Sepp Rakhmonov 2024 Log Normal Stochastic Volatility Model With Quadratic Drift
Sepp Rakhmonov 2024 Log Normal Stochastic Volatility Model With Quadratic Drift
2450003
OPEN ACCESS
International Journal of Theoretical and Applied Finance
Vol. 26, No. 8 (2023) 2450003 (63 pages)
c The Author(s)
DOI: 10.1142/S0219024924500031
ARTUR SEPP ∗
Clearstar Labs AG
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
PARVIZ RAKHMONOV
Marex
155 Bishopsgate, London EC2M 3TQ, United Kingdom
[email protected]
In this paper, we introduce the log-normal stochastic volatility (SV) model with a
quadratic drift to allow arbitrage-free valuation of options on assets under money-market
account and inverse martingale measures. We show that the proposed volatility process
has a unique strong solution, despite non-Lipschitz quadratic drift, and we establish
the corresponding Feynman–Kac partial differential equation (PDE) for computation of
conditional expectations under this SV model. We derive conditions for arbitrage-free
valuations when return–volatility correlation is positive to preclude the “loss of mar-
tingality”, which occurs in many traditional SV models. Importantly, we develop an
analytic approach to compute an affine expansion for the moment generating function of
the log-price, its quadratic variance (QV) and the instantaneous volatility. Our solution
allows for semi-analytic valuation of vanilla options under log-normal SV models closing
a gap in existing studies.
We apply our approach for solving the joint valuation problem of vanilla and inverse
options, which are popular in the cryptocurrency option markets. We demonstrate the
accuracy of our solution for valuation of vanilla and inverse options.a By calibrating
the model to time series of options on Bitcoin over the past four years, we show that
the log-normal SV model can work efficiently in different market regimes. Our model can
∗ Corresponding author.
This is an Open Access article published by World Scientific Publishing Company. It is distributed
under the terms of the Creative Commons Attribution 4.0 (CC BY) License which permits use,
distribution and reproduction in any medium, provided the original work is properly cited.
a Github project https://github.com/ArturSepp/StochVolModels provides Python code with the
implementation of option valuation using the affine expansion and examples of model calibration
to implied volatility data applied in this paper.
2450003-1
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
be well applied for modeling of implied volatilities of assets with positive return–volatility
correlation.
1. Introduction
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
Empirical studies strongly support evidence that the volatility of price returns is
itself a stochastic process (see, for example, Shephard (2005) for a comprehensive
survey). It is accepted that the celebrated model by Black & Scholes (1973) and
Merton (1973), which assumes constant volatility, cannot explain implied volatil-
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
ity surfaces observed in option markets, which are inhomogeneous in strike and
maturity dimensions. In contrast, stochastic volatility (SV) models are able to fit
to implied volatility surfaces and their dynamics.
b Exp-OU SV models are studied in Fouque et al. (2000), Detemple & Osakwe (2000), Masoliver
& Perelló (2006), Perelló et al. (2008), Muhle-Karbe et al. (2022), Drimus (2012). Exp-OU SV
models are widespread in industry as a basis for local SV models (Lipton (2002)), see Bergomi
(2015), Bühler (2006), Ren et al. (2007), Henry-Labordère (2009), Bergomi & Guyon (2012).
2450003-2
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
additional state variable. Authors then approximate the payoff function using an
orthogonal polynomial expansion, see Ackerer & Filipović (2020), for option pricing
applications. A related approach for computing moments under SV models using
expansions is developed by Alòs et al. (2020).
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
2450003-3
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(for VIX index, short and leveraged short exchange-traded funds (ETFs)) or on a
transitory regime-based basis (for some commodities, foreign currencies, and cryp-
tocurrencies). In fact, exp-ou and log-normal SV models with the linear drift lead to
the “loss of martingality” of the price process when the return–volatility correlation
is positive (see, for example, Lewis (2000, 2018) and Lions & Musiela (2007)). We
contribute by establishing conditions on model parameters of log-normal SV model
which produce true martingale dynamics.
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
tage of inverse payoffs is that all option-related transactions can be handled using
units of underlying cryptocurrencies, such as Bitcoin or Ether, without using fiat
currencies (see Lucic (2021) and Lucic & Sepp (2023) for a connection between the
valuation of the inverse cryptocurrency options and FX options). In case of cryp-
tocurrencies, cash settled options would require using a stablecoin as a tradable
equivalent of United States Dollar (USD) or any other fiat currency. Given inherent
and frequent depegging risk of stablecoins, Alexander et al. (2023) argue for wider
adoptation of inverse and quanto options. We also suggest that similar payoff can
be advantageous for some commodity markets (such as gold) where option holders
may opt for physical settlement of positive option premium through the delivery of
the underlying commodity.
We contribute to the literature by incorporating the inverse measure in our
valuation framework using log-normal SV model. We formulate and solve the joint
valuation problem of vanilla and inverse options using Fourier transform methods.
2450003-4
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
at maturity T as follows:
1 1
rcall (ST ) = max{ST − K, 0}, rput (ST ) = max{K − ST , 0}. (2.2)
ST ST
We assume that the market is complete and satisfies the so-called No Free Lunch
with Vanishing Risk (NFLVR) and No Dominance (ND) conditions (see Jarrow &
Larsson (2012)). By choosing M as a numéraire, we consider an equivalent martin-
gale measure Q, induced by M . Accordingly, the time-t value of an option, denoted
by U (t, S), with payoff function u(ST ) at time T , equals
1
U (t, S) = M (t)E u(ST )|Ft , (2.3)
M (T )
where expectation E is taken under the MMA martingale measure Q.
Next, by choosing spot price S as a numéraire, we consider an inverse martingale
measure Q̃, induced by S. The option value function, denoted by Ũ (t, S), with the
payoff paid in units of S, equals
1
Ũ (t, S) = St Ẽ u(ST )|Ft , (2.4)
ST
where expectation Ẽ is taken under the inverse martingale measure Q̃.
Theorem 2.1. We assume that market is complete and satisfies NFLVR and ND
conditions. We further assume that both MMA measure Q and inverse measure Q̃
2450003-5
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
are equivalent martingale measures. Then the values of options under MMA measure
in Eq. (2.3) and under inverse measure in Eq. (2.4) are unique and satisfy
1 1
M (t)E u(ST )|Ft = St Ẽ u(ST )|Ft . (2.5)
M (T ) ST
Proof. As shown in Theorem 3.2 of Jarrow & Larsson (2012) and Theorem 2.17
of Herdegen & Schweizer (2018), the market satisfies NFLVR and ND conditions
if and only if Q is a true martingale measure. Hence, asset prices, discounted by
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
2450003-6
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proof. We only provide a sketch of the proof and refer to Theorems 9.2 and 9.3 in
Lewis (2000) and to Sin (1998) for details. As vanilla call option payoff is unbounded,
the price of the call option is only local martingale under Q. Taking into account
“correction” term (so-called martingale defect) reflecting the difference between true
martingale and strict local martingale properties, yields Eq. (2.8). Relationship in
Eq. (2.9) follows from the same argument, but applied to Q̃.
In Theorem 3.2, we show that volatility process σt does not explode under the
proposed log-normal SV dynamics in Eq. (3.12).
where μt is statistical drift, W (0) , W (1) are uncorrelated Brownian motions under
P, κ1 > 0 and κ2 ≥ 0 are linear and quadratic mean-reversion rates, respectively,
θ > 0 is the mean of the volatility, β ∈ R is the volatility beta which measures the
sensitivity of the volatility to changes in the spot price, and ε > 0 is the volatility
of residual volatility.
Empirically, Bakshi et al. (2006) find that the presence of nonlinear terms,
including the quadratic term, in the volatility drift is significant when using econo-
metric estimation of the dynamics of the VIX index.
2450003-7
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
where E(·) denotes Itô exponential, and λ0 (t) and λ1 (t) are equity and volatility
risk-premias, respectively.
Assuming that process ζ(t) is a true martingale, the following processes
t t
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
dσt = [κ1 θ − (κ1 − κ2 θ)σt − κ2 σt2 − (βλ0 (t) + ελ1 (t))σt ]dt
(0) (1)
+ βσt dŴt + εσt dŴt , (3.5)
where λ̄0 and λ̄1 are constants. As a result, we rewrite Eq. (3.5) as follows:
dσt = κ1 θ − (κ1 − κ2 θ)σt − (κ2 + β λ̄0 + ελ̄1 )σt2 dt
(0) (1)
+ βσt dŴt + εσt dŴt . (3.7)
To retain the functional form of the volatility process, we rescale model param-
eters θ̂, κ̂1 , κ̂2 under Q as follows:
c Such linear specification is the simplest functional form that allows to establish simple relationship
between gains on a delta-hedged option portfolio and level of the market volatility. As shown in
Bakshi & Kapadia (2003), such form allows to get precise information about the sign and strength
of the risk premium.
2450003-8
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Thus, under martingale measure Q, the dynamics of price and volatility pro-
cesses become
(0)
dSt = rt St dt + σt St dŴt ,
(3.10)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
(0) (1)
dσt = (κ̂1 + κ̂2 σt )(θ̂ − σt ) dt + βσt dŴt + εσt dŴt .
2450003-9
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(0)
dZt = Zt σt dWt , Z 0 = S0 ,
(3.15)
(0)
dRt = σt2 Rt dt − Rt σt dWt , R0 = 1/Z0 .
Next, we introduce the log of the zero-drift price process Xt = ln Zt :
1 (0)
dXt = − σt2 dt + σt dWt , X0 = ln S0 . (3.16)
2
Using XT under Assumption 2.1, we represent the spot price ST by
ST = er̄(0,T ) eXT . (3.17)
Similarly, under Assumption 2.2, using Eq. (3.17), we represent the futures settled
at time T in Q as
fT (T ) = er̄(0,T )−q̄(0,T ) eXT , (3.18)
where the integrated convenience rate q̄(0, T ) is defined in Eq. (2.6). Accordingly,
for modeling purposes, we focus on the zero-drift price dynamics Zt and its log-price
process Xt defined in Eq. (3.16).
2450003-10
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Karatzas & Shreve (1991) for details. We will establish the existence of the unique
strong solution in Theorem 3.3. We first provide a boundary classification for the
volatility process to show that log-normal volatility process σt in dynamics (3.19)
cannot reach 0 or explode to +∞, which are important properties of the volatility
dynamics. Our proof is based on existence of solutions for general stochastic differen-
tial equations (SDEs) with so-called monotonic coefficients, so that it is potentially
applicable to wider class of diffusion processes.
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
κ1 ≥ 0, κ2 ≥ 0, θ > 0, ϑ ≥ 0. (3.20)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Theorem 3.2 (Regularity). Under Assumption 3.1, the boundary points {0, +∞}
are unattainable for process σt in SDE (3.19).
