<<

A rough SABR formula

Masaaki Fukasawa Graduate School of Engineering Science, Osaka University 1-3 Machikaneyama, Toyonaka, Osaka, JAPAN [email protected]

Jim Gatheral Baruch College, City University of New York [email protected]

May 11, 2021

Abstract

Following an approach originally suggested by Balland in the context of the SABR model, we derive an ODE that is satisfied by normalized smiles for short maturities under a rough volatility extension of the SABR model that extends also the rough Bergomi model. We solve this ODE numerically and further present a very accurate approximation to the numerical solution that we dub the rough SABR formula.

1 Introduction

The now twenty years old SABR (stochastic alpha-beta-rho) model remains very popular arXiv:2105.05359v1 [q-fin.MF] 11 May 2021 amongst practitioners, particularly those in foreign exchange and interest rate markets. One key reason for the popularity of the SABR model is the existence of a closed- form approximation to the implied , the celebrated SABR formula of Hagan et al. [HKLW02]. SABR surfaces are, however, not really consistent with market data. In particular, the SABR implied volatility surface and its approximation by the SABR formula cannot reproduce the power-law type term structure of at-the-money (ATM) skew typically observed in markets. As a result, practioners are forced to use different SABR parameters for different maturities. Recently, a class of models where volatility is not a but has a rougher path, so called rough volatility models, have been shown to generate better fits to the observed volatility surface with very few parameters; see [ALV07, BFG16, Fuk17, FZ17] for more details. In particular, the rough Bergomi model of [BFG16] has only three parameters. In this article, we introduce a rough SABR model that includes both the SABR model and the rough Bergomi model as particular cases, and extend the SABR formula to the rough SABR model, demonstrating the accuracy of our formula using numerical simulation. The SABR formula is a small-time asymptotic approximation, and there are several approaches to its derivation, including [BBF04, Osa07]. Balland’s derivation [Bal06] of the lognormal SABR formula is a particularly simple and elegant one. Balland’s idea is to start with the drift condition for arbitrage-free implied volatility processes, and then to derive an ordinary differential equation (ODE) to be satisfied by normalized implied volatility smiles. In this article, we apply this idea to the rough SABR model to derive a modified ODE, which we solve numerically in the general case. We also provide an accurate closed- form approximation to this numerical solution, which we dub the rough SABR formula. In contrast to the classical SABR formula, the rough SABR formula generates a reasonable shape for the entire volatility surface, not just a single smile.

2 No arbitrage dynamics of the implied volatility

As a preliminary step, here we describe the drift condition for arbitrage-free dynamics of implied volatility processes. We treat both the Black-Scholes and Bachelier (normal) implied volatilities. We follow Balland [Bal06] for the Black-Scholes case, and apply the same idea to the Bachelier case. We also introduce the notion of asymptotically arbitrage- free approximation of the implied volatility.

Let S = {S t} be the underlying asset price process of an market, and C = BS BS B B {Ct(K, T)}, Ct(K, T) = P (S t, K, T −t, Σt ) = P (S t, K, T −t, Σt ) be a price pro- cess with strike K and maturity T, where PBS(S, K, τ, σ) and PB(S, K, τ, σ) are respectively the Black-Scholes and Bachelier call prices with volatility σ

S √ BS log K σ τ P (S, K, τ, σ) = S Φ(d+) − KΦ(d−), d± = √ ± , σ τ 2 √ ! !! B K − S 0 K − S 0 P (S, K, τ, σ) = σ τφ √ − (K − S 0) 1 − Φ √ σ τ σ τ We fix K > 0 and T > 0, and study the Black-Scholes and Bachelier implied volatility BS BS B B processes Σ = {Σt } and Σ = {Σt }. Proposition 2.1. Assume that S and C are continuous Itˆoprocesses, and that there is an equivalent measure Q for S and C. Then, Σ := ΣBS is a continuous Itˆo process and, denoting by D dt the drift part of dΣ under Q, we have d d d hlog S i + 2k hlog S, log Σi + k2 hlog Σi t t t d d d (2.1) d Σ4τ2 d = Σ2 − 2ΣτD − Σ2τ hlog S, log Σi + hlog Σi, dt 4 dt where τ = T − t and k = log K/S.

