Next Article in Journal
ESG Disclosure and Firm Performance: An Asset-Pricing Approach
Previous Article in Journal
Investigating Causes of Model Instability: Properties of the Prediction Accuracy Index
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Stochastic Volatility in the Generalized Black-Scholes-Merton Model

Laboratory of Control under Incomplete Information, V.A. Trapeznikov Institute of Control Sciences of RAS, Profsoyuznaya 65, 117997 Moscow, Russia
Risks 2023, 11(6), 111; https://doi.org/10.3390/risks11060111
Submission received: 25 April 2023 / Revised: 30 May 2023 / Accepted: 6 June 2023 / Published: 8 June 2023

Abstract

:
This paper discusses the generalized Black-Scholes-Merton model, where the volatility coefficient, the drift coefficient of stocks, and the interest rate are time-dependent deterministic functions. Together with it, we make the assumption that the volatility, the drift, and the interest rate depend on a gamma or inverse-gamma random variable. This model includes the models of skew Student’s t- and variance-gamma-distributed stock log-returns. The price of the European forward-start call option is derived from the considered models in closed form. The obtained formulas are compared with the Black-Scholes formula through examples.

1. Introduction

The standard Black-Scholes-Merton model (see Black and Scholes 1973; Merton 1973) assumes that the volatility coefficient is constant over the contract time. However, when the parameters of the model are calibrated with respect to the market option prices, one may find out that the volatility is a function of the exercise time and the strike coefficient. The dependence curve of the implied volatility of stock prices has the form of a convex function. This effect is called the “volatility smile”. Taking it into account, Merton (1973) proposed to regard the drift and the volatility coefficients as functions of time. The problem of detecting the term structure of volatility is widely discussed in the literature.
Derman and Kani (1994) and Dupire (1994) suggested a general diffusion model for volatility under the hypothesis that it is a function of state and time. This model is called the local volatility model. Its multidimensional extensions were introduced and considered in the papers by Brigo and Mercurio (2002) and Brigo et al. (2003). Results in the local volatility model that are close to the Black-Scholes formula are given in the monograph by Oksendal (2003).
Diffusion stochastic volatility models infer that the stock price and the volatility coefficient are driven by correlated Brownian motions. A survey of the early papers in this direction is given in Section 7.4.2 of the monograph by Musiela and Rutkowski (2005). Among more recent papers, let us mention those where combined models are considered. The Heston-Hull-White models are studied by Grzelak and Oosterlee (2011), Levendis and Maré (2022) and Liu et al. (2023). The Heston-Cox-Ingersoll-Ross model is discussed in Cao et al. (2016) and Mao et al. (2022).
Modern research uncovers a more complicated term structure for volatility. Non- parametric methods of volatility assessment are given in Alghalith et al. (2020), Cuchiero et al. (2020), and Liu et al. (2019) (neural network estimation). The fat-tailedness of the volatility is detected by Mahieu and Schotman (1998) and Kim et al. (1998). Jumps of finite activity of the stock and the stock index volatilities are discovered by Bates (1996), Andersen et al. (2007), Eraker et al. (2003), Taylor et al. (2018). The stochastic volatility modeled by the inverse-gamma distribution is investigated in Nakajima and Omori (2012), Liu (2021) and Men et al. (2021). Nakakita and Nakatsuma (2021) modeled the stochastic volatility by the both gamma and inverse-gamma distributions. Processes with Lévy jumps of infinite activity are used for the stochastic volatility modeling as well. Namely, the gamma process is considered in this direction by Nzokem (2023). The parameters of the distributions are estimated in these papers for a variety of financial indices using the maximum likelihood technique and the methods of Bayesian statistics.
Together with the volatility coefficient, the Black-Scholes-Merton model implies that the drift coefficient of the stock log-return is also constant. However, the financial data often suggests the presence of jumps in the log-return. Andersen et al. (2001) include in the model of stock log-returns jumps of the general form. Several sources discuss Lévy’s jumps of finite activity. Shackleton et al. (2010) and Hong et al. (2023) augment the Heston and the lognormal Ho-Lee models with compound Poisson jumps. Chib et al. (2002) discuss stochastic volatility models with jumps whose distribution is generated by the Bernoulli distribution. Nakajima and Omori (2012) and Nakakita and Nakatsuma (2021) study the skew Student’s t-distribution as a model for stock prices. Being the normal-inverse mixture with the inverse-gamma mixing density, the skew Student’s t-distribution has the stochastic mean, which is modeled by the inverse-gamma distribution.
Moreover, processes with jumps of infinite activity are also discussed for asset drift modeling. The skew Student’s t-process is investigated as a model for financial index dynamics in the form of subordinated Brownian motion with drift by Aas and Haff (2006), Bibby and Sørensen (2003), Finlay and Seneta (2006), and Finlay and Seneta (2008). Aas and Haff (2006) study the asymptotic properties of the skew Student’s t-distribution and estimate its parameters in their application to currency exchange rate modeling. Bibby and Sørensen (2003) compare the properties of various generalized hyperbolic distributions, including the skew Student one, and propose again that the skew Student’s t-process is a good fit for the exchange rate simulation. Finlay and Seneta (2006) approve the skew Student’s t-process for the modeling of the S&P500 data. Finlay and Seneta (2008) develop different techniques for parameter estimation in the skew Student’s t-model and discuss the modeling of the S&P500 index and the oil prices. The variance-gamma process (the variance-gamma distribution is the normal-inverse mixture with the gamma mixing density, and hence the variance-gamma process has the stochastic drift modeled by the gamma process) is considered in Madan et al. (1998), Seneta (2004), Daal and Madan (2005), Ivanov (2018), Ivanov (2022), Linders and Stassen (2016), and Mozumder et al. (2015), among others. Madan et al. (1998) summarize the basic properties of the variance-gamma distribution and suggest the method of receiving analytical results in the variance-gamma model. Daal and Madan (2005) confirm the use of the variance-gamma model for the exchange rate simulation. Linders and Stassen (2016) model with the variance-gamma process the Dow Jones index dynamics. Mozumder et al. (2015) study the S&P500 index options in the variance-gamma model. Ivanov (2022) proceeds from the ideas of Madan et al. (1998) and obtains closed-form results for an extension of the variance-gamma model.
The model of this paper relates to the generalized Black-Scholes-Merton model of Section 7.1.10 of Musiela and Rutkowski (2005) and Chapter 12.3 of Oksendal (2003). We assume that the drift and volatility of the stock depend on the gamma or the inverse-gamma distribution, which are independent of the Brownian motion. It appears that it is possible in this model to obtain the price of a forward-start call option in closed form. We exploit the methodology of integral transformations of special functions in their application to mathematical finance, which was introduced in Madan et al. (1998) and then developed in particular in Ano and Ivanov (2016). A comparison of the results with the Black-Scholes formula is given in the section on numerical examples.

