Corollary 3.10

Suppose β<2H/3. Then, as t0, (17) σimpl(kt,t)=σ0+ρσ0σ0K1,1ktβ+O(φ3,H,β,θ(t)).(17)

Corollary 3.10 follows from Theorem 3.6 with n=3, formula (Equation16), the equality (18) I′′′(0)=6ρσ0σ04K1,1(18) (see Theorem 3.4), and the expansion 1+h=1+12h+O(h2) as h0.

Using Corollary 3.10, we establish the following implied volatility skew formula in the moderate deviation regime.

Corollary 3.11

Let 0<H<12, 0<β<23H, and fix y,z>0 with yz. Then as t0, (19) σimpl(yt1/2H+β,t)σimpl(zt1/2H+β,t)(yz)t1/2H+βρσ0σ0K1,1tH1/2.(19)

Remark 3.12

Corollary 3.11 complements earlier works of Alòs et al. (Citation2007) and Fukasawa (Citation2011Citation2017). For instance, the following formula can be found in Fukasawa (Citation2017, p. 6), see also Fukasawa (Citation2011, p. 14): (20) σimpl(yt1/2,t)σimpl(zt1/2,t)(yz)t1/2ρC(H)σ0σ0tH1/2.(20) In formula (Equation20), we employ the notation used in the present paper. Our analysis shows that the applicability range of skew approximation formulas is by no means restricted to the Central Limit Theorem type log-moneyness deviations of order t1/2. It also includes the moderate deviations regime of order t1/2H+β. The previous rate is clearly t1/2 as t0.

Remark 3.13

Symmetry

Write Φ1(W,B,Bˆ;ρ;σ) for the ‘Itô-type map’ Φ1(W,B,Bˆ):=01σ(Bˆ)dρ¯W+ρB. It equals, in law, Φ1(W,B,Bˆ;ρ;σ()), and indeed all our formulae are invariant under this transformation. In particular, the skew remains unchanged when the pair (ρ,σ0) is replaced by (ρ,σ0).

4. Simulation results

We verify our theoretical results numerically with a variant of the rough Bergomi model (Bayer et alCitation2016) which fits nicely into the general rough volatility framework considered in this paper. As before, the model has been normalized such that S0=1 and r=0. We let (W,B) be two independent Brownian motions and ρ(1,1) with ρ¯2=1ρ2 such that Z=ρ¯W+ρB is another Brownian motion having constant correlation ρ with B. For some spot volatility σ0 and volatility of volatility parameter η, we then assume the following dynamics for some asset S: (21) dStSt=σ(Bˆt)dZt(21) (22) σ(x)=σ0exp12ηx(22) where Bˆ is a Riemann-Liouville fBM given by Bˆt=2H0t|ts|H1/2dBs. The approach taken for the Monte Carlo simulations of the quantities we are interested in is the one initially explored in the original rough Bergomi pricing paper (Bayer et alCitation2016). That is, exploiting their joint Gaussianity, where we use the well-known Cholesky method to simulate the joint paths of (Z,Bˆ) on some discretization grid D. With (Equation22) being an explicit function in terms of the rough driver, an Euler discretisation of the Ito SDE (Equation21) on D then yields estimates for the price paths.

The Cholesky algorithm critically hinges on the availability and explicit computability of the joint covariance matrix of (Z,Bˆ) whose terms we readily compute below.Footnote2

Lemma 4.1

For convenience, define constants γ=12H[0,12) and DH=2H/(H+12) and define an auxiliary function G:[1,)R by (23) G(x)=2H11γxγ+γ1γx(1+γ)×12γ2F1(1,1+γ,3γ,x1)(23) where 2F1 denotes the Gaussian hypergeometric function (Olver et alCitation2010). Then the joint process (Z,Bˆ) has zero mean and covariance structure governed by Var[Bˆt2]=t2H,for t0,Cov[BˆsBˆt]=t2HG(s/t),for s>t0,Cov[BˆsZt]=ρDH(sH+1/2(smin(t,s))H+1/2),for t,s0,Cov[ZtZs]=min(t,s),for t,s0.

Numerical simulationsFootnote3 confirm the theoretical results obtained in the last section. In particular – as can be seen in figure  – the asymptotic formula for the implied volatility (Equation17) captures very well the geometry of the term structure of implied volatility, with particularly good results for higher H and worsening results as H0. Quite surprisingly, despite being an asymptotic formula, it seems to be fairly accurate over a wide array of maturities extending up to a single year.

Figure 1. Illustration of the term structure of implied volatility of the Modified Rough Bergomi model in the Moderate deviations regime with time-varying log-strike kt=0.4tβ. Depicted are the asymptotic formula (equation (Equation17), dashed line) and an estimate based on N=108 samples of a MC Cholesky Option Pricer (solid line) with 500 time steps. Model parameters are given by spot vol σ00.2557, vvol η=0.2928 and correlation parameter ρ=0.7571.

Figure 1. Illustration of the term structure of implied volatility of the Modified Rough Bergomi model in the Moderate deviations regime with time-varying log-strike kt=0.4tβ. Depicted are the asymptotic formula (equation (Equation17(17) σimpl(kt,t)=σ0+ρσ0′σ0⟨K1,1⟩ktβ+O(φ3,H,β,θ(t)).(17) ), dashed line) and an estimate based on N=108 samples of a MC Cholesky Option Pricer (solid line) with 500 time steps. Model parameters are given by spot vol σ0≈0.2557, vvol η=0.2928 and correlation parameter ρ=−0.7571.

5. Proof of the energy expansion

Consider dX=12σ2(Y)dt+σYdρ¯dW+ρdB,X0=0dY=dBˆ,Y0=0 where Bˆt=0tK(t,s)dBs for a fixed Volterra kernel (recall (Equation3) in the previous section). We study the small noise problem (Xε,Yε) where (W,B,Bˆ) is replaced by (εW,εB,εˆBˆ). The following proposition roughly says that PX1εεεˆxexpIxεˆ2.

Proposition 5.1

Forde and Zhang Citation2017

Under suitable assumptions (cf. Section 2), the rescaled process ((εˆ/ε)X1ε:ε0) satisfies an LDP (with speed εˆ2) and rate function (24) Ix=inffH01xρG(f)22ρ¯2F(fˆ)+12EfinffH01IxfIxfx,(24) where Gf=01σKf˙sf˙sdsσKf˙,f˙σ(fˆ),f˙Ff=01σKf˙s2dsσ2Kf˙,1σ2(fˆ),1Ef=01f˙s2dsf˙,f˙

The rest of this section is devoted to analysis of the function I as defined in (Equation24). First, we derive the first order optimality condition for the above minimization problem.

Proposition 5.2

First order optimality condition

For any xR we have at any local minimizer f=fx of the functional Ix in (Equation24) that (25) ftx=ρxρGfxσKf˙x,10,t+σKf˙xf˙x,K10,tρ¯2Ffx+xρGfx2ρ¯2F2fxσσKf˙x,K10,t,(25) for all t[0,1].

Proof.