2 Proof. By Ito’sˆ formula, 1 1 dC − P dS = −P dt + P dΣ + P dhS i + P dhS, Σi + P dhΣi S τ σ 2 SS S σ 2 σσ and this is (the differential of) a local martingale under Q, where the are σ √ PS = Φ(d+), Pτ = S φ(d+) √ , Pσ = S φ(d+) τ, 2 τ √ φ(d+) φ(d+)d− S τ PSS = √ , PS σ = − , Pσσ = d+d−φ(d+). S σ τ σ σ Substituting these, we obtain ! Σ2 1 S Σ2τ − dt + ΣτDdt + dhlog S i − log − dhlog S, log Σi 2 2 K 2 ! 1 S 2 Σ4τ2 + log − dhlog Σi = 0. 2 K 4



In light of Proposition 2.1, we introduce the following notion.

Definition 2.1. A continuous Itˆoprocess Σˆ = {Σˆ t} is said to be an asymptotically arbitrage- free approximation of ΣBS under Q if, denoting by Dˆ dt the drift part of dΣˆ under Q, there exist a continuous function ϕ on R and a continuous process Ψ = {Ψt} on [0, T] such that

d d d hlog S i + 2k hlog S, log Σˆi + k2 hlog Σˆi dt dt dt (2.2) ˆ 4 2 2 2 d Σ τ d − Σˆ + 2ΣˆτDˆ + Σˆ τ hlog S, log Σˆi − hlog Σˆi ≤ ϕ(ΨΣˆ) · op(1) dt 4 dt as τ = T −t → 0, where k = log K/S and op(1) is a term which converges to 0 in probability.

Now we give a Bachelier version.

Proposition 2.2. Assume that S and C are continuous Itˆoprocesses, and that there is an equivalent local martingale measure Q for S and C. Then, Σ := ΣB is a continuous Itˆo process and, denoting by D dt the drift part of dΣ under Q, we have d d d hS i + 2k hS, log Σi + k2 hlog Σi = Σ2 − 2ΣτD, (2.3) dt dt dt where τ = T − t and k = K − S.

3 Proof. By Ito’sˆ formula, 1 1 dC − P dS = −P dt + P dΣ + P dhS i + P dhS, Σi + P dhΣi S τ σ 2 SS S σ 2 σσ and this is (the differential of) a local martingale under Q, where the (Bachelier) greeks are ! ! ! k σ k √ k PS = 1 − Φ √ , Pτ = √ φ √ , Pσ = τφ √ , σ τ 2 τ σ τ σ τ ! ! ! 1 k k k k2 k PSS = √ φ √ , PS σ = √ φ √ , Pσσ = √ φ √ . σ τ σ τ σ2 τ σ τ σ3 τ σ τ Substituting these, we obtain Equation (2.3). 

In light of Proposition 2.2, we introduce the following notion.

Definition 2.2. A continuous Itˆoprocess Σˆ = {Σˆ t} is said to be an asymptotically arbitrage- free approximation of ΣB under Q if, denoting by Dˆ dt the drift part of dΣˆ under Q, there exist a continuous function ϕ on R and a continuous process Ψ = {Ψt} on [0, T] such that

d d 2 d 2 hS i + 2k hS, log Σˆi + k hlog Σˆi − Σˆ + 2ΣˆτDˆ ≤ ϕ(ΨΣˆ) · op(1) (2.4) dt dt dt as τ = T − t → 0, where k = K − S and op(1) is a term which converges to 0 in probability.