Proof. We will refer to following theorem that guarantees the existence and unique-
ness of general SDEs with monotonic coefficients.
Theorem 3.4 (Gyöngy & Krylov (1980)). Let at (x) : [0, T ] × R → R, bt (x) :
[0, T ] × R → R be functions, continuous in x, satisfying following conditions:
T
(1) 0 sup|x|≤R (|at (x)| + |bt (x)|2 )dt < ∞ for any T ≥ 0, R ≥ 0,
(2) for x, y, z ∈ R such that |x| ≤ R, |y| ≤ R, |z| ≤ R following conditions hold:
(a) (monotonicity condition) :
2(x − y)(at (x) − at (y)) + (bt (x) − bt (y))2 ≤ Kt (R)(x − y)2 , (3.21)
We show that SDE (3.19), where κ1 , κ2 , θ, ϑ satisfy (3.20), does indeed satisfy
conditions 1–3. We emphasize that the positivity of the process demonstrated in
2450003-11
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Theorem 3.2 is important. Since in our case at (x) = (κ1 + κ2 x)(θ − x), bt (x) = ϑx,
we get
2(x − y)(at (x) − at (y)) + (bt (x) − bt (y))2
= (ϑ2 − 2(κ1 − κ2 θ))(x − y)2 − 2κ2 (x + y)(x − y)2
≤ (ϑ2 − 2(κ1 − κ2 θ) + 2κ2 R)(x − y)2
t
when |x| ≤ R, |y| ≤ R. As for any R > 0 we have 0 [ϑ2 −2(κ1 −κ2 θ)+2κ2 R]ds < ∞,
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
3
z < 0 valid for z > 0, we can estimate the right hand side of (3.23) as follows:
2κ1 θ(z 2 + 1) + [ϑ2 − 2(κ1 − κ2 θ)]z 2 ≤ 2κ1 θ + [2κ1 θ + ϑ2 − 2(κ1 − κ2 θ)]z 2 .
We conclude that the growth condition 2b is also satisfied.
Proof. Let us first consider the case p ≥ 1. We use same approach as in Lions &
Musiela (2007). Applying Itô’s lemma to σtp , we have
ϑ2
dσtp = pσtp−1 κ1 θ − κ1 − κ2 θ − p(p − 1) σt − κ2 σt2 dt + pϑσtp dWt .
2
The function U (τ ) = E(σtp ) then solves the problem
∂U
− + L(σ) U = 0, U (0, σ) = σ p ,
∂τ
(3.24)
∂ 1 2 2 ∂2
L = (κ1 θ − (κ1 − κ2 θ)σ − κ2 σ )
(σ) 2
+ ϑ σ .
∂σ 2 ∂σ 2
Let us find supersolution Ū to (3.24) using ansatz Ū (σ) = σ p exp(aσ + bτ ), a > 0,
b > 0. Inserting it into (3.24), we see that its left-hand side equals
ϑ2
σ p exp(aσ + bθ) −b + aκ1 θ − p(κ1 − κ2 θ) + p(p − 1) + pκ1 θσ −1
2
ϑ2
+ (pϑ2 − a(κ1 − κ2 θ) − κ2 p)σ + −κ2 a + a2 σ2 .
2
2450003-12
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
∂ 1 ∂2
L̃ = (κ2 + (κ1 − κ2 θ + ϑ )σ̃ − κ1 θσ̃ )
(σ) 2
+ ϑ2 σ̃ 2 2 . 2
∂ σ̃ 2 ∂ σ̃
Inserting ansatz into (3.25), we see that its left-hand side equals
σ̃ q−1 exp(aσ̃ + bθ)[−b + (· · ·) + (· · ·)σ̃ + (−2aκ1 θ + a2 ϑ2 )σ̃],
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
where we omitted the terms not depending on b and σ̃ for brevity. Since the term in
square brackets is a second-order polynomial in σ̃, Ũ is a supersolution if κ1 θ > 0.
t
+ pϑE sup σup dWu(∗) .
t∈[0,T ] 0
Theorem 3.5. The dynamics of the volatility σt in (3.12) under Q̃ are given by
t(0) + εσt dW
dσt = (κ1 θ − (κ1 − κ2 θ)σt − (κ2 − β)σt2 )dt + βσt dW t(1) . (3.26)
MMA measure Q and inverse measure Q̃ are related by the density process
Λt = Et [dQ̃/dQ] = Zt /Z0 , Λ0 = 1, (3.27)
where zero-drift stochastic driver Zt is defined in Eq. (3.15).
2450003-13
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
and Eq. (3.26) follows. We note that Q̃ and Q are related by the density Λt as
follows:
t
1 t 2
dΛt /Λt = σt dWt ⇒ Λt := exp σs dWs − σ ds = Zt /Z0 . (3.29)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
0 2 0 s
We highlight that the idea of using change of measure to study the martingale
property is widely used in the literature, see Ruf (2015) and references therein. We
now derive sufficient conditions for the equivalence of measures Q ∼ Q̃ which is
required for the application of Girsanov theorem in the proof of Theorem 3.5.
Theorem 3.6. Under Assumption 3.1, measures Q and Q̃ are equivalent if and
only if κ2 ≥ β.
Proof. By Theorem 3.5, the density Et [dQ̃/dQ] = Zt /Z0 . Therefore, the martingale
property for Zt under Q guarantees equivalence Q ∼ Q̃. Using Theorem 3.7(1)
concludes the proof.
Proof. (1) Theorem 9.2 in Lewis (2000) or Sin (1998) shows that Zt is Q-martingale
if and only if the process
dσt = ((κ0 + κ1 σt )(θ − σt ) + βσt2 )dt + ϑσt dW
(∗)
t
has unique strong solution under Q̃. Using Theorem 3.3, we immediately obtain
that the uniqueness holds iff κ2 ≥ β.
(2) According to Theorem 3.5, the dynamics of {σt , Rt } under Q̃ become
t
dσt = ((κ1 + κ2 σt )(θ − σt ) + βσt2 )dt + βσt dW
(0) t ,
+ εσt dW
(1)
(3.30)
t(0) .
dRt = (−σt )Rt dW
2450003-14
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
which holds if κ2 ≥ 2β. Rt is not a martingale if for some smooth, positive, increasing
function ϕ we have
lim inf (βξ 2 + (κ1 + κ2 ξ)(θ − ξ) + βξ 2 )ϕ(ξ)−1 > 0, (3.31)
ξ↑+∞
+∞
ϕ(ξ)−1 dξ < ∞, ε > 0. Choosing ϕ(ξ) = ξ 2 , (3.31) holds if κ < 2β.
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
where ε
Importantly we conclude that the log-normal SV model with linear drift leads
to arbitrages when the return–volatility correlation is positive.
2450003-15
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Corollary 3.3 (Dynamics under the inverse measure Q̃). The joint dynamics
of log-price Xt , the mean-adjusted volatility process Yt , and the QV It under the
inverse measure Q̃ are driven by
1 t(0) ,
dXt = (Yt + θ)2 dt + (Yt + θ)dW X0 = X,
2
t
dYt = (λ̃ − κ̃Yt − κ̃2 Yt2 )dt + β(Yt + θ)dW
(0) t ,
+ ε(Yt + θ)dW
(1)
Y0 = Y,
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
2450003-16
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(a) Steady state distribution of the volatility (b) Skewness of volatility as function of k2
Fig. 1. (a) The steady state PDF of the volatility computed using Eq. (3.38) with fixed κ1 = 4
and κ2 = {0, 4, 8}. (b) and (c) The skewness of the volatility computed using Eq. (3.39) and the
excess kurtosis of the unconditional returns distribution computed using Eq. (3.44), respectively,
both as functions of κ2 , for the three choices of κ1 = 1, 4, 8. Other model parameters are fixed to
θ = 1 and ϑ = 1.5.
2450003-17
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
where we exchange the integration order. It is clear that for odd n, m(n) = 0. For
the second and the fourth moments, we have
∞ ∞
2 2
m(2) = δt σ G(σ)dσ, m(4) = 3(δt) σ 4 G(σ)dσ. (3.43)
0 0
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
2450003-18
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
⎛ ⎞
0 0 0 0 0 0 ···
⎜ 0 −κ −κ2 0 0 0 · · ·⎟
⎜ ⎟
⎜c(2)θ2 2c(2)θ (c(2) − 2κ) −2κ2 0 0 · · ·⎟
Λ (0,∞)
=⎜
⎜ 0
⎟.
⎜ c(3)θ 2 2c(3)θ (c(3) − 3κ) −3κ 0 · · ·⎟
⎟
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
2
⎝ 0 0 c(4)θ 2 2c(4)θ (c(4) − 4κ) −4κ2 · · ·⎠
···
An approximate solution to ODEs (3.47) is obtained by truncating the number
∗
of terms to k ∗ and defining a finite dimensional vector of mth moments M (1,k ) ,
m = 1, . . . , k ∗ , as follows:
∗ ∗ ∗ ∗
∂τ M (1,k )
= Λ(1,k ) M (1,k )
+ C (1,k ) ,
∗ ∗
M (1,k ) (0) = (Y0 , Y02 , . . . , Y0k )T ,
∗ ∗
C (1,k )
= (0, c(2)θ2 , 0, 0, . . . , −k ∗ κ2 Y0k +1 T
) ,
⎛ ⎞
−κ −κ2 0 0 0 ···
⎜2c(2)θ (c(2) − 2κ) −2κ2 0 0 ··· ⎟
⎜ ⎟
⎜c(3)θ2 2c(3)θ (c(3) − 3κ) −3κ2 0 ··· ⎟
=⎜ ⎟.
∗
Λ(1,k )
⎜ 0 ⎟
⎜ c(4)θ 2 2c(4)θ (c(4) − 4κ) −4κ2 ··· ⎟
⎝ ··· ⎠
··· 0 0 c(k ∗ )θ 2 2c(k ∗ )θ (c(k ∗ ) − k ∗ κ)
(3.48)
The analytic solution to ODE system (3.48) is given by
∗ ∗ ∗ ∗
M (1,k ) (τ ) = expm{Λ(1,k ) t} · M (1,k ) (0) + (Λ(1,k ) )−1
∗ ∗ ∗
· (expm{Λ(1,k ) t} − I (k ) ) · C (1,k ) , (3.49)
−1
where expm() is the matrix exponent, · and are the matrix product and inverse,
∗
respectively, and I (k ) is k ∗ × k ∗ identity matrix.
2450003-19
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(n)
Thus, moment mτ solves recursive equation (omitting subscript t)
Using the fact that m(0) = 1 and c(1) = 0 we can present Eq. (3.50) for n =
0, 1, . . . as a matrix equation for infinite-dimensional column vector in Eq. (3.47).