We denote ab whenever a=b+o(δ) for a small parameter δ. We expand Ef+δgEf+2δf˙,g˙Ff+δgFf+δσ2Kf˙,Kg˙Gf+δgGf+δσKf˙,g˙+σKf˙f˙,Kg˙ If f=fx is a minimizer then δIx(f+δg) has a minimum at δ=0 for all g. We expand Ixf+δg=xρGf+δg22ρ¯2Ff+δg+12E(f+δg)xρGfδρσKf˙,g˙+σKf˙f˙,Kg˙22ρ¯2Ff+δσ2Kf˙,Kg˙+12E(f)+δf˙,g˙xρGf2δ2ρxρGfσKf˙,g˙+σKf˙f˙,Kg˙2ρ¯2Ff1+δFfσ2Kf˙,Kg˙+12Ef+δf˙,g˙xρGf2δ2ρxρGfσKf˙,g˙+σKf˙f˙,Kg˙2ρ¯2FfxρGf22ρ¯2FfδFfσ2Kf˙,Kg˙+12Ef+δf˙,g˙. As a consequence, we must have, for f=fx and every g˙L2[0,1] 0=ddδIxf+δgδ=0=ρxρGfσKf˙,g˙+σKf˙f˙,Kg˙ρ¯2FfxρGf2ρ¯2F2fσσKf˙,Kg˙+f˙,g˙. Recall f0x=0, any x. We now test with g˙=1[0,t] for a fixed t[0,1] and obtain ftx=ρxρGfxσKf˙x,10,t+σKf˙xf˙x,K10,tρ¯2Ffx+xρGfx2ρ¯2F2fxσσKf˙x,K10,t.

5.1. Smoothness of the energy

Having formally identified the first order condition for minimality in (Equation24), we will now show that the energy xI(x) is a smooth function. More precisely, we will use the implicit function theorem to show that the minimizing configuration fx is a smooth function in x (locally at x=0). As Ix is a smooth function, too, this will imply smoothness of xIx(fx)=I(x), at least in a neighborhood of 0.

As the Cameron-Martin space H of the process Bˆ continuously embeds into C([0,1]), K maps H01 continuously into C([0,1]), i.e. there is a constant C>0 such that for any fH01 we have (26) Kf˙CfH01.(26) This result will follow from

Lemma 5.3

Let (Vt:0t1) be a continuous, centred Gaussian process and H its Cameron-Martin space. Then we have the continuous embedding HC[0,1]. That is, for some constant C, hChH.

Proof.

By a fundamental result of Fernique, applied to the law of V as Gaussian measure on the Banach space (C[0,1],), the random variable V has Gaussian integrability. In particular, σ2:=E(V2)<, On the other hand, a generic element hH can be written as ht=E[VtZ] where Z is a centred Gaussian random variable with variance hH2, see, e.g. Friz and Hairer (Citation2014, page 150). By Cauchy–Schwarz, htEVt1/2hHσhH and conclude by taking the sup over on the l.h.s. over t[0,1].

Remark 5.4

Assume V is of Volterra form, i.e. Vt=0tK(t,s)dBs. Then it can be shown (see Decreusefond Citation2005, Section 3) that H is the image of L2 under the map K:f˙fˆ:=t0tKt,sf˙sds and Kf˙H=a˙fL2. In particular then, applying the above with h=Kf˙H, gives Kf˙CKf˙H=Cf˙L2=CfH01.

5.1.1. The uncorrelated case

We start with the case ρ=0 as the formulas are much simpler in this case.

By Proposition 5.2, any local optimizer f=fx of the functional Ix:H01R in the uncorrelated case ρ=0 satisfies for any t[0,1] ft=x2F2fσσKf˙,K1[0,t]. We define a map H:H01×RH01 by (27) H(f,x)(t):=ftx2F2fσσKf˙,K1[0,t].(27) Hence, for given xR, any local optimizer f must solve H(f,x)=0. As one particular solution is given by the pair (0,0), we are in the realm of the implicit function theorem. We need to prove that

  • (f,x)H(f,x) is locally smooth (in the sense of Fréchet);

  • DH(f,x):=(/f)H(f,x) is invertible in (0,0).

Note that invertibility should hold for x small enough, as DH(f,x)=idH01x2R for some R, which is invertible as long as R has a bounded norm for sufficiently small x.

Remark 5.5

The method of proof in this section is purely local in H01. Hence, we only really need smoothness of σ locally around 0. Note, however, that stochastic Taylor expansions used in Section 6 will actually require global smoothness of σ.

Lemma 5.6

The functions F:H01R and R1:H01C([0,1]) defined by R1(f)(t):=σσKf˙,K1[0,t],t[0,1], are smooth in the sense of Fréchet.

Proof.

For N1 we note that the Gateaux derivative of F satisfies DNF(f)g1,,gN=01dNdxNσ2(Kf˙)Kg˙1Kg˙Nds. By Lemma 5.3, we can bound DNF(f)g1,,gNconst01Kg1˙(s)KgN˙(s)dsconstKg˙1Kg˙NconstCNg1H01gNH01, for const=(dn/dxn)σ2.Footnote4 Thus, DNF(f) is a multi-linear form on H01 with operator norm DNF(f)(dn/dxn)σ2CN independent of f. As fDNF(f) is continuous, we conclude that DNF(f) as given above is, in fact, a Fréchet derivative.

Let us next consider the functional R1. Note that DNR1(f)(g1,,gN)(t)=sN(Kf˙)Kg˙1Kg˙N,K1[0,t] for sN(x):=(dN/dxN)σ(x)σ(x). Hence, Assumption 2.5 implies that DNR1(f)(g1,,gN)H012=01t1sN(Kf˙)(s)i=1N(Kg˙i)(s)K(s,t)ds2dtsN2i=1NKg˙i201t1K(s,t)2dsdtsN2C2Ni=1NgiH012010sK(s,t)2dtdssN2C2N010sK(s,t)2dtdsi=1NgiH012. We see that the multi-linear map DNR1(f) has operator norm bounded by DNR1(f)sNCN010sK(s,t)2dtds, independent of f. From continuity of fDNR1(f), it follows that DNR1(f) is the N'th Fréchet derivative.

Theorem 5.7

Zero correlation

Assuming ρ=0, the energy I(x) (as defined in (Equation24)) is smooth in a neighborhood of x=0.

Proof.

By construction, we have DH(f,x)=idH01x2A(f) for A:H01L(H01,H01) defined by A(f):=R1(f)DF2(f)+F2(f)DR1(f). Here, R1(f)DF2(f)g=(DF2(f)g)RR1(f)H01. As verified above, H is smooth in the sense of Fréchet. Trivially, DH(0,0)=idH01 is invertible and H(0,0)=0. Therefore, the implicit function theorem implies that there are open neighborhoods U and V of 0H01 and 0R, respectively, and a smooth map xfx from V to U such that H(fx,x)0 and fx is unique in U with this property.

For the energy, we prove that I(x)=Ix(fx) in a neighborhood of x=0. First of all, we show that a minimizer exists. If not, there is a function gH01 with Ix(g)<Ix(fx). For small enough x such a g must be inside a ball with radius ε around 0H01, as Ix(g)12gH012 and limx0Ix(fx)=0. Then note that for any gH01 D2I0(0)(g,g)=gH012>0, where D2Ix(f) denotes the second derivative of fIx(f). By continuity, D2Ix(f) stays positive definite for (x,f) in a neighborhood of (0,0). As noted, for x small enough, both g and fx (and the line connecting them) lie in this neighborhood. For h:=gfx, this implies Ix(g)Ix(fx)=DIx(fx)h+01D2Ix(fx+th)(h,h)dt>0, since DIx(fx)h=0 and D2Ix(fx+tsh)(h,h)>0. This contradicts the assumption that Ix(g)<Ix(fx), and we conclude that fx is, indeed, a minimizer of Ix, implying that I(x)=Ix(fx) locally.

Finally, as xfx is smooth and (f,x)Ix(f)=x2/2F(f)+12fH012 is smooth, we see that xI(x)=Ix(fx) is smooth in a neighborhood of 0. (Note that this arguments relies on σ(0)0, implying that F(f)0 for f in a neighborhood to 0.)