It should be noted that the condition for an approximation to be an asymptotically arbitrage-free is necessary but not sufficient for the approximation to be reasonable. In particular, an asymptotically arbitrage-free approximation is not unique and is not neces- sarily accurate. Still, under the lognormal SABR model dS dα η = αdZ, = dW, (2.5) S α 2 where (Z, W) is a 2-dim correlated Brownian motion with dhZ, Wit = ρdt, and ρ ∈ (−1, 1) and η > 0 are constants, Balland [Bal06] found that η K Σˆ := α f (Y), Y = log α S is (in our terminology) an asymptotically arbitrage-free approximation of ΣBS if f is a solution of the ODE !2 ! f 0(y) y2 1 − y 1 + ρy + = f (y)2, f (y) 4 and that solving this ODE, the lognormal SABR formula of Hagan et al. [HKLW02] is obtained: p y 1 + ρy + y2/4 − ρ − y/2 f (y) = , g(y) = −2 log . g(y) 1 − ρ See also [AH13] for a related work. This simple and elegant approach to reach this accurate formula motivates us to seek an asymptotically arbitrage-free approximation to derive a useful formula under rough volatility models.

4 3 Implied volatility under rough SABR

Here we present our rough SABR model with an asymptotically arbitrage-free approxima- tion of the implied volatility. The model is

dS t dξt(s) = αt dZt, = κ(s − t)dWt, t < s (3.1) β(S t) ξt(s) p under an equivalent martingale measure Q, where αt = ξt(t), β is a positive continuous function,√ (Z, W) is a 2-dim correlated {Ft}-Brownian motion with dhZ, Wit = ρdt, κ(t) = H−1/2 η 2Ht , ρ ∈ [−1, 1], η > 0 and H ∈ (0, 1/2]. We assume {ξ0(s)}s≥0 to be a family of F0 measurable random variables and the curve s 7→ ξ0(s) to be continuous. Note that we have an explicit expression ( √ Z t ) Q H−1/2 1 2 2H 2H ξt(s) = E [ξs(s)|Ft] = ξ0(s) exp η 2H (s − u) dWu − η (s − (s − t) ) 0 2 (3.2) for 0 ≤ t ≤ s. The case H = 1/2 with

2 n 1 2 o ξ0(s) = α0 exp 4 η s is the classical SABR model; see (2.5) for the log normal case (β(s) = s). When β(s) = s, this is the rough Bergomi model introduced in [BFG16], and s 7→ ξt(s) is the forward variance curve at time t: Z s Z s "Z s # Q h 2 i Q ξt(u)ds = E αu|Ft du = E dhlog S i|Ft . t t t In general, ξ can be determined from weighted variance rates: Z s Z s h i "Z s S 2 # ξ u u EQ α2|F u EQ u h S i |F . t( )d = u t d = 2 d log u t t t t β(S u) See [FHT21] for the infinite dimensional of this model with application to hedging. Extending Balland [Bal06], we obtain the following result. Theorem 3.1. Let f be a solution of the ODE !2 ! ! f 0(y) y  y 2 y f 0(y) 1 − y 1 + 2ρ + = f (y)2 1 − (1 − 2H) (3.3) f (y) 2H + 1 2H + 1 f (y) with f (0) = 1. Let β(s) = s, that is, consider the rough Bergomi model. Then, Σˆ := U f (Y) is an asymptotically arbitrage-free approximation of ΣBS under Q, where s 1 Z T κ(T − t) K Ut = ξt(s)ds, Yt = log . T − t t Ut S t

5 The proof of Theorem 3.1 is given in Appendix A.

Remark 3.1 (Consistency with the asymptotic skew formula of [EFGR19]). Substituting a formal series expansion f (x) = 1 + ax + bx2/2 + ... to (3.3), we find ρ a = , 2(H + 1/2)(H + 3/2) consistent with the asymptotic skew formula for the rough Bergomi model we obtained in [EFGR19]. That said, the beta function coefficient that appears in the curvature formula of [EFGR19] precludes consistency with the second order coefficient

(2H + 3)2 − 12(2H + 1)ρ2 b = . 2(H + 1)(2H + 1)2(2H + 3)2

Replacing the role of log S by S in the proof of Theorem 3.1, we obtain the following Bachelier version.