2. Materials and Methods

In this section, we present a mathematical model in which we formulate our results. Together with it, necessary definitions and a theoretical background are given. Used designations are also included.

2.1. Model

Throughout the paper, we designate as ( S t ) t T and ( R t ) t T the stock price and the bank account value dynamics over a period of time [ 0 , T ] . Set 0 T 0 < T . A forward-start option (see Section 6.2 of the monograph by Musiela and Rutkowski (2005)) is a contract in which the holder receives at time T 0 an option with expiry date T and exercise price K S T 0 with some strike coefficient K > 0 . On the other hand, the holder must pay at time 0 an up-front fee, the price of the option. Thus, a forward-start call option has payoffs.
( S T K S T 0 ) +
at the maturity T.
Next, we give a mathematical specification of the discussed model. Let Ω , F , F t t T , P be the probability space with filtration F t t T . We assume that the Brownian motion B t t T , B 0 = 0 , and the random variable z are defined on this probability space and the both processes B t t T and z t t T with
z t = I t < t 0 + z I t t 0 , t 0 0 ,
are adapted to the filtration F t t T . It is supposed that the processes are independent with each other.
If t 0 = 0 , the additional uncertainty that is generated by the random variable z relates to the “volatility smile” effect. When t 0 > 0 , the additional uncertainty appears because of supplemental factors. There can be the economic annual reporting, the political shifts, the microeconomic merger decisions among those factors. Then t 0 is suggested to be larger than the expected time of the factor forthcoming. If, for example, we discuss the annual reporting factor, it is obviously fair to assume that the time t 0 is not random but actually fixed.
The generalized Black-Scholes-Merton model implies that the stock volatility is random and even not necessarily adapted to the filtration generated by the underlying Brownian motion (see Section 7.1.10 of Musiela and Rutkowski (2005)). We consider the generalized Black-Scholes-Merton model under the assumption that the logarithm of the stock price ( S t ) t T follows an equation.
log S t / S 0 = 0 t μ ˜ s d s + 0 t θ ˜ s z s d s + 0 t σ ˜ s z s d B s ,
where ( μ ˜ t ) t T , ( θ ˜ t ) t T and ( σ ˜ t ) t T are time-dependent deterministic processes and σ ˜ t > 0 . From the economic point of view, the independence between z t and B t relates to the situation when the volatility randomness is induced by macroeconomic factors which are independent with the firm microstructure development.
The logarithm of the bank account value is suggested to evaluate via an equation
log R t = 0 t r ˜ s d s + 0 t ρ ˜ s z s d s
with time-dependent deterministic ( r ˜ t ) t T , r ˜ t 0 , and ( ρ ˜ t ) t T , ρ ˜ t 0 . The new terms after the time t 0 take into account both the ordinary Black-Merton-Scholes uncertainty and the supplemental uncertainty, which is indicated by the random variable z. The model concerns a the small or average additional indeterminacy of financial markets.
Let the initial probability measure P be a martingale measure for the discounted stock price process S t / R t t T . That is, we assume that
μ ˜ t + θ ˜ t = r ˜ t + ρ ˜ t σ ˜ t 2 2 , t 0 ,
and additionally
μ ˜ t = r ˜ t , θ ˜ t = ρ ˜ t σ ˜ t 2 2 , t t 0 .
Indeed, then
S t / R t = S 0 exp 0 t σ ˜ s z s d B s 1 2 0 t σ ˜ s 2 z s d s
and hence the martingale property is satisfied. The problem of equivalent martingale change of measure for semimartingales is widely discussed in the literature. We refer, in particular, to Subsection III.3 of the monograph by Jacod and Shiryaev (1987) and to the papers by Kallsen and Shiryaev (2002) and Eberlein et al. (2009).
Example 1. 
If we set t 0 = , then the model (2) and (3) becomes the diffusion Black-Scholes-Merton model (see Chapter 12.3 of Oksendal (2003) and Chapter VIII.1 of Shiryaev (1999)) with
log S t B S / S 0 B S = 0 t μ ˜ s + θ ˜ s d s + 0 t σ ˜ s d B s
and
log R t B S = 0 t r ˜ s + ρ ˜ s d s .
Example 2. 
Let t 0 = 0 . Then
log S 1 / S 0 = 0 1 μ ˜ s d s + z 0 1 θ ˜ s d s + z 0 1 σ ˜ s d B s
and therefore the model (2) can be discussed as an extension of the variance-gamma and the skew Student’s t-models in the sense of the stock log-return behavior. Indeed, if we set μ ˜ t μ , θ ˜ t θ and σ ˜ t σ in (8), then
log S 1 / S 0 = μ + θ z + σ z N ,
where N = N ( 0 , 1 ) is the standard normal random variable. Hence the log-return (9) is variance-gamma distributed if z is gamma distributed and skew Student’s t-distributed if z is inverse-gamma distributed.
Due to the initial probability measure is assumed to be martingale for the discounted stock price, the risk-neutral forward-start call option price C F S can be computed as
C F S = E R T 1 ( S T K S T 0 ) + .
We use in the subsequent sections the designations S t G and S t I G for the stock prices in the gamma and inverse-gamma models. If T 0 = 0 , then a forward-start call option becomes a standard European call one with the price
C = E R T 1 ( S T K S 0 ) + .