Remark 5.8

Classical counter-examples in the context of the direct method of calculus of variations show that the step of verifying the existence of a minimizer should not be taken too lightly. For instance, the functional J(u):=01(u(s)21)2+u(s)2ds does not have a minimizer in H01, but J can be made arbitrarily close to 0 by choosing piecewise-linear functions u with slope u=1 oscillating around 0. We refer to any text book on calculus of variations. In the situation above, local ‘convexity’ in the sense of a positive definite second derivative prevents this phenomenon. An alternative method of proof for the existence of a minimizer is to show that J is (lower semi-) continuous in the weak sense.

5.1.2. The general case

In the general case (cf. Proposition 5.2), we define the function H:H01×RH01 by (28) H(f,x)(t):=ftρxρGfσKf˙,1[0,t]+σKf˙f˙,K1[0,t]ρ¯2Ff+xρGf2ρ¯2F2fσσKf˙,K1[0,t]=ftρxρG(f)ρ¯2F(f)R2(f)(t)+R3(f)(t)+xρG(f)2ρ¯2F(f)2R1(f)(t),(28) where R2,R3:H01H01 are defined by (29) R2(f)(t):=σ(Kf˙),1[0,t],(29) (30) R3(f)(t):=σ(Kf˙)f˙,K1[0,t],(30) t[0,1].

One easily checks that G, R2, R3 are smooth in the Fréchet sense.

Lemma 5.9

The functions G:H01R, R2:H01H01 and R3:H01H01 are smooth in Fréchet sense.

Proof.

The proof of smoothness is clear. We report the actual derivatives. For G we get DNG(f)g1,,gN=σ(N)Kf˙f˙,i=1NKg˙i+k=1Nσ(N1)Kf˙,g˙kikKg˙i. For R2 and, respectively, R3, we obtain DNR2(f)(g1,,gN)(t)=0tσ(N)(Kf˙)(s)i=1N(Kg˙i)(s)ds, and DNR3(f)(g1,,gN)(t)=σ(N+1)Kf˙f˙K1[0,t],i=1NKg˙i+k=1Nσ(N)Kf˙K1[0,t],g˙kikKg˙i.

Theorem 5.10

Let σ be smooth with σ(0)0. Then the energy I(x) as defined in (Equation24) is smooth in a neighborhood of x=0.

Proof.

The proof is similar to the proof of Theorem 5.7. In fact, the only difference is in establishing invertibility of DH(0,0) and the existence of a minimizer.

Note that (Equation28) contains three terms. The derivative of the first term (ff) is always equal to idH01. For the second term, we note that xρG(f)x=0, f=0=0. Hence, the only non-vanishing contribution to the derivative of the second term evaluated in direction gH01 at x=0, f=0 and t[0,1] is ρ2DG(0)gρ¯2F(0)(R2(0)+R3(0))=ρ2σ0g(1)ρ¯2σ02σ0t+0=ρ2ρ¯2g(1)t. For the same reason, the derivative of the third term at (f,x)=(0,0) vanishes entirely. Hence, (DH(0,0)g)(t)=g(t)+ρ2ρ¯2g(1)t. It is easy to see that gDH(0,0)g is invertible. Indeed, let us construct the pre-image g=DH(0,0)1h of some hH01. At t=1 we have ρ¯2+ρ2ρ¯2g(1)=h(1), implying g(1)=ρ¯2h(1). For 0t<1, we then get g(t)+ρ2ρ¯2g(1)t=g(t)+ρ2ρ¯2ρ¯2h(1)t=g(t)+ρ2h(1)t=h(t), or g(t)=h(t)ρ2h(1)t.

For existence of the minimizer, note that D2J0(0)(g,g)=ρ2ρ¯2g(1)2+gH012, which is again positive definite.

Remark 5.11

Though only formulated in terms of ‘smoothness’, it is easy to show that σCk implies that ICk1 (locally at 0).

5.2. Energy expansion

Having established smoothness of the energy I as well as of the minimizing configuration xfx locally around x=0, we can proceed with computing the Taylor expansion of fx around x=0. We will once more rely on the first order optimality condition given in Proposition 5.2. Plugging the Taylor expansion of fx into Ix will then give us the local Taylor expansion of I(x).

5.2.1. Expansion of the minimizing configuration

Theorem 5.12

We have ftx=αtx+βtx22+Ox3,αt=ρσ0t,βt=2σ0σ03ρ2K1,1[0,t]+K1[0,t],13ρ2tK1,1.

Remark 5.13

Non-Markovian transversality

In the RL-fBM case, K(t,s)=2H|ts|γ with γ=H1/2 one computes 1,K10,t=11+γ2+γ11t2+γC10,1. Interestingly, the transversality condition known from the Markovian setting (q1=0, which readily translates to f˙1x=0 there) remains valid here (for ρ=0), at least to order x2, in the sense that f˙txβtx22=const1t1+γ|t=1=0

Proof of Theorem 5.12

First order expansion:

Up to the order needed in order to get the first order term, we have ftx=αtx+O(x2),ft˙x=αt˙x+O(x2),σ(Kf˙x)=σ0+σ0Kα˙x+O(x2),σ(Kf˙x)=σ0+σ0Kα˙x+O(x2),F(fx)=σ2(Kf˙x),1=σ02+O(x),G(fx)=σ(Kf˙x),f˙x=σ0,α˙x+O(x2). Therefore, σ(Kf˙x),1[0,t]=σ0t+O(x),σ(Kf˙x)f˙x,K1[0,t]=O(x),σσ(Kf˙x),K1[0,t]=O(1),xρG(fx)=(1ρσ0α1)x+O(x2),(xρG(fx))2=O(x2). This yields for the first order term in (Equation25) αt=ρ(1ρσ0α1)ρ¯2σ0t. Setting t=1, we get α1=ρρ¯2σ0ρ2ρ¯2α1, which is solved by α1=ρ/σ0. Inserting this term back into the equation for αt, we get (31) αt=ρσ0t.(31)

Second order expansion:

Using (Equation31) and the ansatz ftx=αtx+12βtx2+O(x3), we re-compute the relevant terms appearing in the (Equation25). We have σ(Kf˙x(s))=σ0+σ0ρσ0(K1)(s)x+O(x2) and analogously for σ replaced by σ, σσ. This implies σ(Kf˙x),1[0,t]=σ0t+σ0ρσ0K1,1[0,t]x+O(x2),σ(Kf˙x)f˙x,K1[0,t]=ρσσ0K1[0,t],1x+O(x2),σσ(Kf˙x),K1[0,t]=σ0σ0K1[0,t],1+O(x). Using the notation introduced earlier, we have F(fx)=σ02+2σ0ρK1,1x+O(x2),G(fx)=ρx+12σ0β1+ρ2σ0σ02K1,1x2+O(x3). This directly implies xρG(fx)=ρ¯2xρ12σ0β1+ρ2σ0σ02K1,1x2+O(x3),xρG(fx)2=ρ¯4x22ρ¯2ρ×12σ0β1+ρ2σ0σ02K1,1x3+O(x4). We next compute some auxiliary terms appearing in (Equation25). N1:=ρ(xρG(fx))σ(Kf˙x),1[0,t]+σ(Kf˙x)f˙x,K1[0,t]=ρρ¯2σ0tx+ρ2ρ¯2σ0σ0K1,1[0,t]+K1[0,t],1ρ4σ0σ0tK1,112ρ2σ02tβ1x2+O(x3) The corresponding denominator is ρ¯2F(fx). Using the formula a1x+a2x2+O(x3)b0+b1x+O(x2)=a1b0x+a2b0a1b1b02x2+O(x3), we obtain (32) N1ρ¯2F(fx)=ρσ0tx+ρ2σ0σ03K1,1[0,t]+K1[0,t],1ρ4ρ¯2+2ρ2σ0σ03tK1,112ρ2ρ¯2β1tx2+O(x3)(32) For the second term in (Equation25), let N2:=xρG(fx)2(σσ)(Kf˙x),K1[0,t]=ρ¯4σ0σ0K1[0,t],1x2+O(x3). The corresponding denominator is ρ¯2F(fx)2=ρ¯2σ04+O(x). Hence, (33) N2ρ¯2F(fx)2=ρ¯2σ0σ03K1[0,t],1x2+O(x3).(33) Combining (Equation32) and (Equation33), we get ftx=ρσ0tx+ρ2σ0σ03K1,1[0,t]+K1[0,t],1ρ4ρ¯2σ0σ03tK1,112ρ2ρ¯2β1t2ρ2σ0σ03tK1,1+ρ¯2σ0σ03K1[0,t],1x2+O(x3) We shall next compute β1. Taking the second order terms on both sides and letting t=1, we obtain 12β1=ρ2σ0σ032K1,1ρ4ρ¯2σ0σ03K1,112ρ2ρ¯2β12ρ2σ0σ03K1,1+ρ¯2σ0σ03K1,1. Moving β1 to the other side with 1+ρ2/ρ¯2=1/ρ¯2 and collecting terms on the right hand side, we arrive at 121ρ¯2β1=σ0σ03K1,12ρ2ρ4ρ¯22ρ2+ρ¯2=12ρ2ρ¯2σ0σ03K1,1 We conclude that β1=2(12ρ2)σ0σ03K1,1 Hence, we obtain βt=2σ0σ03ρ2K1,1[0,t]+K1[0,t],13ρ2tK1,1.