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
Clearly, that under regularity conditions (real parts of eigenvalues of Λ(0,∞) are
negative), we must have limk→∞ ∂τ m(k) = 0. Thus, to make a finite-dimensional
approximation of Eq. (3.47), we fix k ∗ and assume
∗ ∗ ∗
∂τ m(k +1)
= 0 ⇒ m(k +1)
= Y0k +1
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
. (3.51)
∗
Finally, by substituting fixed values for m(0) and m(k +1) , we present (3.47) for
finite-dimensional vector of mth moments, m = 1, . . . , k ∗ , in Eq. (3.48). Solution in
Eq. (3.49) is obtained by integration.
In Fig. 2, we plot first four moments of the volatility process computed using
Eq. (3.49) with truncation order k ∗ = 4 (Fig. 2(a)) and k ∗ = 8 (Fig. 2(b)). For
comparison, we show estimates and 95% confidence intervals obtained by MC sim-
ulations. We see that the low-order truncation with k ∗ = 4 is consistent with MC
estimates for the first and second moments. The high-order truncation with k ∗ = 8
is consistent with the first four moments compared to MC estimates.
Remark 3.1. We note that if κ2 = 0, the system (3.48) becomes lower tridiagonal
so that it can be solved exactly for all moments up to k ∗ by the sequential integration
from n = 1 to n = k ∗ .
Proposition 3.4 (Bounded moments). Under Assumption 3.1 and κ1 θ > 0, the
QV is well defined under the MMA measure and E[(It )p ] < ∞ for any p, |p| ≥ 1.
If κ2 ≥ β, then the QV is well defined under the inverse measure and Ẽ[(It )p ] < ∞
for any p, |p| ≥ 1.
t
Proof. As E[(It )p ] = E[( 0 σs )p ] ≤ tp E[sups∈[0,t] σsp ] < ∞, the first statement fol-
lows from Proposition 3.1. The second statement follows using the volatility dynam-
ics under Q̃ in Eq. (3.28).
2450003-20
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Fig. 2. The first four moments of mean-adjusted volatility computed using Eq. (3.49) with the
truncation order k ∗ = 4 (a) and k ∗ = 8 (b) as functions of τ . The model parameters σ0 = 1.5,
θ = 1.0, κ1 = κ2 = 4, ϑ = 1.5. Dots and error bars denote the estimate and 95% confidence interval,
respectively, computed using MC simulations of model dynamics using scheme in Eq. (3.59) with
number of path equal 100 000 and daily time steps.
2450003-21
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Corollary 3.4 (Expected value). The expected value of the QV using k ∗ th order
truncation in Proposition 3.3 is given by
1 τ (2)
I!τ ≡ (m0 (τ ) + 2θm0 (τ ))dτ + θ2
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
(1)
τ 0
(3.53)
1 ! (2) ! (1) 2 ∗
= (m (τ ) + 2θm (τ )) + θ + O(k ),
τ
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
where m !(1)
and m
(2)
! are the first and second elements in vector M "(1,k ) (τ ) computed
∗
using Eq. (3.54), and O(k ∗ ) is the truncation error. The integrated moments of the
volatility with the k ∗ -order truncation are computed by
τ
"
M (1,k∗ )
(τ ) ≡
∗ ∗ ∗ ∗
M (1,k ) (τ ) = (Λ(1,k ) )−1 · (expm{Λ(1,k ) t} − I) · M (1,k ) (0)
0
∗ ∗ ∗ ∗
+ (Λ(1,k ) )−1 · ((Λ(1,k ) )−1 · (expm{Λ(1,k ) t} − I) − τ I) · C (1,k ) .
(3.54)
The approximation term O(k ∗ ) includes only the truncation error O(k ∗ ) because
Eq. (3.54) is computed analytically using matrix exponentiation. In Fig. 3, we
present comparison of the analytical solution for the expected QV using Eq. (3.53) to
the estimate and 95% confidence interval based on MC simulations. The truncation-
based analytic solution is consistent with MC estimates.
Fig. 3. The expected QV computed using Eq. (3.53) with k ∗ = 4. The top and bottom three
lines correspond to σ0 = 1.5 and σ0 = 0.5, respectively. The dot and the error bars denote the
estimate and 95% confidence interval computed using MC simulations of model dynamics with
scheme in SDE (3.59) with number of path equal 100 000 and daily time steps. The model cases of
mean-reversion correspond to κ1 = 4 and three choices κ2 = {0, 4, 8} with other model parameters
set to θ = 1.0, ϑ = 1.5.
2450003-22
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
2
T
We consider a time horizon T > 0 and an equidistant time grid tk = kΔ, Δ = n,
0 ≤ k ≤ n. Backward Euler–Maruyama (BEM) scheme for Lt is defined by
(∗) (∗)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Proof. As G(l) is continuous, G (l) > 0 and liml→±∞ G(l) = ±∞, the result
follows.
Then, there exists a constant Kp > 0 such that (E supt∈[0,T ] |L̂t − Lt |p )1/p ≤ Kp Δ.
Instead of Eq. (3.57), we prove slightly stronger estimate in Eq. (3.58). The
latter follows from the combination of Hölder inequality and Proposition 3.1. We
2450003-23
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Corollary 3.5. MC discretization scheme of the model dynamics (3.12) under the
MMA measure using log-price Xt in SDE (3.16) is given by
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
1 (0) (0)
Xtk+1 = Xtk − σt2k (tk+1 − tk ) + σtk (Wtk+1 − Wtk ),
2
(0) (0) (1) (1)
Ltk+1 = Ltk + ζ(Ltk+1 )(tk+1 − tk ) + β(Wtk+1 − Wtk ) + ε(Wtk+1 − Wtk ),
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
with Xt0 = ln S0 , Lt0 = ln σ0 , It0 = I0 , and where W (0) and W (1) are independent
Brownian motions.
We note that the first two dynamics are defined in (−∞, +∞), which does not
require extra handling of boundary conditions, in contrast to some affine models
for the variance process.
where the expectation E under the MMA measure Q is computed using dynam-
ics (3.34) and the expectation Ẽ under the inverse measure Q̃ is computed using
dynamics (3.36). Here we apply the fact that the dynamics are Markovian under
the both measures.
Theorem 4.1. The value function U (τ, X, I, Y ; p) solves the following PDE on the
domain [0, T ] × R × R+ × (−θ, ∞)
2450003-24
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
where diffusive operators L(Y ; p) , L(X; p) , and L(I; p) are defined as follows:
1 2 (p)
L(Y ; p) U = ϑ (Y + θ)2 UY Y + (λ(p) − κ(p) Y − κ2 Y 2 )UY ,
2
2 1
L(X; p)
U = (Y + θ) (UXX − pUX ) + βUXY ,
2
(p) (p)
κ(p) = κ1 − κ2 θ + 2κ2 θ, λ(p) = (κ2 − κ2 )θ2 ,
#
(p) κ2 , p = 1,
κ2 =
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
κ2 − β, p = −1.
Proof. We note that the classic Feynman–Kac formula (for an example, see Sec.
A.17 in Musiela & Rutkowski (2009)) cannot be applied directly to Eq. (4.1) because
the model dynamics (3.34) and (3.36) do not satisfy the linear growth condition. In
Sec. A.2 we prove that the existence and the uniqueness is guaranteed by Theorem
A.1. This theorem states that if the covariance matrix of diffusive operator L ≡
L(Y ; p) + L(X; p) + L(I; p) is nonnegative definite and the state variables neither
explode nor leave an open domain, then there exists a unique solution to the PDE
representation of the conditional expectation under corresponding model dynamics.
Now using the diffusive operators in the valuation PDE (4.2), we obtain that
the diffusion matrix a in Eq. (A.14) equals
⎛ ⎞
(Yt + θ)2 β(Yt + θ)2 0
⎜ ⎟
a=⎜
⎝β(Yt + θ)
2
ϑ2 (Yt + θ)2 0⎟
⎠ (4.4)
0 0 0
Since β < ϑ, we observe that 0 = λ1 < λ2 < λ3 , hence matrix a is only nonnegative
definite and its smallest eigenvalue is always zero.
We now set the domain D of state variables {Xt , Yt , It } in PDE (4.2) by D =
R×(−θ, +∞)×R+ . As we showed in Theorem (3.2), under the MMA measure (3.34)
and the inverse measure (3.36), the boundary points {0, +∞} are unattainable for
the volatility process σt , hence solution for mean-adjusted volatility process Yt and
QV It cannot explode or leave D before T . Same applies to Xt as {0, +∞} are
unattainable for σt . Thus, condition 1 of Theorem A.1 is satisfied. As boundedness
condition A.1 is satisfied by construction, the statement of Theorem 4.1 follows
from Theorem A.1 stated in Sec. A.2.
2450003-25
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
G(τ, X, I, Y ; Φ, Ψ, Θ; p)
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
#
E[e−ΦXτ −ΨIτ −ΘYτ |X0 = X, I0 = I, Y0 = Y ], p = 1,
= (4.5)
Ẽ[e−ΦXτ −ΨIτ −ΘYτ |X0 = X, I0 = I, Y0 = Y ], p = −1.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Using model dynamics (3.34) under the MMA measure and model dynamics
(3.36) under the inverse measure, respectively, the MGF solves the following PDE
on the domain [0, T ] × R × R+ × (−θ, ∞):
with operators defined in Eq. (4.3). First, we establish a sufficient condition for the
existence of the MGF for the three state variables.
Theorem 4.2. Given the transform variable Φ = ΦR + iΦI ∈ C, the MGF of the
log-price Xτ
exists for ΦR ∈ (−1, 0), if Zτ is a martingale under the MMA measure. By Theorem
3.7(1), the necessary condition is κ2 ≥ β.
Similarly, the MGF of the log-price Xτ
G(τ, I; Ψ; p = 1) = E[e−ΨIτ ]
√
exists for ΨR < 0 if κ2 > ϑ −2ΨR .
2450003-26
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proof. We denote Kt = eIt which satisfies SDE dKt = Kt dIt = Kt σt2 dt and
set p = −ΨR . Feynman–Kac formula allows to recast the solution U (τ, σ, K) =
E(KTp |Kt = K, σt = σ) as the solution of the Cauchy problem
∂U
− + L(σ,K) U = 0, U (0, σ, K) = K, (4.7)
∂τ
where operator L(σ,K) is defined by
∂ 1 ∂2 ∂
L(σ,K) = (κ1 + κ2 σ)(θ − σ) + ϑ2 σ 2 2 + Kσ 2 .
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
∂σ 2 ∂σ ∂K
We build a supersolution Ū (τ, σ, K) to (4.7) using the following ansatz:
Ū = K p u(τ, σ), u(τ, σ) = eaσ+bτ , (4.8)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
which yields
1 2 2
−uK p
−b + κ1 θa + (κ2 θ − κ1 )aσ + ϑ a + p − κ2 a σ .