2.2. Special Functions

Let us define a complementary function I = I ( a , b , u 1 , u 2 , u 3 ) for a > 0 , b > 0 , u 1 , u 2 , u 3 R by the identity
I ( a , b , u 1 , u 2 , u 3 ) = Γ a + 1 2 2 π ( b + u 1 ) a [ B 1 2 , a 2 + + u 2 b + u 1 F a + 1 2 , 1 2 , 3 2 ; u 2 2 2 ( b + u 1 ) ] I u 3 = 0 + + | s | a 1 2 e s ( 1 + q ) a 2 π ( b + u 1 ) a [ B ( a , 1 ) ( | s | K a + 1 2 | s | + s K a 1 2 | s | ) × × Φ 1 a , 1 a , a + 1 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( a + 1 , 1 ) × × K a 1 2 | s | Φ 1 a + 1 , 1 a , a + 2 ; 1 + q 2 , s ( 1 + q ) ] I u 3 0
with
q = q ( u 1 , u 2 ) = u 2 u 2 2 + 2 ( b + u 1 )
and
s = s ( u 1 , u 2 , u 3 ) = u 3 u 2 2 + 2 ( b + u 1 ) .
As it is shown further in the paper, this function determines the forward-start option price in the both gamma and inverse-gamma volatility models. In (10), we denote as
Γ ( u ) , B ( u 1 , u 2 ) , K u 1 ( u 2 )
the gamma function, the beta function and the MacDonald function (the modified Bessel function of the second kind), respectively. Furthermore, we use in (10) the designation
F ( u 1 , u 2 , u 3 ; u 4 )
for the hypergeometric Gauss function and the denotation
Φ 1 ( u 1 , u 2 , u 3 ; u 4 , u 5 )
for the degenerate Appell function (or the Humbert series) which is the double sum
m = 0 n = 0 ( u 1 ) m + n ( u 2 ) m m ! n ! ( u 3 ) m + n u 4 m u 5 n
for | u 4 | < 1 , where ( u ) l , l N { 0 } , is the Pochhammer’s symbol. For more information on these special mathematical functions and their integral representations we refer to the monographs by Bateman and Erdélyi (1953), Srivastava and Karlsson (1985) and the handbooks by Erdélyi et al. (1954), Gradshteyn and Ryzhik (2007).
Together with it, we set the deterministic functions c 1 , c 2 , c 3 , c 4 by the identities
c 1 = c 1 ( T 0 ) = T 0 T σ ˜ s 2 2 + ρ ˜ s d s T 0 T σ ˜ s 2 d s ,
c 2 = c 2 ( T 0 ) = T 0 T r ˜ s d s log K T 0 T σ ˜ s 2 d s ,
c 3 = c 3 ( T 0 ) = T 0 T ρ ˜ s σ ˜ s 2 2 d s T 0 T σ ˜ s 2 d s ,
c 4 = c 4 ( T 0 ) = T 0 T ρ ˜ s d s ,
c 5 = c 5 ( T 0 ) = 1 2 T 0 T σ ˜ s 2 d s .

3. Results

The two subsections below introduce formulas for the prices of forward-start European call options in the models with gamma and inverse-gamma volatilities. As corollaries, the prices of standard European calls are derived. The prices of the related put options could be computed via the call-put parity. For details, see Eberlein et al. (2008) and Herdegen and Schweizer (2018).

3.1. Gamma Volatility

It is assumed throughout this subsection that the random variable z in (1) is gamma distributed, that is z = γ = γ ( a , b ) , where the parameters a > 0 and b > 0 . The properties of gamma distribution are considered in the Section 5.3.3 of the monograph by Schoutens (2003) and in the papers by Madan et al. (1998), Rathgeber et al. (2016), among others. The gamma distribution has the probability density function
f γ ( x ) = b a x a 1 e b x Γ a , x > 0 .
The characteristic function of the gamma random variable is
φ γ ( u ) = E e i u γ = b b i u a .
It has the mean and the variance
E γ = a b and V γ = E γ E γ 2 = a b 2 ,
respectively.
The next theorem gives us the price of the European forward-start call option in the gamma volatility model.
Theorem 1. 
Assume that T 0 t 0 . Then the forward-start call option price in the gamma model
C G F S = = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 1 , c 2 ) K exp T 0 T r ˜ s d s I ( a , b , c 4 , c 3 , c 2 ) ) ,
where the function I is defined in (10) and c 1 , c 2 , c 3 , c 4 are set by (11), (12), (13), (14), respectively.
Remark 1. 
In the classical Black-Scholes-Merton framework, the forward-start call option price is computed through the conditional expectation with respect to S T 0 , that is
C B S F S = R T 1 E S T K S T 0 + = R T 1 E E S T K S T 0 + | S T 0 = = R T 0 R T E S T 0 R T 0 E S T S T 0 K + | S T 0 = S 0 R T T 0 E S T T 0 K +
since S T S T 0 1 and S T 0 are independent with each other, where
R T T 0 E S T T 0 K +
is again a Black-Scholes price. If T 0 > t 0 , then this direct approach does not work in our model because of S T S T 0 1 and S T 0 both depend on z. Hence we pass to the conditional expectations with respect to z in the proof of Theorem, see (A2) in Appendix A.
As a straightforward corollary of Theorem 1, we get the standard European call price.
Corollary 1. 
Let r ˜ t r , ρ ˜ t ρ , σ ˜ t σ . Then the standard European call price in the gamma model
C G = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 1 , c 2 ) K e r T I ( a , b , c 4 , c 3 , c 2 ) )
with
c 1 = ρ σ + σ 2 T , c 2 = r T log K σ T ,
c 3 = ρ σ σ 2 T , c 4 = ρ T .