5.2.2. Energy expansion in the general case

Now we compute the Taylor expansion of I(x) as defined in Proposition 5.1. We start with the second term. Plugging in the optimal path ftx=αtx+12βtx2+O(x3) (and using a˙β,1=β1 as β0=0) we obtain 12f˙x,f˙x=12ρ2σ02x2+12ρσ0β1x3+O(x4). Inserting β1=2(12ρ2)(σ0/σ03)K1,1 into the above formula for (xρG(fx))2, we get xρG(fx)2=ρ¯4x22ρ¯4ρσ0σ02K1,1x3+O(x4). Recall the denominator 2ρ¯2F(fx)=2ρ¯2σ02+4ρ¯2σ0ρK1,1x+O(x2). Using the expansion of a fraction a2x2+a3x3+O(x4)b0+b1x+O(x2)=a2b0x2+a3b0a2b1b02x3+O(x4), we obtain from xρG(fx)22ρ¯2F(fx)=ρ¯42ρ¯2σ02x2+2ρ¯4ρσ0σ02K1,12ρ¯2σ02ρ¯44ρ¯2σ0ρK1,14ρ¯4σ04x3+O(x4)=ρ¯22σ02x22ρ¯2ρσ0σ04K1,1x3+O(x4). We note that 12ρσ0β12ρ¯2ρσ0σ04K1,1=(12ρ2)2(1ρ2)ρσ0σ04K1,1=ρσ0σ04K1,1. Adding both terms, we arrive at the

Proposition 5.14

The energy expansion to third order gives I(x)=12σ02x2ρσ0σ04K1,1x3+O(x4).

5.2.3. Energy expansion for the Riemann-Liouville kernel

Let us specialize the energy expansion given in Proposition 5.14 for the Riemann-Liouville fBm. Choose γ=H12 and recall that the kernel K takes the form K(t,s)=(ts)γ. We get (K1)(t)=0tK(t,s)ds=0t(ts)γds=t1+γ1+γ. The key term K1,1 appearing in the energy expansion now gives K1,1=01(K1)(t)dt=01t1+γ1+γdt=1(1+γ)(2+γ)=1(H+1/2)(H+3/2). Plugging the above formula into the energy expansion, we obtain the energy expansion for the Riemann-Liouville fractional Browian motion I(x)=12σ02x2ρ(H+1/2)(H+3/2)σ0σ04x3+O(x4). For completeness, let us also fully describe the time-dependence of the second order term βt in the expansion of the optimal trajectory ftx. Unlike the first order time, here we do not have a linear movement any more. Indeed (34) K1,1[0,t]=0t(K1)(s)ds=0ts1+γ1+γds=t2+γ(1+γ)(2+γ),(34) (35) K1[0,t],1=1(1+γ)(2+γ)1(1t)2+γ.(35)

6. Proof of the pricing formula

Fix x0 and xˆ=(ε/εˆ)x where ε=t1/2 and εˆ=tH=ε2H. We have c(xˆ,t)=EexpXtexpxˆ+=EexpX1εexpxˆ+=EexpεεˆXˆ1εexpεεˆx+ where we recall Xˆ1εεˆεX1ε=01σ(εˆBˆ)εˆdρ¯W+ρB12εεˆ01σεˆBˆt2dt. Consider a Cameron-Martin perturbation of Xˆ1ε. That is, for a Cameron-Martin path h=(h,f)H01×H01 consider a measure change corresponding to a transformation εˆ(W,B)εˆ(W,B)+(h,f) (transforming the Brownian motions to Brownian motions with drift), we obtain the Girsanov density (36) Gε=exp1εˆ01h˙sdWs1εˆ01f˙sdBs12εˆ201h˙s2+f˙s2ds.(36) Under the new measure, Xˆ1ε becomes Zˆ1ε, where Zˆ1ε=01σ(εˆBˆt+fˆt)εˆdρ¯Wt+ρBt+dρ¯ht+ρft12εεˆ01σ(εˆBˆt+fˆt)2dt.

Definition 6.1

For fixed x0, write (h,f)Kx if Φ1(h,f,fˆ)=x. Call such (h,f) admissible for arrival at log-strike x. Call (hx,fx) the cheapest admissible control, which attains Ix=infh,fH011201h˙2dt+1201f˙2dt:Φ1h,f,fˆ=x, where we recall that fˆ=Kf˙ and Φ1(h,f,fˆ)=01σ(fˆ)dρ¯h+ρf.

For any Cameron-Martin path (h,f), the perturbed random variable Zˆ1ε admits a stochastic Taylor expansion with respect to εˆ.

Lemma 6.2

Fix (h,f)Kx and define Zˆ1ε accordingly. Then (37) Zˆ1ε=x+εˆg1+εˆ2R2ε,(37) where g1 is a Gaussian random variable, given explicitly by (38) g1=01{σ(fˆt)dρ¯Wt+ρBt+σ(fˆt)Bˆtdρ¯ht+ρft},(38) and (39) R2ε=01σfˆtBˆtdρ¯Wt+ρBt12εεˆ01σ(εˆBˆt+fˆt)2dt+12εˆ20εˆ01σζBˆt+fˆtBˆt2×εˆdρ¯Wt+ρBt+dρ¯ht+ρftεˆζdζ.(39)

Proof.

By a stochastic Taylor expansion for the controlled process Zˆtε with control (h,f)Kx as in Definition 6.1 and thanks to σC2, we have at t=1 Zˆ1ε=01σ(εˆBˆ+fˆ)εˆdρ¯W+ρB+dρ¯h+ρf12εεˆ01σ(εˆBˆt+fˆt)2dt=01σ(fˆ)dρ¯h+ρf+εˆ01{σ(fˆ)dρ¯W+ρB+σ(fˆ)Bˆdρ¯h+ρf}+εˆ201σfˆtBˆtdρ¯Wt+ρBt12εεˆ01σ(εˆBˆt+fˆt)2dt+120εˆ01σζBˆt+fˆtBˆt2×εˆdρ¯Wt+ρBt+dρ¯ht+ρftεˆζdζ. Collecting terms in powers of εˆ and with the random variable g1 as in (Equation38) (recalling that εˆεO(εˆ2)), we have Zˆ1ε=01σ(fˆ)dρ¯h+ρf+εˆg1+O(εˆ2), furthermore, since (h,f)Kx, by the definition of Φ1, it holds that 01σ(fˆ)dρ¯h+ρf=x. This proves the statement (Equation37) and the statement that g1 is Gaussian is immediate from the form (Equation38).