Theorem 3.2. Let f be a solution of the ODE (3.3). Let β(s) = 1. Then,

Σˆ := U f (Y) is an asymptotically arbitrage-free approximation of ΣB under Q, where s Z T 1 K − S t Ut = ξt(s)ds, Yt = κ(T − t) . T − t t Ut

According to [BBF02], for a model dS t = σ(S t) dZt, we have

log K ΣBS ≈ S . R K ds S σ(s) In particular for the Bachelier model σ(s) = σ, we have

log K ΣBS ≈ σ S K − S that connects the Black-Scholes and Bachelier volatility parameters. Combining this and Theorem 3.2, we obtain an approximation formula for our model with β(s) = 1:

log K U  K − S  K ΣBS ≈ U f (Y) S = f κ(τ) log . K − S K − S U S This further suggests a formula for general β:

U  X  K Z K ds ΣBS ≈ f κ(τ) log , X = . (3.4) X U S S β(s)

6 4 Solving the ODE

Now we study the solution of the ODE (3.3). For g(y) = y/ f (y), from (3.3), we have ! ! y y2 yg0(y) g0(y)2 1 + 2ρ + = 1 − (1 − 2H) 1 − (4.1) 2H + 1 (2H + 1)2 g(y) with g(y)/y → 1 as y → 0. When, H = 1/2, this is solvable and p 1 + ρy + y2/4 − ρ − y/2 g(y) = −2 log (4.2) 1 − ρ which gives the familiar SABR formula. For H = 0, we also have the explicit solution tv   y 2ρ  ρ y + ρ  g(y) = log(1 + 2ρy + y2) + p arctan p − arctan p . (4.3) |y| 1 − ρ2  1 − ρ2 1 − ρ2 

We plot the resulting solutions f (y) = y/g(y) for various values of ρ in Figure 4.1.

Figure 4.1: The function f for H = 1/2 (in red) and H = 0 (in blue).

For general H, the ODE has to be solved numerically. By (4.1), we have p (1 − 2H) f (y) + (1 − 2H)2 f (y)2 + 8Hq(y) g0(y) = , g(0) = 0 2q(y) where y y y2 f (y) = , f (0) = 1, q(y) = 1 + 2ρ + . g(y) 2H + 1 (2H + 1)2 There is no difficulty in obtaining a numerical solution for such a one dimensional first order ODE. See Figure 4.2 for numerical examples of f functions for various values of H when ρ = 0 (right) and ρ = −0.9 (left).

7 Figure 4.2: The function f : numerical solutions for various values of H

5 A closed-form approximation

In order to derive a closed-form approximation, it turns out to be convenient to recast (4.1) in terms of G(y):= g(y)2. We find ( ) 1 y2 2 ρ y G0(y)2 + + 1 = (1 − 2H) y 1G0(y) + 2HG(y). (5.1) 4 (2H + 1)2 2H + 1 2 with initial condition G(0) = 0. Denote the solutions of (5.1) with H = 0 and H = 1/2 respectively by G0(·) and G1/2(·). Then from (4.2) and (4.3),

 p 2  1 + ρy + y2/4 − ρ − y/2 G (y) = 4 log  1/2  1 − ρ    ρ ρ y ρ 2 2  +  G0(y) = log(1 + 2ρy + y ) + p arctan p − arctan p  . 1 − ρ2  1 − ρ2 1 − ρ2 

Substituting G(y) = y2 + a y3 + b y4 + ... into (5.1) and matching coefficients gives ! ρ 1 4γ + 1 a = − , b = 3 ρ2 − 1 . γ(γ + 1) 4γ2 (2γ + 1) (γ + 1)2