3.2. Inverse Gamma Volatility

We suggest in this subsection that z in (1) has the inverse-gamma distribution ς = ς ( a , b ) , a > 0 , b > 0 . This distribution may be used, in particular, for the modeling of the skew Student’s t-distribution through a normal mean-variance mixture (see McNeil et al. (2005)). The probability density function of the inverse-gamma distribution is
f ς ( x ) = b a Γ ( a ) x a 1 e b x , x > 0 .
with a , b > 0 . It has the characteristic function (see, for example, Witkovský (2001))
φ ς ( u ) = 2 ( i u b ) a 2 Γ ( a ) K a 2 i u b .
The inverse-gamma distribution has the mean
E ς = b a 1 I a > 1 + I a 1
and the variance
V ς = b 2 ( a 1 ) 2 ( a 2 ) I a > 2 + I a 2 .
The next theorem computes the forward-start call option price in the model with inverse-gamma volatility.
Theorem 2. 
Let T 0 t 0 and ρ ˜ t 0 for t T 0 . Then the forward-start call option price in the inverse-gamma model
C I G F S = = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 2 , c 5 ) K exp T 0 T r ˜ s d s I ( a , b , 0 , c 2 , c 5 ) ) ,
where I , c 2 and c 5 are defined in (10), (12) and (15).
A corollary below determines the standard European call price.
Corollary 2. 
Assume that r ˜ t r , σ ˜ t σ . Then the standard European call price in the inverse-gamma model
C I G = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 2 , c 5 ) K e r T I ( a , b , 0 , c 2 , c 5 ) )
with
c 2 = r T log K σ T , c 5 = σ T 2 .

3.3. Discussion

We have established the new analytical expressions that give us the formulas for the forward-start call option prices in the generalized Black-Scholes-Merton model, additionally driven by the gamma and the inverse-gamma distributions. These distributions are widely regarded in the literature for their aim of modeling different financial indices. One may mention the papers by Luciano and Schoutens (2016), Moosbrucker (2006), Wallmeier and Diethelm (2012), Göncü et al. (2016) on exploiting the gamma distribution, and the works by Fung and Seneta (2010), Takahashi et al. (2021), and Nakajima (2020) on using the inverse-gamma one in the stock market simulation.
The considered models extend the generalized Black-Scholes-Merton, the variance-gamma and the skew Student’s t-models. The obtained formulas depend on the values of special mathematical functions but can be computed over 0.5 s on modern software. The prices of forward-start put options can be calculated by exploiting the duality principle (Eberlein et al. 2008; Herdegen and Schweizer 2018). Future research may relate to the computation of the prices of exotic stock options and options on bonds in the discussed models. Furthermore, developments of these models based on the assumption of the linear drift jump (see Ivanov 2022) could be considered. Since the Bermudan options can be viewed as a spread of the forward-start contracts (see Schweizer 2002), the results of the paper may be used for the American-style options pricing as well.

4. Numerical Examples

In this section, we compare the prices of standard European call options in the models with the gamma and the inverse-gamma volatilities with the Black-Scholes formula. The idea is to pick the parameters of the gamma and the inverse-gamma distributions so that they even out the maximal number of the characteristics of the underlying processes. In fact, it has become possible to get the identities
E S t I G = E S t G = E S t B S ,
E log S t I G = E log S t G = E log S t B S
and
V log S t G = V log S t B S .
Furthermore, the parameters of the distributions are suggested, and the option prices are computed for the three models in separate subsections.

4.1. Black-Scholes Price

We discuss in this subsection the Black-Scholes-Merton model (6) and (7) under the condition (4) assuming that S 0 B S = 1 , r ˜ t = ρ ˜ t 0 and σ ˜ t 1 . Then
log S t B S = B t t 2
and the variance of the stock price
V S t B S = E S t B S 2 1 = e t 1 .
The logarithm of the stock price (27) has the mean
E log S t B S = t 2
and the variance
V log S t B S = t .
Let T 0 = 0 and T = 1 . Then the standard European call option price is
C B S = E S 1 B S K + = N 1 2 log K K N 1 2 log K ,
where N ( u ) is the normal distribution function.