Finally, we determine an explicit form of the Girsanov density Gε for the choice where (hx,fx) in (Equation36) are chosen the cheapest admissible control (cf. Definition 6.1. Similarly to classical works of Azencott, Ben Arous and others, see, for instance, Ben Arous (Citation1988), we show that the stochastic integrals in the exponent of Gε are proportional to the first order term g1 (with factor I(x)) when evaluated at the minimizing configuration (hx,fx).

Lemma 6.3

We have 01h˙txdWt+01f˙txdBt=Ixg1.

Proof.

See Lemma A.2.

With these preparations in place, we are now ready to prove the pricing formula from Section 3.

Proof of Theorem 3.2

With a Girsanov factor (all integrals on [0,1]) Gε=e1/εˆh˙dW1εˆf˙dB12εˆ2h˙2+f˙2dt and (evaluated at the minimizer) Gε|=eIx/εˆ2eIxg1ω/εˆ, we have, setting Uˆε:=Zˆ1εx=εˆg1+εˆ2R2ε c(xˆ,t)=EexpεεˆZˆ1εexpεεˆx+Gε|=eε/εˆxEexpεεˆUˆε1+Gε|=eIx/εˆ2eε/εˆxEexpεεˆUˆε1+eIxg1/εˆ=eIx/εˆ2eε/εˆxEexpεεˆUˆε1e(Ix/εˆ2)Uˆε×eIxR2ε1Uˆε0.=eIx/εˆ2eε/εˆxJε,x.

7. Proof of the moderate deviation expansions

In Section 2, we pointed out that (iiic) is exactly what one gets from (call price) large deviations (Equation8), if heuristically applied to xε2β. We now give a proper derivation based on moderate deviations.

Lemma 7.1

Assume (iiia-b) from Assumption 2.4. Then an upper moderate deviation estimate holds both for calls and digital calls. That is, we have

  1. For every β(0,H), and every fixed x>0, and xˆε:=xε12H+2β, E[(eX1εexˆε)+]expx2+o(1)2σ02ε4H4β

and also (40) P[X1ε>xˆε]expx2+o(1)2σ02ε4H4β.(40)

Proof.

Recall σ(.) smooth but unbounded and recall xˆε:=xε12H+2β. In case of β=0 and H=1/2 a large deviation principle (LDP) for (X1εεˆ/ε) is readily reduced, via exponential equivalence, to a LDP for the family of stochastic Itô integrals given by σ(εˆBˆ)εˆdZ for some Brownian Z, ρ-correlated with B. There are then many ways to establish a LDP for this family. A particularly convenient one, that requires no growth restriction on σ, uses continuity of stochastic integration with respect to the rough path (B,Z,BdZ)=(B,Z,BˆdZ) in suitable metrics, for which a LDP is known (Friz and Hairer Citation2014, Ch 9.3). It was pointed out in Bayer et al. (Citation2017) that a similar reasoning is possible when H<1/2, the rough path is then replaced by a ‘richer enhancement’ of (B,Z), the precise size of which depends on H, for which again one has a LDP. A moderate deviation priniple (MDP) for (X1εεˆ/ε) is a LDP for (ε2βX1εεˆ/ε) for β(0,H). This can be reduced to a LDP, with ε¯:=ε2βεˆ=ε2H2β, for ε2β01σ(εˆBˆ)εˆdZ=01σ(εˆBˆ)ε¯dZ01σε(ε¯Bˆ)ε¯dZ with speed ε¯2. Since σε()σ(ε2β) converges (with all derivatives) locally uniformly to the constant function σ0, and one checks that the above is exponentially equivalent to the (Gaussian) family given by σ0ε¯Z1, with law N(0,σ02ε¯2)=N(0,σ02ε4H4β) which gives (Equation40), even with equality. (By localization, exponential equivalence can again be done for σ without growth restrictions.)

We have not yet used either assumption (iiia-b). These become important in order to extend estimate (Equation40) to the case of genuine call payoffs. We can follow here a well-known argument (e.g. Forde and Jacquier Citation2009; Pham Citation2010; Forde and Zhang Citation2017) with the ‘moderate’ caveat to carry along a factor ε2β. In fact, this is follows precisely the argument of Forde and Zhang (Citation2017) where the authors carry along a factor εˆ/ε=ε2H1. (This provides a unified view on rough and moderate deviations.) The remaining details then follow essentially ‘Appendix C. Proof of Corollary 4.13., part (ii) upper bound’ of Forde and Zhang (Citation2017), noting perhaps that the authors use their assumptions to show validity of what we simply assumed as condition (iiib), and also that one works with the quadratic rate function I(0)x2=x2/2σ02 throughout.

Remark 7.2

By an easy argument similar to ‘Appendix C. Proof of Corollary 4.13., part (i) lower bound’ of Forde and Zhang (Citation2017) one sees that validity of the call price upper bound (iiic) implies the corresponding digital call price upper bound (Equation40.) For this reason, we only emphasized (iiic) but not (Equation40) in Section 2.

In a classical work Azencott (Citation1982) (see also Azencott Citation1985; Ben Arous Citation1988, Théorème 2) obtained asymptotic expansions of functionals of Laplace type on Wiener space, of the type ‘E[exp(F(Xε)/ε2)]’, for small noise diffusions Xε. This refines the large deviation (equivalently: Laplace) principle of Freidlin–Wentzell for small noise diffusions. In a nutshell, for fixed X0=x, Azencott gets expansions of the form ec/ε2(α0+α1ε). His ideas (used by virtually all subsequent works in this direction) are a Girsanov transform, to make the minimizing path ‘typical’, followed by localization around the minimizer (justified by a good large deviation principle), and finally a local (stochastic Taylor) type analysis near the minimizer. None of these ingredients rely on the Markovian structure (or, relatedly, PDE arguments). As a consequence (and motivation for this work) such expansions were also obtained in the (non-Markovian) context of rough differential equations driven by fractional Brownian motion (Inahama Citation2013; Baudoin and Ouyang Citation2015) with H<1/2.

And yet, our situation is different in the sense that call price Wiener functionals do not fit the form studied by Azencott and others, nor can we in fact expect a similar expansion: Example 3.3 gives a Black-Scholes call price expansion of the form constant times ecε2(ε3+). Azencott's ideas are nonetheless very relevant to us: we already used the Girsanov formula in Theorem 3.2 in order to have a tractable expression for J. It thus ‘only’ remains to carry out the localization and do some local analysis.

Proposition 7.3

Let x>0 and β(0,H). Then the factor J is negligible in the sense that, for every θ>0, εθlogJ(ε,xε2β)0as ε0.

Proof.