8 1 where γ = H + 2 . To the same order, we have ρy3 1   G (y) = y2 − + 15ρ2 − 4 y4 + ... 1/2 2 48 4ρy3 1   G (y) = y2 − + 4ρ2 − 1 y4 + .... 0 3 2 Matching coefficients of y2 and y3, we arrive at the following interpolation of the extreme solutions G0 and G1/2, which by construction gives the correct ATM skew: ( !) 3 (1 − 2H)  y  2H 2y G (y) = (2H + 1)2 G + G . (5.2) A 2H + 3 0 2H + 1 2H + 3 1/2 2H + 1

Obviously, the approximate solution GA(y) agrees with G0(y) when H = 0 and with G1/2(y) when H = 1/2. To give a sense for the accuracy of the approximate solution is for general H, in Figure 5.1 we plot the numerical solution f versus the approximation p fA(y):= |y|/ GA(y) for two values of H: H = 0.05 which is a typical calibrated value, and H = 0.25 which should approximately maximize the approximation error.

Figure 5.1: The function f and its approximation fA for two values of ρ. H = 0.05 is in blue, H = 0.25 is in red; solid line is the numerical solution f and dashed, the approximation fA(y).

6 Final formula with Monte-Carlo comparison

Let τ = T − t as before and s κ(τ) x K Z K ds 1 Z t+τ y(x, τ) = , k = log , kβ = , Ut(τ) = ξt(s)ds. U(τ) S S β(s) τ t

9 Then (3.4) and (5.2) suggest the following approximate rough SABR formula for the Black- Scholes implied volatility Σ(k, τ) of an option with time to τ and log-strike k:

|y(k, τ)| Σ(k, τ) = Σ(0, τ) p (6.1) GA(y(kβ, τ)) under the rough SABR model (3.1).

In order to confirm the accuracy of (6.1) in the lognormal√ case β(s) = s, we simulate the rough Bergomi model [BFG16] with kernel κ(τ) = η 2H τH−1/2 using the hybrid scheme [BLP17, FH21]1 with 1 million paths, 210 = 1, 024 time steps for H ∈ {0.10, 0.20} and 213 = 8, 192 time steps for the case H = 0.05.

We take η = 1 and a flat forward variance curve, ξ0(s) = 0.04. In Figures 6.1, 6.2, and 6.3, for H ∈ {0.05, 0.1, 0.2}, we plot smiles for ρ ∈ {−0.9, −0.6, 0, 0.6}. Specifically, each subplot has the graph of ! ! Σ(k, τ) κ(τ)k Σ(k, τ) y(k, τ), = , , Σ(0, τ) 0.2 Σ(0, τ)

In each case, the dashed red curve (“rSABR”) is the function f obtained by solving the ODE (3.3) numerically. From these plots, we first notice that the scaling of y(k, τ) works remarkably well to offset the maturity dependence of the normalized smile. Also, note in particular that the quality of our approximation decreases as H decreases. √ As a further experiment, we simulate the rough SABR model (3.1) with β(s) = s with 1 million paths and 212 = 4, 096 time steps for H = 0.05 and ρ = −0.9. Again, we take η = 1 and a flat forward variance curve, ξ0(s) = 0.04. The quality of the rough SABR formula (6.1) is demonstrated in Figure 6.4. Here we plot against k on the x-axis, rather than y(k, τ). Through these numerical experiments, we observe that for small values of H such as H = 0.05 with η = 1, we need as many as 212 time steps to achieve convergence. A compu- tation with so many time steps is obviously extremely time-consuming, as if to emphasize to us the value of analytical approximations. We emphasize that our rough SABR formula (6.1) is a short-dated approximation, as is the original SABR formula of [HKLW02]. As with the classical SABR formula, we can only expect the approximation to work well for values of the dimensionless expansion parameter η τH < 1 and indeed this is what we find in our extensive numerical experiments. For more extreme rough Bergomi parameter sets such as the pair η = 2.3, H = 0.05 arising from calibration of the rough Bergomi model to the SPX surface [BFG16], the approximate formula (6.1) does not match the Monte Carlo smile sufficiently well for it to be useful in model calibration. 1Implementation of the refinement of [FH21] seems to make a significant difference in resolving the smile for out-of-the-money calls. Here we set the parameters κ = 2 and κ0 = n for the scheme of [FH21].