2
2
We show there exists a ≥ 0 such that
1 2 2
lim −b + κ1 θa + (κ2 θ − κ1 )aσ + ϑ a − κ2 a + p σ 2 < +∞. (4.9)
σ↑+∞ 2
We notice that inequality (4.9) holds if there exist a > 0 such that 12 ϑ2 a2 −κ2 a+p <
0. As latter defines a parabola in a, we can always find required positive a, if
√ √
κ2 > ϑ 2p = ϑ −2ΨR .
Theorem 4.4. Given the transform variable Θ = ΘR + iΘI ∈ C, the MGF of the
mean-adjusted volatility process Yτ
#
E[e−ΘYτ |Y0 = Y ], p = 1,
G(τ, Y ; Θ; p) =
Ẽ[e−ΘYτ |Y0 = Y ], p = −1
ϑ2 ϑ2
exists for ΘR < 0, if κ2 > 2 |ΘR |, p = 1 and κ2 − β > 2 |ΘR |, p = −1.
2450003-27
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
where coefficients are defined in (4.3). We show there exists a ≥ c such that
(p) 1
lim b − (λ(p) − κ(p) Y − κ2 Y 2 )a − ϑ2 (Y + θ)2 a2 < ∞.
Y ↑+∞ 2
(p)
We notice that last inequality is satisfied if there exist a ≥ c > 0 such that −κ2 a +
1 2 2 (p) ϑ2
2 ϑ a < 0. Dividing by a > 0 we conclude that it is satisfied if κ2 > 2 c. Rewriting
the last condition in terms of κ2 , β using (4.3), completes the proof.
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
of coefficients. Instead, we can show that the solution to the PDE (4.6) can be rep-
resented by an exponential-affine ansatz E [∞] using an infinite dimensional vector
A(τ ) = {A(k) (τ ; Φ, Ψ, Θ; p)}, k = 0, 1, . . . , ∞, as
# ∞
%
$
E (τ, X, I, Y ; Φ, Ψ, Θ; p) = exp −ΦX − ΨI +
[∞] (k)
A (τ ; Φ, Ψ, Θ; p)Y k
,
k=0
(4.13)
A(k)
τ =A M
(k)
A + (L(k) ) A + H (k) , k = 0, 1, . . . , ∞, (4.14)
where M (k) , L(k) , H (k) are infinite-dimensional sparse matrices and vectors.
We note the similarity between the current ansatz (4.13) for problem (4.6) and
the problem of computing the moments of the volatility process addressed in Propo-
sition 3.3. There we first represent the recursive solution to the moments as an
infinite dimensional vector in Eq. (3.47). We then obtain an approximate solution
by a truncation of infinite series and specifying a boundary condition to reduce the
problem to a finite-dimensional vector (3.48), which can be solved using analytical
methods. Here we follow this insight for solving PDE (4.6) and term our (truncated)
solution as the affine expansion for a given order.
Theorem 4.5 (First-order affine expansion). The MGF in Eq. (4.5) can be
decomposed as follows:
where E [1] is the leading first-order term and R[1] is the remainder term, such that
E [1] (τ = 0) = 1 and R[1] (τ = 0) = 0. The leading term E [1] is given by the
2450003-28
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
exponential-affine form
# %
$
2
[1] (k) k
E (τ, Y ; Φ, Ψ, Θ; p) = exp A (τ ; Φ, Ψ, Θ; p)Y , (4.16)
k=0
⎪
⎪ 0 0 0 ⎛ ⎞ 0 0 0 ⎪
⎪
⎪
⎪ ⎟⎪
⎨⎜ ⎟ 0 ⎜ ⎪
0 0
⎜ 2 2
θ ϑ ⎟ ⎜ ⎟ ⎜ ϑ 2 ⎟ ⎬
M (k)
= ⎜ ⎜ 0 0⎟ ⎜
, ⎝0 θϑ 2 2 2 ⎟
θ ϑ ⎠, ⎜ ⎜ 0 2θϑ 2⎟
,
⎪ ⎟ ⎟
⎪
⎪ ⎝ 2 ⎠ ⎝ 2 ⎠⎪⎪
⎪
⎪
⎩ 0 θ ϑ2 2
0 ⎪
⎭
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
0 0 0 0 2θϑ2 2θ2 ϑ2
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎪
⎪ 0 0 0 ⎪
⎪
⎨⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎬
⎜
L (p) = ⎝λ − θ βΦ⎠, ⎝
(k) ⎟ ⎜ −κ − 2θβΦ ⎟ ⎜ (p) ⎟
⎠, ⎝ −βΦ − κ2
(p) 2 (p)
⎪ ⎠ ⎪,
⎪
⎩ ⎪
⎭
2 2
θ ϑ 2(λ + θϑ − θ βΦ)
(p) 2 2
ϑ − 2κ(p) − 4θβΦ
) *
1 2 2 1
H (k) (p) = θ (Φ + pΦ − 2Ψ), θ(Φ2 + pΦ − 2Ψ), (Φ2 + pΦ − 2Ψ)
2 2
(4.17)
with the initial condition A(0) = (0, −Θ, 0) .
The remainder term R[1] solves the following PDE (omitting arguments) :
−Rτ[1] + L(Y ; p) R[1] = −E [1] F [1] (Y, A(1) , A(2) ),
(4.18)
R[1] (0, Y ; Φ, Ψ; p) = 0
with operator L(Y ) defined in Eq. (4.3) and residual F [1] defined by
$
4
[1] (1) (2)
F (Y, A ,A )= C (n) (τ ; A(1) , A(2) )Y n ,
n=3
(4.19)
C (3) (τ ) = 2A(2) (ϑ2 (2θA(2) + A(1) ) − βΦ − κ2 ),
2450003-29
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(a) R[A(τ ),Φ = −0.50 + 2.00i (b) J[A(τ ),Φ = −0.50 + 2.00i
Fig. 4. The solution of ODEs of the first order expansion in Eq. (4.17) as function of τ for
Φ = −0.5 + 2i, p = 1. We use the model parameters calibrated to Bitcoin options data and
reported in Eq. (6.4). (a) The real part of the solution, (b) the imaginary part of the solution, (c)
the the real and imaginary parts of the exponential term E [1] in Eq. (4.16).
Proof. The proof is identical to that of Theorem 4.7 for the second-order expansion.
We skip it for brevity.
In Fig. 4, we show the real and imaginary parts of ODE solutions A(τ ) and E [1] .
We see that functions A(1) and A(2) quickly reach an equilibrium point, while A(0)
is the nonstationary part which contributes to the leading term E [1] .
2450003-30
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
$
4
|R[1] (τ, Y ; Φ, Ψ, Θ; p)| ≤ C (4) (τ ∗ ) × Mσ(n) (τ ∗ ), (4.20)
n=3
(n)
where Mσ (τ ) is the nth central moment of the volatility defined by
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
Corollary 4.2. First-order affine expansion for the MGF (4.5) is obtained by using
the leading term E [1] in Eq. (4.15):
where the approximation error arises from the truncation of the infinite dimensional
affine anzats in Eq. (4.13) with the error magnitude estimated using Eq. (4.20).
In Fig. 6, we illustrate that the first-order approximation (4.22) for the MGF
produces valid and accurate PDFs for all three state variables.
where E [2] is the leading term such that E [2] (τ = 0) = 1 and R[2] (τ = 0) = 0 is the
remainder term.
The leading term E [2] (τ, Y ; Φ, Ψ; p) is given by the exponential-affine form
# 4 %
$
E [2] (τ, Y ; Φ, Ψ, Θ; p) = exp A(k) (τ ; Φ, Ψ, Θ; p)Y k , (4.24)
k=0
A(k)
τ = A M (k) A + (L(k) (p)) A + H (k) (p),
2450003-31
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
⎧⎛ ⎞
⎪
⎪ 0 0 0 0 0 ⎛ ⎞
⎪
⎪ ⎜ ⎟ 0 0 0 0 0
⎪
⎪ ⎜ 2
θ ϑ 2
⎟ ⎜
⎪
⎪ ⎜0 0 0 0⎟ ⎜0 θϑ2 θ2 ϑ2 0 0⎟
⎪
⎨⎜ ⎟ ⎜ ⎟
⎜
2
⎟ ⎜ ⎟
M (k) = ⎜ ⎟, ⎜0 θ2 ϑ2 0 0 0⎟ ⎟,
⎪⎜0
⎪ 0 0 0 0⎟ ⎜ ⎟
⎪
⎪ ⎜ ⎟ ⎝0 0 0⎠
⎪
⎪ ⎜0 ⎟ 0 0
⎪
⎪ ⎝ 0 0 0 0⎠
⎪
⎩ 0 0 0 0 0
0 0 0 0 0
⎛ ⎞
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
0 0 0 0 0
⎜ ⎛ ⎞
⎜ ϑ2
3 2 2 ⎟ ⎟ 0 0 0 0 0
⎜0 2θϑ2 θ ϑ 0⎟ ⎜
⎜
⎜
2 2 ⎟ ⎜0 0
⎟ ⎜ ϑ2 3θϑ2 2θ2 ϑ2 ⎟ ⎟
⎜ ⎟ ⎜ ⎟
⎟,
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
⎜0 2θϑ2 2θ2 ϑ2 0 0⎟ , ⎜ 0 ϑ 2
4θϑ 2
3θ 2 2
ϑ 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ 3 2 2 ⎟ ⎝0 3θϑ2 3θ2 ϑ2 ⎠
⎜0 θ ϑ 0 0 0⎟
0 0
⎜ 2 ⎟
⎝ ⎠ 2
0 2θ ϑ 2
0 0 0
0 0 0 0 0
⎛ ⎞⎫
0 0 0 0 0 ⎪
⎪
⎜ ⎟⎪⎪
⎪
⎜0 3 2 2 ⎟⎪
4θϑ ⎟⎪
⎜ 0 0 ϑ ⎪
⎜ 2 ⎟⎪⎪
⎜ ⎟⎪⎬
⎜ 2 2⎟
⎜0 0 2ϑ 2
6θϑ 2
4θ ϑ ⎟ ,
⎜ ⎟⎪
⎜ ⎟⎪⎪
⎟⎪
3 2 9 2 2
⎜0 ϑ 6θϑ 2
θ ϑ 0 ⎪
⎪
⎜ ⎟ ⎪
⎝ 2 2 ⎠⎪⎪
⎪
⎪
⎭
0 4θϑ2 4θ2 ϑ2 0 0
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎪
⎪ 0 0 0
⎪
⎪
⎪
⎪⎜λ(p) − θ2 βΦ⎟ ⎜ −κ(p) − 2θβΦ ⎟ ⎜ ⎟
⎟ ⎜ ⎟
(p)
⎨⎜
⎪ ⎟ ⎜ ⎜ −βΦ − κ2 ⎟
⎜ ⎟ ⎜ ⎟
L (p) = ⎜
(k) 2
θ ϑ 2 ⎟, ⎜2(λ + θϑ − θ βΦ)⎟, ⎜ ϑ2 − 2κ(p) − 4θβΦ ⎟
(p) 2 2 ⎜
⎟,
⎪⎜ ⎟ ⎜ ⎟
⎪
⎪⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎪
⎪⎝ 0 ⎠ ⎝ 2
3θ ϑ 2 ⎠ ⎝3(2qϑ2 + λ(p) − θ2 βΦ)⎟
⎜
⎠
⎪
⎪
⎩ 0 0 2 2
6θ ϑ
⎛ ⎞ ⎛ ⎞⎫
0 0 ⎪
⎪
⎜ ⎟ ⎜ ⎟⎪⎪
⎪
⎜ ⎟ ⎜ ⎟ ⎪
⎟⎪
0 0
⎜ ⎟ ⎜ ⎬
⎜ (p) ⎟ ⎜ ⎟
⎜ −2(βΦ + κ2 ) ⎟, ⎜ 0 ⎟ ,
⎜ ⎟ ⎜ ⎟⎪
⎜ 3(ϑ2 − κ(p) − 2θβΦ) ⎟ ⎜ −3(βΦ + κ2 )
(p) ⎟⎪⎪
⎪
⎝ ⎠ ⎝ ⎠⎪⎪
⎪
⎭
4(3θϑ + λ − θ βΦ)
2 (p) 2
2(3ϑ − 2κ − 4θβΦ)
2 (p)
) *
1 2 2 1
H (k) (p) = θ (Φ + pΦ − 2Ψ), θ(Φ2 + pΦ − 2Ψ), (Φ2 + pΦ − 2Ψ), 0, 0
2 2
(4.25)
2450003-32
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
The remainder term R[2] solves the following PDE (omitting arguments)
with operator L(Y ) defined in Eq. (4.3) and function F [2] defined by
$
8
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
F [2] (τ, Y ; A(1) , A(2) , A(3) , A(4) ) = C (k) (τ ; A(1) , A(2) , A(3) , A(4) )Y k ,
k=5
C (5) 2
(τ ) = 3ϑ A (3) (3)
(3θA + 2A (2)
) + 4A(4) (ϑ2 (3θ2 A(3) + 4θA(2) + A(1) ) − κ2 ),
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
1 2
C (6) (τ ) = ϑ (16θ2 (A(4) )2 + 16A(4) (3θA(3) + A(2) ) + 9(A(3) )2 ),
2
C (7) (τ ) = 4ϑ2 A(4) (4θA(4) + 3A(3) ), C (8) (τ ) = 8ϑ2 (A(4) )2 . (4.27)
Proof. The proof follows by analogy to the first-order expansion in Theorem 4.5.