Step 1. Localization Write xε:=xε2β,xˆε:=xεε12H=xε12H+2β. By definition, E[(eX1εexˆε)+]eIxε/εˆ2exˆε=Jε,xε. Fix x,δ>0 and write δε=δε2β. We claim that (the positive quantity) (41) Jε,xεJδεε,xε=eIxε/εˆ2exˆεE[(eX1εexˆε)1Xˆ1ε>xε+δε](41) is exponentially small, in the sense that, for some c>0 and ε¯2=ε4H4β, Jε,xεJδεε,xε=Oec/ε¯2. There is a battle here between the exploding factor eI(xε)/εˆ2, with exponent Ixεεˆ2I(0)xε22εˆ2=I(0)x22ε4H4β, and on the other hand E[(eX1εexˆε)1Xˆ1ε>xε+δε]exp(x+δ)2+o(1)2σ02ε4H4β where the given estimate is an easy consequence of Lemma 7.1. Since I(0)=1/σ02 we see that the last factor ‘exponentially over-compensates’ the rest, so that the difference is indeed exponentially negligible.

Step 2. Upper bound. For any x>0, recall that Uˆε,x=Uˆε decomposes into a Gaussian random variable g1=g1x and remainder R2ε,x=R2ε. In order to control this remainder without imposing boundedness assumption on σ(.), we will crucially used a ‘localized remainder tail estimate’ as given in Proposition 7.4 below. We have, for any ε(0,1], (42) Jδε,x=Ee(Ix/εˆ2)UˆεexpεεˆUˆε1eIxR2ε1Uˆε[0,δε](eδ1)E[eIx/εˆg1x;Uˆε,x0,δ].(42) To proceed, recall εˆ1g1x=εˆ2Uˆε,xR2ε,x so that, for any κ>0, e(Ix/εˆ)g1x=e(Ix/εˆ)g1x1εˆBˆ;0,1κ+e(Ix/εˆ2)Uˆε,xeIxR2ε,x1εˆBˆ;0,1<κ. Since I(x)>0 for small enough x>0, it follows that (I(x)/εˆ2)Uˆε,x<0 on the event {Uˆε,x[0,δ]}, which leads us to Jδε,x(eδ1)E[e(Ix/εˆ)g1x;|εˆBˆ|;0,1κ]+(eδ1)E[eIxR2ε,x;|εˆBˆ|;0,1<κ](eδ1)E[e(2Ix/εˆ)g1x]P|εˆBˆ|;0,1κ+(eδ1)C where, by Proposition 7.4, the constant C=C(κ) is uniform in small ϵ and x. The square-root terms are computed resp. (Fernique) estimated by exp(Ix)2V(g1x)εˆ2×exp(cκ2/εˆ2) for some c>0 which depends on the law of B (hence H), but is uniform in ϵ and x. Hence, for x small enough, the resulting exponent (I(x))2V(g1x)cκ2 is negative, which is more than enough to conclude the upper bound.

Step 3. Lower bound. Write Eδ,κ[]=E[1Uˆε,x[0,δε]1|εˆBˆ|;[0,1]<κ] and estimate Eδ,κe(Ix/εˆ2)Uˆε/2expεεˆUˆε11/2=Eδ,κe(Ix/εˆ2)Uˆε/2expεεˆUˆε11/2eIxR2ε/2×eIxR2ε/2Jδε,x1/2Eδ,κeIxR2ε1/2 where we used Cauchy–Schwarz and discarded the event {|εˆBˆ|;[0,1]<κ}. The localized remainder estimate provides an upper bound on Eδ,κ[eI(x)R2ε], uniformly over small (enough) ϵ and x.

It then suffices to get a suitable lower bound of the left-hand side above. Indeed, for u[0,εˆ2η]=[0,ε4Hη], with η small enough, not dependent on ϵ, (43) u(e(ε/εˆ)u1)1/2e(Ix/εˆ2)u/2γεεˆu1/2(43) for a constant γ>0 which can also be taken uniformly in small x,ε. Then estimate Eδ,κ[(e(ε/εˆ)Uˆε1)1/2e(Ix/2ε2)Uˆε]γε1/2HE[|Uˆε|1/21Uˆε0,εˆ2η1εB;0,1<κ]. As a quick sanity check, pretend zero remainder so that Uˆε=εˆg1: dropping further the (exponentially close to probability one) event {|εB|;[0,1]<κ}, a Gaussian computation then shows that we are left with (γε1/2H times εˆ1/2 times) E[|g1|1/2;g1[0,εˆ]](const)εˆ3/2. In general, set Vε=Uˆε/εˆ=g1+εˆR2sε, so thatFootnote5 Eκ|Uˆε|1/2;Uˆε[0,εˆ2η]=εˆ1/2Eκ[Vε1/2;Vε[0,εˆη]]. At this stage, it is difficult to treat εˆRε as perturbation of g since, on the given event {Vε[0,εˆη]}, all terms are of order εˆ. We can solve this issue by realizing that we can replace, throughout, x by xε=xε2β. Since I(xε)(const)xε, with see from (Equation43), that in the above estimate the event Uˆε[0,εˆ2η]=[0,ε4Hη] (resp. Vε[0,εˆη]=[0,ε2Hη]) can be replaced by Uˆε[0,ε4H2βη] (resp. Vε[0,ε2H2βη]), possibly with an insignificantly modified constant η. It is now straight-forward to show that the behavior of Eκ[|Vε|1/2;Vε[0,ε2H2βη]] is of the same order as E[g1/2;g[0,ε2H2βη]], the correct behavior (i.e. positive power of ε) is obtained by spelling out the (Gaussian) integral.

Proposition 7.4

Localized remainder tail estimate

For every κ>0, there exists c1,c2>0 such that, for all r and uniformly in small ε,x we have PR2ε>r,|εˆBˆ|;0,1<κc1expc2r

Proof.

We decompose εˆ2R2ε=Mε+Nε in terms of the (local) martingale Mε:=εˆ0σεˆBˆ+fˆσfˆd[ρ¯W+ρB] and the (bounded variation) process Nε:=0σεˆBˆ+fˆσfˆσfˆεˆBd[ρ¯h+ρf]12εεˆ0σ2εˆBˆ+fˆdt. Let τε,κ be the stopping time when εˆBˆ first leaves the uniform ball of radius κ. Then Mtκ,ε:=Mtτε,κε still yields a (local) martingale. The point is that {|εˆBˆ|;[0,1]<κ}={τε,κ>1}. On this event, Mε|[0,1]=Mκ,ε|[0,1] and we can thus replace Mε, in the definition of the remainder, by Mκ,ε. Let K=Kκ,x be the κ-fattening of {f(t):0t1}, recall f=fx, then, for t[0,1], dMκ,εt/dt=εˆ2(σ(εˆBˆt+fˆt)σ(ft))2εˆ4σ;K2|Bˆt|2. Clearly, we can replace K by K~κ which contains all Kκ,x for small x. To summarize, we have, on the event {|εˆBˆ|;[0,1]<κ}, Rε=εˆ2Mκ,ε+εˆ2Nε with [εˆ2Mκ,ε]=O(|Bˆ|;[0,1]2) and, as seen by a similar (but easier) reasoning, εˆ2Nε=O(|Bˆ|;[0,1]2), always for fixed κ>0, but uniformly in small ϵ (equivalently, εˆ) and small x>0. This clearly shows that εˆ2Nε has exponential tails. The same is true for the martingale part, whose bracket is O(Gaussian2). This is exactly the situation for the ‘model’ martingal increment 201BdB=B121 which clearly has exponential tails. To make this rigorous, recall that Gaussian resp. exponential tails are characertized by O(p) resp. O(p)-growth of the Lp-norms. The statement is then an easy consequence of the sharp (upper) BDG constant (Carlen and Kree Citation1991), known to be O(p).

8. Proof of the implied volatility expansion

With Theorem 3.2 in place, we now turn to the proof of the implied volatility expansion, formulated in Theorem 3.6.