10 Figure 6.1: With β(s) = s and parameters H = 0.05, η = 1, the dashed red line is the nu- merical solution f ; Monte Carlo estimates of normalized implied volatility Σ(k, τ)/Σ(0, τ) for τ = 1, 3, 6, and 12 months are as in the legend.

11 Figure 6.2: With β(s) = s and parameters H = 0.10, η = 1, the dashed red line is the nu- merical solution f ; Monte Carlo estimates of normalized implied volatility Σ(k, τ)/Σ(0, τ) for τ = 1, 3, 6, and 12 months are as in the legend.

12 Figure 6.3: With β(s) = s and parameters H = 0.20, η = 1, the dashed red line is the nu- merical solution f ; Monte Carlo estimates of normalized implied volatility Σ(k, τ)/Σ(0, τ) for τ = 1, 3, 6, and 12 months are as in the legend.

13 √ Figure 6.4: With β(s) = s and parameters H = 0.05, η = 1, Monte Carlo estimates of normalized implied volatility Σ(k, τ)/Σ(0, τ) for τ = 1.5, 3, 6, and 12 months are as in the legend. Dashed lines are corresponding plots of the rough SABR formula (6.1).

7 Summary and conclusions

Following an approach originally suggested by Balland [Bal06] in the context of the clas- sical SABR model, we derived an ODE that is satisfied by the rough SABR normalized volatility smile for short maturities. We solved this ODE numerically and further derived a very accurate approximation to the numerical solution. The resulting analytical formula 1 coincides with the classical one in the case H = 2 . Numerical simulation of rough Bergomi smiles confirms that our small time approximation works well for values of the expansion parameter η τH < 1. In FX and interest rate applications, parameters of the classical SABR formula are allowed to depend on time to expiration; effectively a different model for each expiration. In our setup, effective classical SABR parameters also depend on time to expiration, according to the chosen value of H. Our conjecture is that the time dependence of market-implied classical SABR parameters can be parameterized by H – confirmation of this is left for further research. If so, given that our rough SABR formula is hardly more complicated than the classical SABR formula, it has the potential to be widely adopted by practitioners.

14 References

[AH13] Jesper Andreasen and Brian Huge. Expanded forward volatility. Risk, 26(1):101–107, 2013. [ALV07] Elisa Alos,` Jorge A Leon,´ and Josep Vives. On the short-time behavior of the implied volatility for jump-diffusion models with stochastic volatility. Finance and Stochastics, 11(4):571–589, 2007. [Bal06] Philippe Balland. Forward smile. Presentation at Global Derivatives, Paris, September 2006. [BBF02] Henri Berestycki, Jer´ omeˆ Busca, and Igor Florent. Asymptotics and calibra- tion of local volatility models. Quantitative Finance, 2:61–69, 2002-02. [BBF04] Henri Berestycki, Jer´ omeˆ Busca, and Igor Florent. Computing the implied volatility in stochastic volatility models. Communications on Pure and Ap- plied Mathematics, 57:1–22, 2004. [BFG16] Christian Bayer, Peter Friz, and Jim Gatheral. Pricing under rough volatility. Quantitative Finance, 16(6):887–904, 2016. [BLP17] Mikkel Bennedsen, Asger Lunde, and Mikko S Pakkanen. Hybrid scheme for Brownian semistationary processes. Finance and Stochastics, 21(4):931–965, 2017. [EFGR19] Omar El Euch, Masaaki Fukasawa, Jim Gatheral, and Mathieu Rosenbaum. Short-term at-the-money asymptotics under stochastic volatility models. SIAM Journal on Financial Mathematics, 10(2):491–511, 2019. [FH21] Masaaki Fukasawa and Asuto Hirano. Refinement by reducing and reusing random numbers of the hybrid scheme for Brownian semistationary processes. Quantitative Finance, available online, 2021. [FHT21] Masaaki Fukasawa, Blanka Horvath, and Peter Tankov. Hedging under rough volatility. arXiv:2105.04073 [q-fin], 2021. [Fuk17] Masaaki Fukasawa. Short-time at-the-money skew and rough fractional volatility. Quantitative Finance, 17(2):189–198, 2017. [FZ17] Martin Forde and Hongzhong Zhang. Asymptotics for rough stochastic volatility models. SIAM Journal on Financial Mathematics, 8(1):114–145, 2017. [HKLW02] Patrick S Hagan, Deep Kumar, Andrew S Lesniewski, and Diana E Woodward. Managing smile risk. Wilmott Magazine, pages 84–108, 2002. [Osa07] Yasufumi Osajima. The asymptotic expansion formula of implied volatility for dynamic SABR model and FX hybrid model. papers.ssrn.com, 2007.