Remark 4.1. We emphasize that sparse matrices M (k) in the quadratic term do
not depend on the measure parameter p, while the linear L(k) (p) and free terms
H (k) (p) depend on p in ODEs (4.17), (4.25) for the first- and the second-order
expansion.
We now present the result on the existence of continuous solution to Eq. (4.25)
in Proposition 4.2 and discuss it further in Sec. 4.2.3.
In Fig. 5, we show the real and imaginary parts of A(τ ) for the second-order
expansion and the leading term E [2] . We see that, similarly to the first order expan-
sion illustrated in Fig. 4, functions A(1) , A(2) , A(3) , A(4) reach an equilibrium point,
while A(0) is the nonstationary part which contributes to the leading term E [2] .
By comparing the low order terms A(0) , A(1) , A(2) between the first and the sec-
ond expansions, we notice no visible difference, which suggest that the expansion is
rather recursive and higher order terms make insignificant contribution to the low
order terms. We also notice that terms A(3) and A(4) are very small in magnitude
compared to the first three terms, which suggest that higher order term only provide
a small marginal contribution.
2450003-33
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(a) R[A(τ ),Φ = −0.50 + 2.00i (b) J[A(τ ),Φ = −0.50 + 2.00i
Fig. 5. The solution of ODEs in Eq. (4.25) as function of τ for Φ = −0.5 + 2i, p = 1. We use the
model parameters calibrated to Bitcoin options data and reported in Eq. (6.4). (a) The real part
of the solution, (b) the imaginary part of the solution, (c) the real and imaginary parts of the
exponential term E [2] in Eq. (4.24).
$
8
|R (τ, X; Φ, Ψ, Θ; p)| ≤
[2]
Cn (τ ∗ ) × Mσ(n) (τ ∗ ), (4.28)
n=5
(n)
where Mσ is the nth central moment of the volatility defined by (4.21).
2450003-34
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proof. The formal solution for the remainder term R[2] solving problem (4.26) is
obtained by applying the Feynman–Kac formula and is given by
R[2] (τ, Y ; Φ, Ψ)
τ ∞
= [E [2] (τ − t, Y ; Φ, Ψ)F [2] (τ − t, Y )]P (t, Y ; Y )dtdY , (4.29)
0 −θ
τ ∞
≤ |F [2] (τ − t, Y )P (t, Y ; Y )|dtdY . (4.30)
0 −θ
Here, we apply the bound on the MGF |E [2] (τ − t, Y ; Φ, Ψ)| ≤ 1 and integrate
out the log-return X and the QV I. Next we substitute the polynomial function
F [2] in Eq. (4.27). We use the continuity of functions A(k) , k = 1, 2, 3, 4, and apply
the first mean value theorem for definite integrals to estimate the time integral.
Finally, we approximate the expected moments of the mean-adjusted volatility by
(n)
its steady-state nth order central moments Mσ specified in Eq. (4.21).
Corollary 4.4. Second-order affine expansion for the MGF (4.5) is obtained by
using the leading term E [2] in Eq. (4.24) as
where the approximation error arises from the truncation of the infinite dimensional
affine anzats in (4.13) with the error magnitude estimated by Eq. (4.28).
In Fig. 6, we illustrate that the second-order approximation (4.31) for the MGF
produces valid and accurate PDFs for all three state variables.
2450003-35
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(c) Volatilityστ
Fig. 6. PDFs computed using the inversion of the first order expansion term E [1] in Eq. (4.16)
and the second-order expansion term E [2] in Eq. (4.24) for the state variables under the MMA
measure: (a)–(c) PDFs of log-return, Xτ , of QV normalized by τ , Iτ /τ , of volatility στ , στ = Yτ +θ,
respectively. We use the model parameters for Bitcoin options reported in Eq. (6.4) with time to
maturity of one month, τ = 1.0/12.0. The blue histogram is computed using realizations from MC
simulations of joint model dynamics (3.12) using scheme in Eq. (3.59) with the number of path
equal 400,000 and daily time steps.
(1) τ+ (Φ; A) ≥ τ+ ( Φ; A), Φ ∈ C, i.e. real part of A(τ ; Φ) cannot blow up before
A(τ ; Φ).
+2m
k=0 Ak (τ ; Φ; p)Y → +∞ as τ ↑ τ+ (Φ), Φ ∈ R, i.e. leading term of the affine
k
(2)
expansion does not vanish as we approach blow-up time when transform variable
Φ is real.
2450003-36
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(3) limτ ↑τ+(Φ) R[n] (τ ; Φ) > −∞, Φ ∈ R, i.e. remainder term is uniformly bounded
from below.
If MGF in Eq. (4.6) is finite with G(τ0 ; X, I, Y ; Ψ = Φ, Ψ = 0, Θ = 0) < ∞,
then continuous solution A(τ ; Φ, Ψ = 0, Θ = 0) of Eq. (4.17) and of Eq. (4.25) exists
on [0, τ0 ).
By Theorem 4.7, we obtain that continuous solution A(τ ) for the first- and
second-order affine expansions in Eqs. (4.16) and (4.24), respectively, exists on
[0, +∞) if 1, 2, 3 hold.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Proposition 4.4 (Volatility moments). The MGF in Eq. (4.31) with the second-
order affine term E [2] in (4.24) reproduces exactly the expected value and the vari-
ance of the volatility for κ2 = 0.
Proposition 4.5 (Log-price moments). The MGF in Eq. (4.31) with the second-
order affine term reproduces exactly the expected value and the variance of the log-
price for κ2 = 0.
2450003-37
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proposition 4.6 (QV moments). The MGF in Eq. (4.31) with the second-order
affine term reproduces exactly the expected value and the variance of QV for κ2 = 0.
Proof. Propositions 4.4 and 4.5 are proved in Secs. A.3.1 and A.3.2, respectively.
Proof of Proposition 4.6 is similar to the proof of Proposition 4.5.
Remark 4.2. In Remark 3.1 we noted that the system of volatility moments (3.48)
is solved exactly for κ2 = 0, otherwise the solution is approximate. Thus, we cannot
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
generalize above statements for κ2 > 0. However, it is our hypothesis that moments
obtained using (4.24) correspond to moments computed in (3.48) using truncation
order k ∗ = 4 and that truncation error as function of κ2 is very small.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
5. Option Valuation
We apply the zero-drift log-price process Xt in Eq. (3.16) and Eqs. (3.17) and (3.18)
for spot and futures prices, respectively, to consider option valuation on both spot
and futures underlyings. We denote by PT the price of either spot or futures asset
PT = eμ̄(T ) eXT , X0 = ln S0 , (5.1)
where μ̄(T ) = r̄(0, T ) for spot price and μ̄(T ) = r̄(0, T ) − q̄(0, T ) for futures.
As we conclude in Corollaries 4.2 and 4.4, we apply affine expansion given by
either the first order in Eq. (4.16) or the second order in Eq. (4.24) as analytic
solutions to the MGF under the dynamics (3.16). Given that the MFG is available
analytically, we then apply Lewis–Lipton approach for valuation of vanilla options
using Fourier transform, see Lewis (2000), Lipton (2001, 2002).