Proof of Theorem 3.6

We will use an asymptotic formula for the dimensionless implied variance Vt2=tσimpl(kt,t)2,t>0, obtained in Gao and Lee (Citation2014). It follows from the first formula in Remark 7.3 in Gao and Lee (Citation2014) that (44) Vt2kt22Lt=Okt2Lt2(kt+|logkt|+logLt),t0,(44) where Lt=logc(kt,t), t>0.

We will need the following formula that was established in the proof of Theorem 3.4: (45) Lt=I(ktβ)t2H+O(tθ)(45) as t0, for all x0 and β[0,H) and any θ>0. Let us first assume 2H/(n+1)β<2H/n. Using the energy expansion, we obtain from (Equation45) that (46) Lt=i=2nI(i)(0)i!kitiβ2H+Otθ=I′′(0)2k2t2β2H×1+i=3n2I(i)(0)i!I′′(0)ki2t(i2)β+Ot2H2βθ(46) as t0. The second term in the brackets on the right-hand side of (Equation46) disappears if n=2.

Remark 8.1

Suppose n2 and 2H/(n+1)β<2H/n. Then formula (Equation46) is optimal. Next, suppose n2 and 0<β<2H/(n+1). In this case, there exists mn+1 such that 2H/(m+1)β<2H/m, and hence (Equation46) holds with m instead of n. However, we can replace m by n, by making the error term worse. It is not hard to see that the following formula holds for all n2 and 0<β<2H/(n+1): (47) Lt=i=2nI(i)(0)i!kitiβ2H+Ot(n+1)β2H=I′′(0)2k2t2β2H×1+i=3n2I(i)(0)i!I′′(0)ki2t(i2)β+Ot(n1)β(47) as t0 provided we choose θ small enough.

Let us continue the proof of Theorem 3.6. Since ktt1/2H+β and Ltt2β2H as t0, (Equation44) implies that (48) Vt2=k2t12H+2β2Lt+Ot1+2H2βθ,t0.(48) Next, using the Taylor formula for the function u1/(1+u), and setting u=i=3n2I(i)(0)i!I′′(0)ki2t(i2)β+O(t2H2βθ), we obtain from (Equation46) that (2Lt)1=t2H2βk2I′′(0)j=0n2(1)juj+O(un1) as t0. It follows from 2H/(n+1)β<2H/n that (n1)β2H2β, and hence (2Lt)1=t2H2βk2I′′(0)j=0n2(1)juj+O(t4H4βθ)=t2H2βk2I′′(0)j=0n2(1)ji=3n2I(i)(0)i!I′′(0)ki2t(i2)βj+O(t4H4βθ) as t0. Now, (Equation48) gives Vt2=tI′′(0)j=0n2(1)ji=3n2I(i)(0)i!I′′(0)ki2t(i2)βj+Ot1+2H2βθ as t0. Finally, by canceling a factor of t in the previous formula, we obtain formula (Equation14) for 2H/(n+1)β<2H/n. The proof in the case where β2H/(n+1) is similar. Here we take into account Remark 8.1. This completes the proof of Theorem 3.6.

Acknowledgments

Two referees are thanked for their useful comments. We further thank Martin Forde for valuable feedback.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

We gratefully acknowledge financial support through DFG research grants FR2943/2 and BA5484/1 (C. Bayer, P.K. Friz, B. Stemper), European Research Council Grant CoG-683164 (P.K. Friz), and SNF Early Postdoc Mobility Grant 165248 (B. Horvath) respectively.

Notes

† More terms in the expansion of Φ are needed.

† Note that expressions for the exact same scenario have have been computed before in the original pricing paper (Bayer et alCitation2016), yet in that version the expression for the autocorrelation of the fBM Bˆ was incorrect. We compute and state here all the relevant terms for the sake of completeness.

† The Python 3 code used to run the simulations can be found at github.com/RoughStochVol.

† More precisely, since neither σ nor its derivatives need to be bounded, we need to actually work with a local version of the above estimate, for instance by replacing the max with a sup over a compact set containing {(Kf˙)(t):0t1}.