4.2. Gamma Volatility Price

It is assumed that in the gamma volatility model of Section 3.1  S 0 G = 1 , r ˜ t = ρ ˜ t 0 and σ ˜ t 1 . Then we have that
log S t G = γ B t γ t 2
in this case and
V S t G = E e γ t 1 = b b t a 1 I t < b + I t b
from (17). The logarithm of the stock price has the mean
E log S t G = t E γ 2 = a t 2 b
and the variance
V log S t G = E γ B t γ t 2 2 a 2 t 2 4 b 2 = a t b + t 2 E γ 2 4 a 2 t 2 4 b 2 = a t b + t 2 a 4 b 2
in point of (18). We have that
E S t G = E S t B S = 1
and if we set a = b = 1 , then it follows from (28) and (31) that
E log S t G = E log S t B S = t 2 .
Let t 0 = T 0 = 0 and T = 1 . Then according to (20)
C G = I ( a , b , 0 , c 1 , c 2 ) K I ( a , b , 0 , c 3 , c 2 )
with c 1 = 1 2 , c 2 = log K , c 3 = 1 2 . Since a = b = 1 , we get that
C G = I 1 , 1 , 0 , 1 2 , k K I 1 , 1 , 0 , 1 2 , k ,
where k = log K .
It follows immediately from (10) that
I 1 , 1 , 0 , 1 2 , k = Γ 3 2 2 π B 1 2 , 1 2 + 1 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + + | s | 1 2 e s ( 1 + q ) 2 π [ B ( 1 , 1 ) ( | s | K 3 2 | s | + s K 1 2 | s | ) × × Φ 1 , 0 , 2 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( 2 , 1 ) × × K 1 2 | s | Φ 2 , 0 , 3 ; 1 + q 2 , s ( 1 + q ) ] I K 1
with
q = c 1 c 1 2 + 2 = 1 3 and s = c 2 c 1 2 + 2 = 3 k 2
and
I 1 , 1 , 0 , 1 2 , k = Γ 3 2 2 π B 1 2 , 1 2 1 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + + | s | 1 2 e s ( 1 + q ) 2 π [ B ( 1 , 1 ) ( | s | K 3 2 | s | + s K 1 2 | s | ) × × Φ 1 , 0 , 2 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( 2 , 1 ) × × K 1 2 | s | Φ 2 , 0 , 3 ; 1 + q 2 , s ( 1 + q ) ] I K 1
with
q = c 3 c 3 2 + 2 = 1 3 and s = c 2 c 3 2 + 2 = 3 k 2 .
We get from (33) that
I 1 , 1 , 0 , 1 2 , k = 1 2 + 1 4 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + + | k | 1 2 3 K 3 2 π [ { | k | K 3 2 3 | k | 2 k K 1 2 3 | k | 2 } Φ 1 , 0 , 2 ; 2 3 , 2 k + + 2 k 3 K 1 2 3 | k | 2 Φ 2 , 0 , 3 ; 2 3 , 2 k ] I K 1
and from (34) that
I 1 , 1 , 0 , 1 2 , k = 1 2 1 4 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + + | k | 1 2 3 2 K 3 2 π [ { | k | K 3 2 3 | k | 2 k K 1 2 3 | k | 2 } Φ 1 , 0 , 2 ; 1 3 , k + + k 3 K 1 2 3 | k | 2 Φ 2 , 0 , 3 ; 1 3 , k ] I K 1 .
Since
Φ u 1 , u 2 , u 3 ; u 4 , u 5 = = 1 B u 1 , u 3 u 1 0 1 x u 1 1 ( 1 x ) u 3 u 1 1 ( 1 u 4 x ) u 2 e u 4 x d x
for u 1 > 0 , u 3 > u 1 and u 4 < 1 according to formula 4.3.24 of Erdélyi et al. (1954), we get that
Φ 1 , 0 , 2 ; A , B = J 0 ( B )
and
Φ 2 , 0 , 3 ; A , B = 2 J 1 ( B ) ,
where
J 0 ( B ) = 0 1 e B x d x and J 1 ( B ) = 0 1 x e B x d x .
Hence we obtain that
I 1 , 1 , 0 , 1 2 , k = = 1 2 + 1 4 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + | k | 1 2 3 K 3 2 π × × [ { | k | K 3 2 3 | k | 2 k K 1 2 3 | k | 2 } J 0 ( 2 k ) + 4 k 3 K 1 2 3 | k | 2 J 1 ( 2 k ) ] I K 1
and
I 1 , 1 , 0 , 1 2 , k = = 1 2 1 4 2 F 3 2 , 1 2 , 3 2 ; 1 8 I K = 1 + | k | 1 2 3 2 K 3 2 π × × [ { | k | K 3 2 3 | k | 2 k K 1 2 3 | k | 2 } J 0 ( k ) + 2 k 3 K 1 2 3 | k | 2 J 1 ( k ) ] I K 1 .