2450003-38
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
× E [m]
(τ, Y ; Φ = iy − 1/2, Ψ = 0, Θ = 0; p = 1)dy , (5.4)
y 2 + 1/4
where X ∗ = ln(S0 /K)+μ(T ) is log-moneyness, E [1] and E [2] are given in Eqs. (4.16)
and (4.24), respectively.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Proof. Using affine expansion for MGF G with either first-order in Eq. (4.22) or
second-order in Eq. (4.31), the PDF of Xτ is computed by Fourier inversion
PDF(τ, X, Y ; X ; p = 1)
∞
1
= exp{Φ(X − X)}E [m](τ, Y ; Φ, Ψ = 0; p = 1)dΦ . (5.5)
π 0
Assuming that inner integrals are finite, we exchange the integration order as
∞
1 −r(T )
U (τ, X) = e !(Φ)E (τ, Y ; Φ, Ψ = 0; p = 1)dΦ ,
u [m]
(5.6)
π 0
2450003-39
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
As a result, calls and puts on spot underlying are valued using Eqs. (5.2)
Proposition 5.2. The valuation formula (5.6) for the inverse capped payoff
becomes
∞
, (τ, X) = 1
U
∗
e−(iy+1/2)X 2
1
E [m]
π 0 y + 1/4
× (τ, Y ; Φ = iy + 1/2, Ψ = 0, Θ = 0, p = −1)dy, (5.13)
where X ∗ = ln(S0 /K)+μ(T ) is log-moneyness, E [1] and E [2] are given in Eqs. (4.16)
and (4.24), respectively.
2450003-40
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
∗ 1
= −e−ΦX
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
, (5.14)
Φ(Φ − 1)
[Φ] < 1. Integral in Eq. (5.14) is finite for 0 < [Φ] < 1. We set Φ = iy + 1/2,
then
∗ 1
!(Φ = iy + 1/2) = e−(iy+1/2)x
u . (5.15)
y2 + 1/4
As a result, calls and puts are valued using capped payoff (5.12):
with the last equality following from Eq. (4.33). Given the term structure of futures
prices f0 (T ) in Eq. (3.18), the put option on the future is
K , (τ, X)
p(τ, F, K) = −U (5.18)
f0 (T )
with log-moneyness X ∗ = ln(f0 (T )/K). We emphasize that Eq. (5.16) holds only if
Q ∼ Q̃, which requires restriction κ2 ≥ β, see Theorem 3.6. In addition, we assume
that the inverse process Rt = exp(−Xt ), t ∈ (0, T ], is a true Q̃-martingale. For this
we need κ2 ≥ 2β by Theorem 3.7.
2450003-41
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proposition 5.3. The value of the call option on the QV under Q is given by
U (τ, X, Y, I) = e−r̄(T ) E[u(IT )|Ft ]
e−r̄(T ) 1 +∞
= [û(Ψ)E [m] (τ, Y ; Φ = 0, Ψ; Θ = 0; p = 1)]dΨ,
T π 0
(5.20)
where
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
e−ΨI ΨT K
û(Ψ) = e (5.21)
Ψ2
provided [Ψ] < 0, with E [1] and E [2] given in Eqs. (4.16) and (4.24), respectively.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
For a call option on QV with strike K, provided [Ψ] < 0, Eq. (5.22) becomes
+∞
û(Ψ) = e−ΨI u(I )eΨI dI
0
+∞
e−ΨI ΨT K
= e−ΨI max{I − T K, 0}eΨI dI = e .
0 Ψ2
Sepp (2008) derives transforms û(Ψ) for typical payoff on the QV including puts
and calls on the square root of the QV. Assuming that the inverse measure Q̃ is a
martingale measure for generic spot and futures underlying Pt defined in Eq. (5.1),
the inverse option on the QV is given under Q̃ by
u(IT )
Ũ (τ, X, Y, I) = Ẽ Ft . (5.23)
PT
Similarly to Eq. (2.5), values Ũ(τ, X, Y, I) and U (τ, X, Y, I) in Eqs. (5.23) and (5.20),
respectively, are related by
U (τ, X, Y, I) = Pt Ũ (τ, X, Y, I).
Proposition 5.4. The value of the call option on the QV under the inverse measure
Q̃ is given by
e−r̄(T ) 1 +∞
Ũ (τ, X, Y, I) = [û(Ψ)eX−μ(T ) E [m]
T π 0
× (τ, Y ; Φ = 1, Ψ, Θ = 0; p = 1)]dΨ, (5.24)
[m]
provided [Ψ] < 0, with E and û(Ψ) defined as in Eq. (5.21).
2450003-42
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Proof. Using affine expansion for MGF G with either first-order in Eq. (4.22) or
second-order in Eq. (4.31), we compute the joint density of {Xτ , Iτ } using inverse
Fourier transform by
PDFX,I (τ, X, Y, I; X , I )
+∞ +∞
1 Φ(X −X)+Ψ(I −I) [m]
= 2 e E (τ, Y ; Φ, Ψ)dΦ dΨ .
4π −∞ −∞
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
As a result, we obtain
+∞ +∞
e−r̄(T )
Ũ (τ, X, Y, I) = e−X −μ̄(T ) u(I )
T −∞ 0
+∞ +∞
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
provided [Ψ] < 0. We integrated out the complex exponential in Eq. (5.26) using
results of Sepp (2007). Thus, Eq. (5.25) becomes
+∞
e−r̄(T ) 1
Ũ (τ, X, Y, I) = [e−X−μ̄(T ) E [m] (τ, Y ; Φ = 1, Ψ)û(Ψ)]dΨ.
T 2π −∞
(5.27)
2450003-43
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
and J strikes. For option valuation using the MGF, we apply the affine expansion
given by either the first order in Eq. (4.16) or the second order in Eq. (4.24).
First, we fix the space grid with size P (typically, P ≈ 500) of transform variable
Φ for the log-price. We solve the system of ODEs numerically for either the first-
order expansion in Eq. (4.16) or for the second-order expansion in Eq. (4.24) in
time up to the last maturity time. The computational cost of this step is
C (1) = O(P × Nmax ), (6.1)
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
where Nmax is the number of time steps for ODE solver (typically we have one-two
steps per one week) and O is the number of arithmetic (elementary) operations.
Second, we compute values of the MGF on the grid of Φ for each maturity slice.
We then compute values of capped payoffs for the MMA measure in Eq. (5.4) or
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
for the inverse measure in Eq. (5.13) using Simpson rule for numerical integration.
We then value all options in a given chain. The computational cost of this step is
C (2) = O(J × M × P ). (6.2)
The main difference with numerical implementation of affine models is that the
first step can be computed with cost close to O(P × M ) because the solution to the
MGF can be computed analytically without solving ODEs numerically. The cost
ratio between the implementation of the log-normal SV model and an affine SV
model compares favorably when the number of maturities becomes large with M
close to Nmax , and when the number of strikes J becomes large as well.
2450003-44
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
with weight wn (T, K) = Vn (T, K) set to option Black–Scholes vega Vn (T, K).
In Fig. 7(a), we show the time series with the quality of the model fit which is
computed using the square root of average squared differences between model and
market implied volatilities. As a benchmark, we show the average spread between
(a) Average mean-squared error and bid-ask (b) Initial and mean volatilities
volatility spread
Fig. 7. Time series of the quality of the model calibration and of calibrated model parameters
using weekly model calibrations to prices of option on Bitcoin traded on Deribit exchange from
April 2019 to October 2023. (a) Time series of the mean squared error (MSE) between model and
market implied volatilities and the average bid-ask (Bid/Ask) spread of quoted implied volatilities.
(b)–(d) Time series of estimated initial volatility σ0 and mean volatility θ, volatility beta β, and
volatility-of-volatility ε, respectively.
2450003-45
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
implied volatilities of ask and bid quotes for options in the calibration set. We
observe that most of the times, the average model mis-pricing is within the average
bid-ask spread. We see that year 2020 of COVID-19 pandemic is characterized by
high volatility and high spreads, yet the model is able to fit these markets well. In
recent years of 2022 and 2023, bid-ask spreads declined while the model error (the
average difference between market and model implied volatility) became less than
1% most of the times.
In Figs. 7(b)–7(d), we show the time series of calibrated values of initial volatility
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
than relative value of calls and puts is intact despite falling level of the overall
volatility. The volatility beta fluctuates between positive and negative values most
of the times following the market sentiment. When the sentiment is bullish, calls are
bid and volatility beta becomes positive and, vice versa, during bearish sentiment,
puts are bid and the volatility beta becomes negative. Thus, as we discuss in Sec. 3.3,
the conventional SV models may not be arbitrage-free for the valuation of options
on cryptocurrencies when return–volatility correlation become positive.
In Fig. 8, we show the quality of model fit to Bitcoin options traded on Deribit
exchange on 20 June 2023. On this day, fitted model parameters are the following:
σ̂0 = 0.41, θ̂ = 0.38, β̂ = 0.50, ε̂ = 3.06, κ̂1 = 2.21, κ̂2 = 2.18. (6.4)
In Fig. 8, we show the quality of model fit on this date. We observe positive
skewness of implied volatilities, which results in positive volatility beta. Continuous
black line shows model implied volatilities computed using the second-order affine
expansion in Eq. (5.9). Bid and ask displays the market bid and ask quotes, the
ATM is the ATM mid-point volatility. MSE denotes the mean squared error between
the model implied volatilities and the mid of market bid-ask implied volatilities. We
see that the log-normal SV model calibrated to Bitcoin options is able to capture
the market implied skew very well across most liquid maturities with only four
parameters. The average MSE is about 0.5% for longer maturities, which is within
the bid-ask spread. Calibration to the ATM region can be further improved using
a term structure of the mean volatility θ. In a practical setting, the SV model
dynamics can be augmented with a local volatility (Lipton (2002)) to accurately fit
the implied volatility surface.
2450003-46
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(a) (b)
(c)
Fig. 8. Quality of model fit to Bitcoin options on 20 June 2023 reported in Eq. (6.4) using the three
most liquid slices: one week (a), two weeks (b), one month (c). Continuous black line is model
implied volatilities computed using the second-order affine expansion in Eqs. (5.9) and (5.16). Bid
and ask displays the market bid and ask quotes, the ATM is the ATM mid-point volatility. Model
implied volatility is displayed using the continuous black line. MSE is the mean squared error
between the model and the mid of market bid-ask volatilities.
2450003-47
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
Fig. 9. Implied volatilities of Bitcoin options which correspond to the slices used in model calibra-
tion and a uniform range of strikes and which are computed using model parameters reported in
Eq. (6.4) under the MMA measure (MMA) and the inverse measure (Inverse). Continuous aqua
and pink lines show model implied volatilities computed using the second-order affine expansion
in Eqs. (5.9) and (5.16), respectively. Dashed lines labeled by MC − 0.95ci and MC + 0.95ci are
MC 95% confidence intervals computed using scheme in Eq. (3.59) with number of path equal
400, 000 and daily time steps. MSE is the mean squared error between BSM volatilities inferred
from model values and the MC estimates.