† Write Eκ for the expected valued restricted to the event {|εB|;[0,1]<κ}

References

  • Alòs, E., León, J.A. 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. doi: 10.1007/s00780-007-0049-1
  • Azencott, R., Formule de Taylor stochastique et développement asymptotique d'intégrales de Feynman. In Seminar on Probability, XVI, Supplement, volume 921 of Lecture Notes in Math., pp. 237–285, 1982 (Springer: Berlin-New York).
  • Azencott, R., Petites perturbations aléatoires des systemes dynamiques: développements asymptotiques. Bull. Sci. Math., 1985, 109(3), 253–308.
  • Baudoin, F. and Ouyang, C., On small time asymptotics for rough differential equations driven by fractional Brownian motions. In Large Deviations and Asymptotic Methods in Finance, edited by P.K. Friz, J. Gatheral, A. Gulisashvili, A. Jacquier, and J. Teichmann, pp. 413–438, 2015 (Springer International Publishing: Cham).
  • Bayer, C., Friz, P.K., Gassiat, P., Martin, J. and Stemper, B., A regularity structure for rough volatility. Preprint, 2017. arXiv:1710.07481.
  • Bayer, C., Friz, P.K. and Gatheral, J., Pricing under rough volatility. Quant. Finance, 2016, 16(6), 887–904. doi: 10.1080/14697688.2015.1099717
  • Ben Arous, G., Methods de Laplace et de la phase stationnaire sur l'espace de Wiener. Stochastics, 1988, 25(3), 125–153. doi: 10.1080/17442508808833536
  • Bennedsen, M., Lunde, A. and Pakkanen, M.S., Decoupling the short- and long-term behavior of stochastic volatility. Preprint, 2016. arXiv:1610.00332.
  • Bennedsen, M., Lunde, A. and Pakkanen, M.S., Hybrid scheme for Brownian semistationary processes. Finance Stoch., 2017, 21(4), 931–965. doi: 10.1007/s00780-017-0335-5
  • Bismut, J.-M., Large Deviations and the Malliavin Calculus. Progress in Mathematics, Vol. 45. 1984 (Birkhäuser Boston, Inc.: Boston, MA).
  • Carlen, E. and Kree, P., Estimates on iterated stochastic integrals. Ann. Probab., 1991, 19(1), 354–368. doi: 10.1214/aop/1176990549
  • Cass, T. and Friz, P., Densities for rough differential equations under Hörmander's condition. Ann. Math., 2010, 0, 2115–2141. doi: 10.4007/annals.2010.171.2115
  • Decreusefond, L., Stochastic integration with respect to Volterra processes. Ann. l. H. Poincare Probab. Statist., 2005, 41(2), 123–149. doi: 10.1016/j.anihpb.2004.03.004
  • Deuschel, J.-D., Friz, P.K., Jacquier, A. and Violante, S., Marginal density expansions for diffusions and stochastic volatility I: Theoretical foundations. Comm. Pure Appl. Math., 2014a, 67(1), 40–82. doi: 10.1002/cpa.21478
  • Deuschel, J.-D., Friz, P.K., Jacquier, A. and Violante, S., Marginal density expansions for diffusions and stochastic volatility II: Applications. Comm. Pure Appl. Math., 2014b, 67(2), 321–350. doi: 10.1002/cpa.21483
  • Deuschel, J.-D. and Stroock, D.W., Large Deviations, Vol. 137, 1989 (Academic Press: Boston, MA).
  • El Euch, O. and Rosenbaum, M., The characteristic function of rough Heston models. Preprint, 2016. To appear in Math. Finance.
  • Forde, M. and Jacquier, A., Small-time asymptotics for implied volatility under the heston model. Int. J. Theoret. Appl. Finance, 2009, 12(06), 861–876. doi: 10.1142/S021902490900549X
  • Forde, M. and Zhang, H., Asymptotics for rough stochastic volatility models. SIAM J. Financ. Math., 2017, 8(1), 114–145. doi: 10.1137/15M1009330
  • Friz, P.K. and Gassiat, P., Martingality and moments for lognormal rough volatility. In preparation, 2018.
  • Friz, P.K., Gerhold, S. and Pinter, A., Option Pricing in the Moderate Deviations Regime. Math. Finance, 2018, 28(3), 962–988. doi: 10.1111/mafi.12156
  • Friz, P. and Hairer, M., A Course on Rough Paths, 2014 (Springer: Cham).
  • Fukasawa, M., Asymptotic analysis for stochastic volatility: martingale expansion. Finance Stoch., 2011, 15(4), 635–654. doi: 10.1007/s00780-010-0136-6
  • Fukasawa, M., Short-time at-the-money skew and rough fractional volatility. Quant. Finance, 2017, 17(2), 189–198. doi: 10.1080/14697688.2016.1197410
  • Gao, K. and Lee, R., Asymptotics of implied volatility to arbitrary order. Finance Stoch., 2014, 18(2), 349–392. doi: 10.1007/s00780-013-0223-6
  • Gatheral, J., The Volatility Surface: A Practitioner's Guide, 2011 (John Wiley & Sons: Hoboken, NJ).
  • Gatheral, J., Jaisson, T. and Rosenbaum, M., Volatility is rough. Preprint, 2014. To appear in Quant. Finance.
  • Guennoun, H., Jacquier, A. and Roome, P., Asymptotic behaviour of the fractional Heston model. Preprint, 2014. arXiv:1411.7653.
  • Gulisashvili, A., Large deviation principle for Volterra type fractional stochastic volatility models. ArXiv e-prints, October 2017. To appear in SIAM J. Financ. Math.
  • Inahama, Y., Laplace approximation for rough differential equation driven by fractional brownian motion. Ann. Probab., 2013, 41(1), 170–205. doi: 10.1214/11-AOP733
  • Jacquier, A., Pakkanen, M.S. and Stone, H., Pathwise large deviations for the Rough Bergomi model. ArXiv e-prints, June 2017.
  • Jourdain, B., Loss of martingality in asset price models with lognormal stochastic volatility. Int. J. Theoret. Appl. Finance, 2004, 13, 767–787.
  • Lamperti, J., Semi-stable stochastic processes. Trans. Am. Math. Soc., 1962, 104(1), 62–78. doi: 10.1090/S0002-9947-1962-0138128-7
  • Lions, P.-L. and Musiela, M., Correlations and bounds for stochastic volatility models. Ann. Inst. H. Poincaré Anal. Non Linéaire, 2007, 24(1), 1–16.
  • Medvedev, A. and Scaillet, O., A simple calibration procedure of stochastic volatility models with jumps by short term asymptotics. Preprint, 2003. Available at SSRN 477441.
  • Medvedev, A. and Scaillet, O., Approximation and calibration of short-term implied volatilities under jump-diffusion stochastic volatility. Rev. Financ. Stud., 2007, 20(2), 427–459. doi: 10.1093/rfs/hhl013
  • Mijatović, A. and Tankov, P., A new look at short-term implied volatility in asset price models with jumps. Math. Finance, 2016, 26(1), 149–183. doi: 10.1111/mafi.12055
  • Muhle-Karbe, J. and Nutz, M., Small-time asymptotics of option prices and first absolute moments. J. Appl. Probab., 2011, 48(4), 1003–1020. doi: 10.1239/jap/1324046015
  • Olver, F.W.J., Lozier, D.W., Boisvert, R.F. and Clark, C.W (Eds.), NIST Handbook of Mathematical Functions, 2010 (Cambridge University Press: New York).
  • Osajima, Y., The asymptotic expansion formula of implied volatility for dynamic SABR model and FX hybrid model. Preprint, 2007. Available at SSRN 965265.
  • Osajima, Y., General asymptotics of Wiener functionals and application to implied volatilities, In Large Deviations and Asymptotic Methods in Finance, edited by P.K. Friz, J. Gatheral, A. Gulisashvili, A. Jacquier, and J. Teichmann, pp. 137–173, 2015 (Springer International Publishing: Cham).
  • Pham, H., Large deviations in mathematical finance, 2010. Available online at: https://www.lpsm.paris/pageperso/pham/GD-finance.pdf.
  • Sin, C.A., Complications with stochastic volatility models. Adv. Appl. Probab., 1998, 30(1), 256–268. doi: 10.1239/aap/1035228003

Appendix. Auxiliary lemmas

In this section we provide and prove some auxiliary lemmas, which are used in the preparations to the proof of Theorem 3.2. We start with a technical Lemma, that justifies the derivation.

Lemma .1

Assume σ(.)>0 and |ρ|<1. Then Kx is a Hilbert manifold near any h:=(h,f)KxH:=H01×H01.

Proof.

Similar to Bismut (Citation1984, p. 25) we need to show that Dϕ1(h) is surjective where ϕ1(h): HR with ϕ1h=ϕ1h,f=01σ(fˆ)dρ¯h+ρf. From ϕ1h+δh=01σ(fˆ+δfˆ)dρ¯h+ρf+δ(ρ¯h+ρf)=ϕ1h+δ01σ(fˆ)d(ρ¯h+ρf)+δ01σ(fˆ)fˆdρ¯h+ρf+oδ. the functional derivative Dϕ1(h) can be computed explicitly. In fact, even the computation Dϕ1h,h,0=ρ¯01σ(fˆ)dh is sufficient to guarantee surjectivity of Dϕ1(h).

We now give the proof of Lemma 6.3, which determines the form of the Girsanov measure change (Equation36) for the minimizing configuration.

Lemma A.2

(i) Any optimal control h0=(hx,fx)Kx is a critical point of h=h,fIϕ1h+12hH2; (ii) it holds that 01h˙xdW+01f˙xdB=Ixg1.

Proof.

(Step 1) Write h=(h,f) and ϕ1h=ϕ1h,f=01σ(fˆ)dρ¯h+ρf. Let h0=(hx,fx)Kx an optimal control. Then KerDϕ1h0=Th0Kx=hH1:Dϕ1h=0. (This requires Kx to be a Hilbert manifold near h0, as was seen in the last lemma.)

(Step 2) For fixed hH, define ut:=Iϕ1h0+th+12h0+thH20 with equality at t=0 (since x=ϕ1h0 and I(x)=12h0H2) and non-negativity for all t because h0+th is an admissible control for reaching x~=ϕ1h0+th (so that I(x~)=inf{}12h0+thH2.)

(Step 3) We note that u˙(0)=0 is a consequence of uC1 near 0, u(0)=0 and u0. In other words, h0 is a critical point for H1hIϕ1h+12hH2. (Step 4) The functional derivative of this map at h0 must hence be zero. In particular, for all hH,

0Iϕ1h0Dϕ1h0,h+h0,h=IxDϕ1h0,h+h0,h.

(Step 5) With h0=(hx,fx) and h=(h,f) Dϕ1h0,h=ddεε=001σ(fˆx+εfˆ)d×ρ¯hx+ρfx+ερ¯h+ρf=01σ(fˆx)dρ¯h+ρf+01σ(fˆx)fˆdρ¯hx+ρfx

By continuous extension, replace h=(h,f) by (W,B) above and note that Dϕ1h0,W,B=g1 since indeed g1=01σ(fˆt)d(ρ¯Wt+ρBt)+σ(fˆt)Bˆtd(ρ¯ht+ρft). Hence 01h˙xdW+01f˙xdB=Ixg1.