4.3. Inverse-Gamma Volatility Price

We set in the model of Section 3.2  S 0 I G = 1 , r ˜ t = ρ ˜ t 0 and σ ˜ t 1 . Then
log S t I G = ς B t ς t 2
and it issues from (22) that
V S t I G = E e ς t 1 = .
With respect to (23),
E log S t I G = t E ς 2 = t b 2 ( a 1 ) I a > 1 I a 1 .
It follows from (24) that
V log S t I G = E ς B t ς t 2 + t E ς 2 2 = E ς B t t 2 ς E ς 2 = = t E ς + t 2 4 V ς = t b a 1 I a > 1 + I a 1 + + b 2 t 2 4 ( a 1 ) 2 ( a 2 ) I a > 2 + I a 2 = = t b a 1 1 + b t 4 ( a 1 ) ( a 2 ) I a > 2 + I a 2 .
We have that
E S t I G = E S t G = E S t B S = 1 .
If we set a = 3 and b = 2 , then it results from (28), (31) and (38) that
E log S t I G = E log S t G = E log S t B S = t 2 .
Moreover, then
V log S t I G = V log S t G = t 1 + t 4 .
Set t 0 = T 0 = 0 , T = 1 , a = 3 , b = 2 and k = log K . Then in accordance with (26)
C I G = 4 ( I 3 , 2 , 0 , k , 1 2 K I 3 , 2 , 0 , k , 1 2 ) .
Furthermore, we have immediately from (10) that
I 3 , 2 , 0 , log K , 1 2 = | s | 5 2 e s ( 1 + q ) 3 8 2 π [ B ( 3 , 1 ) ( | s | K 7 2 | s | + s K 5 2 | s | ) × ×   Φ 3 , 2 , 4 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( 4 , 1 ) × ×   K 5 2 | s | Φ 4 , 2 , 5 ; 1 + q 2 , s ( 1 + q ) ]
with
q = c 2 c 2 2 + 4 = k k 2 + 4 and s = c 1 c 2 2 + 4 = k 2 + 4 2
and
I 3 , 2 , 0 , log K , 1 2 = | s | 5 2 e s ( 1 + q ) 3 8 2 π [ B ( 3 , 1 ) ( | s | K 7 2 | s | + s K 5 2 | s | ) × ×   Φ 3 , 2 , 4 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( 4 , 1 ) × ×   K 5 2 | s | Φ 4 , 2 , 5 ; 1 + q 2 , s ( 1 + q ) ]
with
q = u 2 c 2 2 + 4 = k k 2 + 4 and s = c 3 c 2 2 + 4 = k 2 + 4 2 .
It results from (40) that
I 3 , 2 , 0 , k , 1 2 = e k 2 + 4 2 k 2 + 4 k 3 k 2 + 4 1 4 128 π × × [ 1 3 { K 7 2 1 2 k 2 + 4 + K 5 2 1 2 k 2 + 4 } × × Φ 3 , 2 , 4 ; k 2 + 4 k 2 k 2 + 4 , k k 2 + 4 2 k 2 + 4 k 4 k 2 + 4 × × K 5 2 1 2 k 2 + 4 Φ 4 , 2 , 5 ; k 2 + 4 k 2 k 2 + 4 , k k 2 + 4 2 ]
and from (41) that
I 3 , 2 , 0 , k , 1 2 = e k 2 + 4 2 k 2 + 4 k 3 k 2 + 4 1 4 128 π × × [ 1 3 { K 7 2 1 2 k 2 + 4 K 5 2 1 2 k 2 + 4 } × × Φ 3 , 2 , 4 ; k 2 + 4 k 2 k 2 + 4 , k 2 + 4 k 2 + k 2 + 4 k 4 k 2 + 4 × × K 5 2 1 2 k 2 + 4 Φ 4 , 2 , 5 ; k 2 + 4 k 2 k 2 + 4 , k 2 + 4 k 2 ] .
Let
J 2 ( A , B ) = 0 1 x 2 ( 1 A x ) 2 e B x d x
and
J 3 ( A , B ) = 0 1 x 3 ( 1 A x ) 2 e B x d x .
Then we have with respect to (35) that
Φ 3 , 2 , 4 ; A , B = 3 J 2 ( A , B )
and
Φ 4 , 2 , 5 ; A , B = 4 J 3 ( A , B ) .
Therefore,
I 3 , 2 , 0 , k , 1 2 = e k 2 + 4 2 k 2 + 4 k 3 k 2 + 4 1 4 128 π × × [ { K 7 2 1 2 k 2 + 4 + K 5 2 1 2 k 2 + 4 } × × J 2 k 2 + 4 k 2 k 2 + 4 , k k 2 + 4 2 k 2 + 4 k k 2 + 4 × × K 5 2 1 2 k 2 + 4 J 3 k 2 + 4 k 2 k 2 + 4 , k k 2 + 4 2 ]
and
I 3 , 2 , 0 , k , 1 2 = e k 2 + 4 2 k 2 + 4 k 3 k 2 + 4 1 4 128 π × × [ { K 7 2 1 2 k 2 + 4 K 5 2 1 2 k 2 + 4 } × × J 2 k 2 + 4 k 2 k 2 + 4 , k 2 + 4 k 2 + k 2 + 4 k k 2 + 4 × × K 5 2 1 2 k 2 + 4 J 3 k 2 + 4 k 2 k 2 + 4 , k 2 + 4 k 2 ] .

4.4. Comparison of the Prices

The Black-Scholes price is calculated according to (30). The gamma volatility price is computed by (32) taking into account (36) and (37). The inverse-gamma volatility price is determined with respect to (39) using (42) and (43).
Table 1 shows the dynamics of the standard call option prices in dependence on the increase in the strike coefficient K from 0.05 to 10. The numbers around K = 1 in Table 1 relate to the skewness of the probability densities in the gamma and the inverse-gamma models. Since the drift coefficients are negative, the probability density of the logarithm of stock price in the gamma model has a larger weight of extremal events than the probability density of the logarithm of stock price in the inverse-gamma one. This fact explains the inequality C G > C I G for large strike coefficients.
One may notice that the simplest call-put parity identity
S 1 K = ( S 1 K ) I S 1 > K ( K S 1 ) I S 1 K
allows us to obtain the standard European put option prices in our examples directly from Table 1. Indeed, the put option price
P = E ( K S 1 ) + = C + K 1
from (44).

5. Conclusions

We have discussed the generalized Black-Scholes-Merton model, in which the stock volatility is modeled using the gamma and the inverse-gamma distributions. The idea for this research is produced by the papers on the statistical analysis of stock market data by many authors. We refer, in particular, to Seneta (2004), Daal and Madan (2005), Nakakita and Nakatsuma (2021), Mozumder et al. (2015), Luciano and Schoutens (2016) on the gamma distribution, and to Nakajima and Omori (2012), Aas and Haff (2006), Finlay and Seneta (2008), and Takahashi et al. (2021) on the inverse-gamma one in the stochastic volatility modeling.
We have selected for the numerical examples the parameters of the models so that the characteristics of the underlying processes are as close to each other as possible. It is clear from the examples that it can be expected that for large strike coefficients the call option prices should be vastly different for the all three models. At the same time, the difference between the put option prices does not exceed 20%. It should be also noticed that the calculation time of our formulas for the gamma and the inverse-gamma models is comparable with the time of the Black-Scholes formula computation.
The obtained results extend the results of Madan et al. (1998) and Ano and Ivanov (2016), which are derived for the variance-gamma model. Furthermore, the price of the forward-start call option is computed in a model of the skew Student’s t-type. The idea of research is confirmed by the variety of works that have approved the use of the gamma and the inverse-gamma distributions for financial modeling. Numerical examples have shown that the results for the standard call option price may differ substantially from the Black-Scholes formula.
Future studies can relate to the modeling and computation of more complicated derivatives, including American-style ones. Moreover, the theoretical developments of the studied models can also be processed for the potential of obtaining closed-form results in option pricing.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