2450003-48
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
(c) slice - 1m
Fig. 10. Implied volatilities of call options on QV of Bitcoin using model parameters in Eq. (6.4)
and for the same maturities as in model calibration in Fig. 8. We use the valuation under the MMA
measure (MMA) and the inverse measure (Inverse). BSM volatilities are implied from model prices
using expected QV Ibτ = {0.18, 0.20, 0.23} computed using Eq. (3.53). Continuous aqua and pink
lines show model implied volatilities computed using the second-order affine expansion under MMA
measure in Eq. (5.20) and under inverse measure in (5.24), respectively. Dashed lines labeled by
MC−0.95ci and MC+0.95ci are MC 95% confidence intervals computed using scheme in Eq. (3.59)
with number of path equal 400, 000 and daily time steps. MSE is the mean squared error between
BSM volatilities inferred from model values and the MC estimates.
2450003-49
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
in Eq. (6.4) for computing model implied volatilities for the same maturities as in
Fig. 8. As benchmark, we use the MC simulation of the model dynamics (3.12)
using scheme in Eq. (3.59) with the number of paths equal 400, 000 and daily time
steps. We show the 95% confidence interval of MC estimates as dashed lines like
bid/ask lines
We observe that option values computed using the second order affine expansion
under the MMA and inverse measures are very close to each other with the differ-
ences being within the numerical accuracy of an ODE solver and Fourier inversion
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
(of order 10−4 ). The difference in terms of implied volatilities does not exceed 10−4 .
In Fig. 10, we show model prices of call options on the QV computed using the
second-order affine expansion under the MMA and inverse measures, and the com-
parison with MC estimates. We use the same expiries as for the model calibration
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
and the same parameters reported in Eq. (6.4). We also observe that the solution
under the both measures are consistent and agree with MC simulations. Interest-
ingly, we observe that the implied volatility skew of options on the QV is upward
sloping, which is observed in market data of options on the QV and listed options on
VIX index. It is known that Heston model, even augmented with jumps, produces
downward sloping skews on options on the QV (see for an example Drimus (2012)
and Sepp (2012)). As a result, the log-normal SV model implies realistic patterns
of implied volatility-of-volatility.
7. Conclusion
We have considered the log-normal SV model with the quadratic drift which plays an
important role for the existence of martingale measures. We have shown that the
log-normal volatility process has the strong solution in spite of having quadratic
drift. We have derived admissible bounds of model parameters for the existence
of both the MMA measure and the inverse measures. Based on that, we applied
proposed SV model for modeling implied volatility surfaces of assets with positive
return–volatility correlation.
Since the log-normal SV model is not affine, there is no analytical solution for
the MGF in this model. To circumvent this, we have developed an affine expansion
approach under which the joint MGF of three state variables (log-price, the QV, and
the volatility) is decomposed into a leading term, which has exponential-affine form,
and a residual term, whose estimate depends on the higher order moments of the
volatility process. We have proved that the second-order leading term is consistent
with the expected values and the variances of the state variables. We have applied
Fourier inversion techniques for valuation of vanilla and inverse options on spot and
futures underlying. By comparison of model values computed using our method
with estimates obtained from MC simulations, we have shown that the second-order
leading term is very accurate for option valuation. We have demonstrated that our
model fits implied volatilities of assets with positive return–volatility correlation.
We have shown that the log-normal SV model fits well different market regimes by
2450003-50
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
calibrating model parameters to time series of options on Bitcoin for the past four
years.
We have derived backward Euler–Maruyama scheme for discretization of the
log-volatility process and show that it has a strong convergence rate of 1. Proposed
scheme for the simulation of the log-normal SV model is robust, computationally
efficient and does not require any special treatment in discretization.
As an extension of the frameworke we shall consider the applications to the
rough SV models proposed by Gatheral et al. (2018) using log-normal volatility. We
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
Acknowledgments
The authors are thankful to Vladimir Lucic for valuable insights and comments.
We thank Alan Lewis, Johannes Muhle-Karbe, Blanka Horvath, Kalev Pärna and
participants at Imperial College Finance and Stochastics Seminar, TU Munich
Oberseminar, ETH Seminar, EAJ Conference 2022 in Tartu for useful discussion
and comments. We are grateful for anonymous referees for helpful comments.
Appendix A. Proofs
A.1. Proof of Theorem 3.2
We consider the interval (l, r) and define the scale function and its density, denoted
by S(x) and s(x), respectively, as follows (see Chap. IV.15 in Borodin (2017)):
) x * x
2μ(σ)
s(x) = exp − dσ , S(x) = s(y)dy, (A.1)
c v(σ)2 c
where x ∈ (l, r) and c is an arbitrary but fixed interior point of (l, r). We define the
speed function and its density, denoted by M (x) and m(x), respectively, by
) x * x
2 2μ(σ)
m(x) = exp dσ , x ∈ (l, r), M (x) = m(y)dy. (A.2)
v(x)2 c v(σ)2 c
e Sepp & Rakhmonov (2023a) apply the developed solution for modeling of SV in Cheyette-type
interest rate model and obtain closed-form solution for valuation of swaptions under this model.
f The volatility beta is an interesting variable to make path-dependent as well. We observe that in
many markets the implied options skewness follows most recent price performance. For example,
in options on Bitcoin, the skewness implied from option prices becomes positive following strong
performance of Bitcoin.
2450003-51
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
x κ1 −κ2 θ
Choosing c such that c 2μ(σ)v(σ)2 dσ = 0 and setting η = 2 ϑ2 , we obtain
) * ) *
2κ1 θ 1 2κ2 2 −η−2 2κ1 θ 1 2κ2
s(x) = xη exp + x , m(x) = x exp − − x .
ϑ2 x ϑ2 ϑ2 ϑ2 x ϑ2
x x ) *
2κ1 θ 1
S(x) = s(y)dy ≤ y η exp dy, x ↓ 0+,
c c ϑ2 y
x x ) *
2κ2
S(x) = s(y)dy ≥ y η exp y dy, x ↑ +∞. (A.3)
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
c c ϑ2
Using upper and lower bounds in (A.3), we find that
S(0+) = lim S(x) = −∞, S(+∞) = lim S(x) = +∞. (A.4)
x↓0+ x↑+∞
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
We see that I1 is finite for all κ1 , κ2 > 0. Using asymptotics of incomplete gamma
function (see Eq. (6.5.32) in Abramowitz & Stegun (1972)), we obtain for y → +∞
+∞ ) * −1
−η−2 2κ2 2κ2
I2 = z exp − 2 z dz = y −η−2
y ϑ ϑ2
) *
2κ2 1
× exp − 2 y 1+O . (A.9)
ϑ y
Substituting asymptotic expansion (A.9) into Eqs. (A.8) and (A.7), we have
r ) * −1 r r
! 2κ2 2κ2 −2
Σ(r) = I1 y η exp y dy − y dy + O y −3 dy ,
x ϑ2 ϑ2 x x
y → +∞.
2450003-52
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
!
We conclude that Σ(+∞) = +∞. Hence, boundary r = +∞ is unattainable.
Next, we show that boundary l = 0 is unattainable as well. By definition of Σ(l)
x y x x
Σ(l) = s(z)dz m(y)dy = m(z)dz s(y)dy
l l l y
x x ) *
2 −η−2 2κ1 θ 1 2κ2
= 2
z exp − 2 − 2 z dz y η
l y ϑ ϑ z ϑ
) *
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
2κ1 θ 1 2κ2
× exp + 2 y dy.
ϑ2 y ϑ
We make an reciprocal transformation {z → 1/z, y → 1/y} to obtain
1/l y *
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
)
2 η 2κ1 θ 2κ2 1
Σ(l) = 2
z exp − 2 z − 2 dz y −η−2
1/x 1/x ϑ ϑ ϑ z
) *
2κ1 θ 2κ2 1
× exp y + dy.
ϑ2 ϑ2 y
The right-hand side of is finite if and only if
1/l y ) * ) *
! 2κ1 θ 2κ1 θ
Σ(l) := z exp − 2 z dz y −η−2 exp
η
y dy (A.10)
1/x 1/x ϑ ϑ2
is finite. Splitting inner integral into two, over [1/x, +∞) and [y, +∞), respectively
y ) * +∞ +∞
2κ1 θ
z η exp − 2 z dz = · · · dz − · · · dz = I1 − I2 , (A.11)
1/x ϑ 1/x y
we see that I1 if finite for all κ1 , κ2 > 0. As before, using asymptotic expansion of
incomplete gamma function we have for y → +∞:
+∞ ) * ) *
2κ1 θ ϑ2 2κ1 θ 1
I2 = z exp − 2 z dz =
η
y exp − 2 y
η
1+O .
y ϑ 2κ1 θ ϑ y
(A.12)
Substituting Eq. (A.12) into Eqs. (A.11) and (A.10), we have for y → +∞
1/l ) * 1/l
1/l
! −η−2 2κ1 θ ϑ2 −2 −3
Σ(l) = I1 y exp y dy − y dy + O y dy .
1/x ϑ2 2κ1 θ 1/x 1/x
!
We conclude that Σ(+∞) = +∞. Hence, boundary l = 0 is unattainable.
2450003-53
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
∂U
+ LU = 0, (t, x) ∈ Q, Q = (0, T ) × D,
∂t (A.13)
U (T, x) = φ(x), x ∈ D,
$
d
∂U 1 $ ik
d
∂2U
LU = bit (x) (t, x) + at (x) (t, x). (A.14)
i=1
∂xi 2 ∂xi ∂xk
i,k=1
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
We assume that matrix (aij t ) is nonnegative definite for every t. Under certain
restrictions on the coefficients of operator L and growth conditions on φ, it is possi-
ble to represent the solution of (A.13) by means of conditional expectation known
as a Feynman–Kac formula. However, standard results require uniform ellipticity
of the operator L, see Friedman (1975), which is not satisfied in our case, because
the QV It is degenerate. Thus, problem (A.13) becomes degenerate parabolic one.
We consider the multi-dimensional SDE
Proof. Consider sequence of bounded domains {Dn }∞ n=1 contained in D such that
-∞
n=1 Dn = D and Dn ⊆ {x ∈ R : x < n}. We fix x ∈ D and find n such that
d
2450003-54
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
∂U
E t,x
σij (Xs ) dWs(j) = 0 ⇒ u(x) = Et,x [u(ηn ∧ T, Xηn ∧T )].
t ∂xi
(A.16)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
−2
We show that P [τn ≤ T ] ≤ c1 (1 + x )n
t,x 2
. Using Markov inequality and
estimate from Karatzas & Shreve (1991), p. 306:
Thus limn→+∞ u(ηn ∧ T, Xηn∧T ) = φ(XT ) a.s. It implies that u(t, x) is bounded
by the maximum principle. Finally, by the dominated convergence theorem:
lim E[u(ηn ∧ T, Xηn ∧T )] = E[φ(XT )]. (A.17)
n→+∞
2450003-55
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
[2] [2]
$
8
(k) [2]
−RΘ,τ + L(Y ) RΘ = − Y k [CΘ (τ ; Θ)E [2] + C (k) (τ ; Θ)EΘ ], (A.22)
k=5
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
[2]
subject to boundary condition RΘ (0, Y ; Θ) = 0, where C (k) (τ ; Θ) are given by
Eq. (4.27). We note that A(k) (τ ; Θ) vanish at Θ = 0.