15 A Proof of Theorem 3.1

Proof. Denote τ = T − t and k = log K/S . By Ito’sˆ formula, we have ! ! dU 1 1 1 α2 1 = κ(τ) R dW + 1 − − κ(τ)2R2 dt, U 2 2 τ U2 4 where R T κ s − t ξ s s t ( ) t( )d Rt = R T . κ T − t ξ s s ( ) t t( )d Further, we have κ0(τ) κ(τ) 1 1 dY = − Ydt + dk + κ(τ)kd + κ(τ)dh , ki κ(τ) U U U Y κ(τ) dU Y κ(τ) = (1 − 2H)dt + dk − Y + dhUi − hU, ki 2τ U U U2 U2 ! ! Y α2 κ(τ)2 3 α κ(τ) Y = − 2H dt + YR2 + Rρ dt + dk − κ(τ)RdW 2τ U2 2 4 U U 2 Y α = − κ(τ)RdW − κ(τ) dZ + drift, 2 U and so, 1 f 0(Y) d log Σˆ = κ(τ)RdW + dY + drift, 2 f (Y) ! 1 Y f 0(Y) f 0(Y) α = κ(τ)R 1 − dW − κ(τ) dZ + drift. 2 f (Y) f (Y) U

This implies d hlog Σˆi dt  !2 !2 !  R2 Y f 0(Y) f 0(Y) α Y f 0(Y) f 0(Y) ρα = κ(τ)2  1 − + − R 1 −   4 f (Y) f (Y) U f (Y) f (Y) U  and ! d ρα Y f 0(Y) f 0(Y) α2κ(τ) hlog S, log Σˆi = κ(τ)R 1 − − . dt 2 t f (Y) f (Y) U

16 Therefore, d d d hlog S i + 2k hlog S, log Σˆi + k2 hlog Σˆi dt dt dt ( ! ) ρα Y f 0(Y) f 0(Y) α2 =α2 + 2kκ(τ) R 1 − − 2 f (Y) f (Y) U  !2 !2 !  R2 Y f 0(Y) f 0(Y) α Y f 0(Y) f 0(Y) ρα + k2κ(τ)2  1 − + − R 1 −   4 f (Y) f (Y) U f (Y) f (Y) U  ( ( ! ) α2 α ρR Y f 0(Y) f 0(Y) α =U2 + 2Y 1 − − U2 U 2 f (Y) f (Y) U  !2 !2 ! ) R2 Y f 0(Y) f 0(Y) α Y f 0(Y) f 0(Y) α  + Y2  1 − + − 1 − ρR  4 f (Y) f (Y) U f (Y) f (Y) U  ( ( ! )!2 !2) α ρR Y f 0(Y) f 0(Y) α R2 Y f 0(Y) =U2 + Y 1 − − + (1 − ρ2) Y2 1 − U 2 f (Y) f (Y) U 4 f (Y) !2 ( ) Y f 0(Y) α2 α R2 =U2 1 − + ρR Y + Y2 . f (Y) U2 U 4