I would like to thank the anonymous referees whose numerous remarks have improved the paper substantially.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A

Proof of Theorem 1. 
Set
g ( z T ) = E R T 1 ( S T K S T 0 ) + | z T .
Then
g ( z T ) = g ( γ ) = E R T 1 ( S T K S T 0 ) + | γ = = E S T 0 R T 0 R T 0 R T S T S T 0 K + | γ .
Since the random variable S T 0 R T 0 1 = S T 0 ( γ ) R T 0 1 ( γ ) is independent with the random variables R T 0 R T 1 = R T 0 ( γ ) R T 1 ( γ ) and S T S T 0 1 = S T ( γ ) S T 0 1 ( γ ) for any particular realization of γ , we get from (A1) that
g ( z T ) = E S T 0 R T 0 | γ E R T 0 R T S T S T 0 K + | γ = = S 0 g 1 ( γ ) g 2 ( γ )
with
g 1 ( γ ) = E R T 0 S T R T S T 0 I S T S T 0 > K | γ
and
g 2 ( γ ) = K E R T 0 R T I S T S T 0 > K | γ .
Because of
C G = E g ( z T ) ,
we get from (A2) that
C G = S 0 E g 1 ( γ ) E g 2 ( γ ) .
Next, one may observe that
g 1 ( γ ) = E exp γ T 0 T σ ˜ s d B s γ 2 T 0 T σ ˜ s 2 d s I γ T 0 T σ ˜ s d B s > ξ 1 ( γ ) | γ .
with
ξ 1 ( γ ) = log K T 0 T r ˜ s d s + γ T 0 T σ ˜ s 2 2 ρ ˜ s d s .
Let
σ ¯ ( γ ) = γ T 0 T σ ˜ s 2 d s .
Since
L a w γ T 0 T σ ˜ s d B s | γ = L a w N 0 , σ ¯ 2 ( γ ) | γ ,
where N is the normally distributed random variable, we get that
g 1 ( γ ) = exp σ ¯ 2 ( γ ) 2 E exp N 0 , σ ¯ 2 ( γ ) I N 0 , σ ¯ 2 ( γ ) > ξ 1 ( γ ) | γ = = exp σ ¯ 2 ( γ ) 2 ξ 1 ( γ ) 1 σ ¯ ( γ ) 2 π exp x x 2 2 σ ¯ 2 ( γ ) d x = = ξ 1 ( γ ) 1 σ ¯ ( γ ) 2 π exp x σ ¯ 2 ( γ ) 2 2 σ ¯ 2 ( γ ) d x = = N σ ¯ ( γ ) ξ 1 ( γ ) σ ¯ ( γ ) .
Furthermore,
g 2 ( γ ) = = K exp T 0 T r ˜ s + γ ρ ˜ s d s ξ 1 ( γ ) 1 σ ¯ ( γ ) 2 π exp x 2 2 σ ¯ ( γ ) 2 d x = = K exp T 0 T r ˜ s + γ ρ ˜ s d s N ξ 1 ( γ ) σ ¯ ( γ ) .
One may notice that
σ ¯ ( γ ) ξ 1 ( γ ) σ ¯ ( γ ) = γ T 0 T σ ˜ s 2 d s log K T 0 T r ˜ s d s + γ T 0 T σ ˜ s 2 2 ρ ˜ s d s γ T 0 T σ ˜ s 2 d s = = c 1 γ + c 2 γ
with
c 1 = T 0 T σ ˜ s 2 2 + ρ ˜ s d s T 0 T σ ˜ s 2 d s , c 2 = T 0 T r ˜ s d s log K T 0 T σ ˜ s 2 d s
and
ξ 1 ( γ ) σ ¯ ( γ ) = c 3 γ + c 2 γ ,
where
c 3 = T 0 T ρ ˜ s σ ˜ s 2 2 d s T 0 T σ ˜ s 2 d s .
Hence
g 1 ( γ ) = N c 1 γ + c 2 γ
and
g 2 ( γ ) = K exp T 0 T r ˜ s + γ ρ ˜ s d s N c 3 γ + c 2 γ .
Set
c 4 = T 0 T ρ ˜ s d s .
Then we get from (A3)–(A5) that
C G = S 0 ( E N c 1 γ + c 2 γ K exp T 0 T r ˜ s d s E e c 4 γ N c 3 γ + c 2 γ ) .
One may see from (A6) that we need to calculate the integral
I G = I G ( v 1 , v 2 , v 3 ) = Γ ( a ) b a 0 e v 1 x N v 2 x + v 3 x f γ ( x ) d x
for v 1 0 and v 2 , v 3 R . Then
C G = S 0 b a Γ ( a ) ( I G ( 0 , c 1 , c 2 ) K exp T 0 T r ˜ s d s I G ( c 4 , c 3 , c 2 ) ) .
We have that
I G = 0 x a 1 e ( b + v 1 ) x N v 2 x + v 3 x d x .
This integral is computed in point of the cases below.
Case 1. v 3 = 0 . Then we have with respect to Case 2.2 at p. 208 of Ano and Ivanov (2016) that
I G ( v 1 , v 2 , v 3 ) = = Γ a + 1 2 2 π ( b + v 1 ) a B 1 2 , a 2 + v 2 b + v 1 F a + 1 2 , 1 2 , 3 2 ; v 2 2 2 ( b + v 1 ) .