A(k) (τ ; Θ = 0) = 0. (A.23)
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
(k)
Calculating CΘ (τ ; Θ) and taking into account conditions (A.23) at Θ = 0
# (4)
(k) −4κ2 AΘ (τ ; Θ)|Θ=0 , k = 5,
CΘ (τ ; Θ)|Θ=0 = (A.24)
0, k ≥ 6.
We verify that last two functions in solution of Eq. (A.26) are zero for κ2 = 0:
A(3) (τ ; Θ) ≡ A(4) (τ ; Θ) ≡ 0. (A.27)
[2]
Now, taking the second derivative w.r.t. Θ in (4.26), we obtain for RΘΘ
[2] [2]
−RΘΘ,τ + L(Y ) RΘΘ
$
8
(k) (k) [2] [2]
=− {Y k [CΘΘ (τ ; Θ)E [2] + 2CΘ (τ ; Θ)EΘ + C (k) (τ ; Θ)EΘΘ ]}
k=5
2450003-56
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
[2]
with RΘΘ (0, Y ; Θ) = 0. Taking the second derivative of C (k) (τ ; Θ) in (4.27) w.r.t.
Θ, and accounting for conditions (A.23) and (A.27) at Θ = 0, we find that
# (4)
(k) −4κ2 AΘ (τ ; Θ)|Θ=0 , k = 5,
CΘΘ (τ ; Θ)|Θ=0 = (A.28)
0, k ≥ 6.
[2]
E [2] (τ ; Θ)|Θ=0 = 1, EΘ (τ ; Θ)|Θ=0 = −Y. (A.29)
[2]
Thus, when κ2 = 0, RΘΘ (τ, Y ; Θ = 0) ≡ 0, and E [2] (τ ; Θ) reproduces the variance
of Yt .
∂ $ (k)
8
[2] [2]
−RΦ,τ + L(Y ) RΦ = − C (τ ; Φ)E [2] Y k
∂Φ
k=5
$
8
(k) [2] (A.31)
=− Y k [CΦ (τ ; Φ)E [2] + C (k) (τ ; Φ)EΦ ],
k=5
[2]
RΦ (0, Y ; Φ) = 0,
2450003-57
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
To show that the second-order expansion reproduces the variance of the log-
(k)
price, we differentiate Eq. (4.25) with respect to Φ and find that AΦ (τ, Φ) solve
the system of ODEs (A.35) at Φ = 0 as functions of τ :
∂ 2 $ (k)
8
[2] [2]
−RΦΦ,τ +L (Y )
RΦΦ =− 2 C (τ ; Φ)E [2] Y k
∂Φ
k=5
$
8
(k) (k) [2] [2]
=− {Y k [CΦΦ (τ ; Φ)E [2] + 2CΦ (τ ; Φ)EΦ + C (k) (τ ; Φ)EΦΦ ]},
k=5
[2]
× RΦΦ (0, Y ; Φ) = 0. (A.37)
(k)
Calculating CΦΦ (τ ; Φ) in (4.27) and taking into account boundary conditions
(A.32) at Φ = 0, we find that
# (4)
(k) −4κ2 AΦ (τ ; Φ)|Φ=0 , k = 5,
CΦΦ (τ ; Φ)|Φ=0 = (A.38)
0, k ≥ 6.
Combining conditions (A.38), (A.33), and (A.39), we are able to simplify prob-
lem (A.34) at Φ = 0 as follows:
[2] [2] (4) (4) [2]
−RΦΦ,τ + L(Y ) RΦΦ = 4κ2 (Y 5 AΦΦ − 2Y 6 AΦ ), RΦΦ (0, Y ; Φ) = 0. (A.40)
[2]
Thus, when κ2 = 0, RΦΦ (τ, Y ; Φ = 0) ≡ 0, so that E [2] (τ, Y ; Φ) is consistent with
the variance of Xt .
2450003-58
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
solution A(τ ; Γ) is defined on maximal interval of existence [0, τ+ (Γ; A)) such that
either τ+ (Γ; A) = +∞ or τ+ (Γ; A) < +∞ and A(τ ; Γ) → +∞ as τ ↑ τ+ (Γ; A).
We combine d nonlinear ODEs in (A.41) into single (nonlinear) vector ODE:
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
Lemma A.2. There exists finite and nonnegative function g such that
t Rs
A(τ ; Γ) ≤ A0 + A0
2 2 2
h(s)e 0 h(r)dr ds, h(s) = g( (A(τ ; Γ)).
0
(A.44)
Proof. The proof is essentially contained in Keller-Ressel & Mayerhofer (2015) and
based on the application of Gronwall’s inequality.
2450003-59
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
G(τ ; X, I, Y ; αΦ, Ψ = 0, Θ = 0)
# %
$2n
= exp −αΦX + Ak (τ ; αΦ, Ψ = 0, Θ = 0)Y k
k=0
ORCID
Artur Sepp https://orcid.org/0000-0002-7038-1748
Parviz Rakhmonov https://orcid.org/0000-0001-9571-7378
References
M. Abramowitz & I. A. Stegun (1972) Handbook of Mathematical Functions with Formulas,
Graphs, and Mathematical Tables, Applied Mathematics Series. National Bureau of
Standards.
D. Ackerer & D. Filipović (2020) Option pricing with orthogonal polyolynomial expansions,
Mathematical Finance 30 (1), 47–84.
C. Alexander, D. Chen & A. Imeraj (2023) Crypto quanto and inverse options, Mathe-
matical Finance 33(4), 1005–1043.
E. Alòs, J. Gatheral & R. Radoicic (2020) Exponentiation of conditional expectations
under stochastic volatility, Quantitative Finance 20, 13–27.
2450003-60
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
A. Alfonsi (2013) Strong order one convergence of a drift implicit Euler scheme: Application
to the CIR process, Statistics & Probability Letters 83(2), 602–607.
G. Bakshi & N. Kapadia (2003) Delta-hedged gains and the negative market volatility risk
premium, The Review of Financial Studies 16(2), 527–566.
J. Baris, P. Baris & B. Ruchlewicz (2008) Blow-up solutions of quadratic differential sys-
tems, Journal of Mathematical Sciences 149 (4), 1369–1375.
O. E. Barndorff-Nielsen & A. N. Shiryaev (2015) Change of Time and Change of Measure,
Vol. 21. World Scientific Publishing Company.
G. Bakshi, J. Ju & H. Ou-Yang (2006) Estimation of continuous-time models with an appli-
by 130.0.122.228 on 10/26/24. Re-use and distribution is strictly not permitted, except for Open Access articles.
2450003-61
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
cations to bond and currency options, The Review of Financial Studies 6(2), 327–
343.
D. H. Jacobson (1974) Extensions of Linear-Quadratic Control; Optimalization and Matrix
Theory. London: Academic Press.
B. Jørgensen (1982) Statistical Properties of the Generalized Inverse Gaussian Distribution.
Springer Science & Business Media.
P. Karasinski & A. Sepp (2012) Beta stochastic volatility model, Risk Magazine 22(9)
66–71.
I. Karatzas & S. E. Shreve (1991) Brownian Motion and Stochastic Calculus, Graduate
Texts in Mathematics. New York: Springer-Verlag.
M. Keller-Ressel & E. Mayerhofer (2015) Exponential moments of affine processes, The
Annals of Applied Probability 25(2), 714–752.
R. A. Jarrow & M. Larsson (2012) The meaning of market efficiency, Mathematical
Finance: An International Journal of Mathematics, Statistics and Financial Eco-
nomics 22(1), 1–30.
A. L. Lewis (2000) Option Valuation Under Stochastic Volatility. Finance Press.
A. L. Lewis (2018) Exact solutions for a GBM-type stochastic volatility model having a
stationary distribution. arXiv:1809.08635.
P.-L. Lions & M. Musiela (2007) Correlations and bounds for stochastic volatility models,
Annales de l’Institut Henri Poincaré C, Analyse non linéaire 24(1), 1–16.
A. Lipton (2001) Mathematical Methods for Foreign Exchange: A Financial Engineer’s
Approach. World Scientific.
A. Lipton (2002) The vol smile problem, Risk Magazine 61–65.
A. Lipton and A. Reghai (2023), ‘SPX, VIX and Scale-Invariant LSV Local Stochastic
Volatility’, Wilmott Magazine, February 78–84.
V. Lucic (2021) BTC inverse call and the standard FX framework, SSRN,
https://ssrn.com/abstract=4113726.
V. Lucic & A. Sepp (2023) Valuation and hedging of cryptocurrency inverse
options: With backtest simulations using Deribit options data, SSRN, https://
ssrn.com/abstract=4606748.
J. Masoliver & J. Perelló (2006) Multiple time scales and the exponential Ornstein–
Uhlenbeck stochastic volatility model, Quantitative Finance 6(5), 423–433.
R. C. Merton (1973) Theory of rational option pricing, The Bell Journal of Economics
and Management Science 4(1) 141–183.
M. Musiela & M. Rutkowski (2009) Martingale Methods in Financial Modelling, Vol. 36.
Springer Science & Business Media.
2450003-62
March 6, 2024 15:25 WSPC/S0219-0249 104-IJTAF SPI-J071
2450003
J. Ruf (2015) The martingale property in the context of stochastic differential equations,
Electron. Commun. Probab 20. 1–10.
A. Sepp (2007) Variance swaps under no conditions, Risk Magazine 20(1) 82–87.
A. Sepp (2008) Pricing options on realized variance in the Heston model with jumps in
returns and volatility, Journal of Computational Finance 11(4), 33–70.
Int. J. Theor. Appl. Finan. 2023.26. Downloaded from www.worldscientific.com
A. Sepp (2012) Pricing options on realized variance in the Heston model with jumps in
returns and volatility — Part II: An approximate distribution of discrete variance,
Journal of Computational Finance 16(22), 3–32.
A. Sepp & P. Rakhmonov (2023) A robust stochastic volatility model for interest rates,
Risk Magazine, September 2023, 1–6, https://ssrn.com/abstract=4315906.
A. Sepp & P. Rakhmonov (2023) What is a robust stochastic volatility model, SSRN,
https://ssrn.com/abstract=4647027.
N. Shephard (2005) Stochastic Volatility: Selected Readings. Oxford University Press.
C. A. Sin (1998) Complications with stochastic volatility models, Advances in Applied
Probability 30(1), 256–268.
M. Tegner and R. Poulsen, Volatility is log-normal — But not for the reason you think,
Risk Magazine 6(2), 1–16.
2450003-63