Now we consider the second line of (2.2). From the above computations, we see that the last two terms are of O(τκ(τ)) = O(τH+1/2). For the second term, since 1 dΣˆ = f (Y)dU + U f 0(Y)dY + U f 00(Y)dhYi + f 0(Y)dhU, Yi, 2 we have ! ! 1 α2 1 Dˆ =U f (Y) 1 − B − κ(τ)2R2 2τ U2 8 ! !! Y α2 κ(τ)2 3 α + U f 0(Y) − 2H + YR2 + Rρ 2τ U2 2 4 U ! ! 1 α2 α Y2R2 YR2 α Rρ + U f 00(Y)κ(τ)2 + YRρ + − U f 0(Y)κ(τ)2 + 2 U2 U 4 4 U 2 !! 1 α2 Y f 0(Y) α2 = U f (Y) 1 − + − 2H 2τ U2 f (Y) U2 ! !! 1 R2 Y f 0(Y) α2 α R2 + Uκ(τ)2 − f (Y) 1 − + f 00(Y) + ρR Y + Y2 . 2 4 f (Y) U2 U 4

Therefore, we have a nonnegligible term ! ! ! Y f 0(Y) α2 Y f 0(Y) 2ΣˆτDˆ ≈ U2 f (Y)2 1 − 1 − + (1 − 2H) . f (Y) U2 f (Y)

17 Now, using that f is a solution of (3.3), d d d hlog S i + 2k hlog S, log Σˆi + k2 hlog Σˆi dt dt dt d Σˆ 4τ2 d − Σˆ 2 + 2ΣˆτDˆ + Σˆ 2τ hlog S, log Σˆi − hlog Σˆi dt 4 dt !2 ! !! Y f 0(Y) α2 α 1 Y2 1 ≈ U2 1 − −1 + + ρY R − + R2 − f (Y) U2 U H + 1/2 4 (H + 1/2)2 ! ! Y f 0(Y) α2 + U2 f (Y)2 1 − 1 − . f (Y) U2

The result then follows from the lemma below, by taking Ψ = 1/U so that ΨΣˆ = f (Y).  Lemma A.1. As τ = T − t → 0,

αt 1 → 1, Rt → Ut H + 1/2 in probability.

Proof. By the explicit expression (3.2), we have Z t Z t ξt(s) ξ0(s) 1 h 2 2i log = log + [κ(s − u) − κ(t − u)] dWu − κ(s − u) − κ(t − u) du ξt(t) ξ0(t) 0 2 0 for t ≤ s ≤ T. Now, as τ = T − t → 0, Z t h i h i κ(s − u)2 − κ(t − u)2 du = η2 s2H − (s − t)2H − t2H → 0, 0 and Z t [κ(s − u) − κ(t − u)]2 du 0 " Z t # = η2 s2H − (s − t)2H + t2H − 4H (s − u)H−1/2(t − u)H−1/2du) 0 " Z 1 # = η2 s2H − (s − t)2H + t2H − 4H t2H (s/t − 1 + x)H−1/2 xH−1/2dx) → 0. 0

Also, the initial forward variance curve is continuous by assumption, so ξt(s)/ξt(t) → 1. Therefore, U2 Z 1 ξ (t + τθ) t t → 2 = dθ 1, αt 0 ξt(t) and R 1 H−1/2 Z 1 θ ξt(t + τθ) dθ 1 R 0 → θH−1/2 θ . t = R 1 d = ξ (t + τθ) dθ 0 H + 1/2 0 t 

18