Case 2. v 3 0 . We get in accordance with (21) of Ano and Ivanov (2016) that
I G ( v 1 , v 2 , v 3 ) = | s | a 1 2 e s ( 1 + q ) a 2 π ( b + v 1 ) a [ B ( a , 1 ) ( | s | K a + 1 2 | s | + + s K a 1 2 | s | ) Φ a , 1 a , a + 1 ; 1 + q 2 , s ( 1 + q ) ( 1 + q ) s B ( a + 1 , 1 ) K a 1 2 | s | Φ a + 1 , 1 a , a + 2 ; 1 + q 2 , s ( 1 + q ) ] ,
where
q = v 2 v 2 2 + 2 ( b + v 1 ) and s = v 3 v 2 2 + 2 ( b + v 1 ) .
Combining together (A7)–(A9) and using the auxiliary function (10), we establish that
C G = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 1 , c 2 ) K exp T 0 T r ˜ s d s I ( a , b , c 4 , c 3 , c 2 ) ) .
Proof of Theorem 2. 
Similarly to (A6), we get that in this case
C I G = S 0 ( E N c 5 ς + c 2 ς K exp T 0 T r ˜ s d s E N c 5 ς + c 2 ς )
with
c 5 = 1 2 T 0 T σ ˜ s 2 d s and c 2 = T 0 T r ˜ s d s log K T 0 T σ ˜ s 2 d s .
Set
I I G = I I G ( v 1 , v 2 ) = Γ ( a ) b a 0 N v 1 x + v 2 x f ς ( x ) d x
for v 1 0 and v 2 R . Then we have from (A10) that
C I G = S 0 b a Γ ( a ) ( I I G ( c 5 , c 2 ) K exp T 0 T r ˜ s d s I I G ( c 5 , c 2 ) ) .
One may observe that
I I G = 0 x a 1 e b x N v 1 x + v 2 x d x .
Let us define for t v 2 functions v ˜ 2 ( t ) , v ˜ 1 ( t ) and u ( t ) as
v ˜ 2 ( t ) = t , v ˜ 1 ( t ) = v 1 v 2 2 + 2 b t 2 + 2 b
and
u ( t ) = v ˜ 1 ( t ) x + v ˜ 2 ( t ) x .
Then
N v 1 x + v 2 x = v 1 x + v 2 x N u ( u ) d u = v 2 N u ( u ) u t d t
and we have that
I I G = 1 2 π 0 x a 1 e b x ( v 2 e v ˜ 1 ( t ) x + v ˜ 2 ( t ) 2 2 x ( x 1 2 x 1 2 v 1 v 2 2 + 2 b ( t 2 + 2 b ) 3 2 t ) d t ) d x = 1 2 π ( I I G 1 I I G 2 ) ,
where
I I G 1 = v 2 e v ˜ 1 ( t ) v ˜ 2 ( t ) ( 0 x a 3 2 e v ˜ 1 2 ( t ) x 2 v ˜ 2 2 ( t ) + 2 b 2 x d x ) d t
and
I I G 2 = s v 2 e v ˜ 1 ( t ) v ˜ 2 ( t ) t ( t 2 + 2 b ) 3 2 ( 0 x a 1 2 e v ˜ 1 2 ( t ) x 2 v ˜ 2 2 ( t ) + 2 b 2 x d x ) d t
with
s = v 1 v 2 2 + 2 b .
Formula 3.471.9 of Gradshteyn and Ryzhik (2007) comprises the identity
0 x u 1 1 e u 2 x u 3 x d x = 2 u 2 u 3 u 1 2 K u 1 2 u 2 u 3 ,
where u 1 R , u 2 > 0 and u 3 > 0 . It results immediately from (A13) that
0 x a 1 2 j e v ˜ 1 2 ( t ) x 2 v ˜ 2 2 ( t ) + 2 b 2 x d x = 2 ( v ˜ 2 2 ( t ) + 2 b v ˜ 1 2 ( t ) ) 1 4 a + j 2 × × K 1 2 a j ( | v ˜ 1 ( t ) | v ˜ 2 2 ( t ) + 2 b ) = 2 ( t 2 + 2 b | s | ) 1 2 a j K 1 2 a j ( | s | )
for j = 0 , 1 . Hence we get that
I I G 1 = 2 | s | a + 1 2 K a + 1 2 ( | s | ) v 2 ( t 2 + 2 b ) a 1 2 e s t t 2 + 2 b d t
and
I I G 2 = 2 s | s | a 1 2 K a 1 2 ( | s | ) v 2 t ( t 2 + 2 b ) a 1 e s t t 2 + 2 b d t .
Set
w = t t 2 + 2 b .
Then
t = w 2 b 1 w 2 , t 2 + 2 b = 2 b 1 w 2 , t w = 2 b ( 1 w 2 ) 3 2
and one may observe that
I I G 1 = | s | a + 1 2 K a + 1 2 ( | s | ) 2 a 1 b a 1 v 2 v 2 2 + 2 b 1 w 2 a 1 e s w d w
and
I I G 2 = s | s | a 1 2 K a 1 2 ( | s | ) 2 a 1 b a 1 v 2 v 2 2 + 2 b w 1 w 2 a 1 e s w d w .
Let
q = v 2 v 2 2 + 2 b
and
y = 1 + w 1 + q .
Then
w = ( 1 + q ) y 1 , w y = 1 + q
and hence
I I G 1 = ( 1 + q ) | s | a + 1 2 e s K a + 1 2 ( | s | ) 2 a 1 b ϰ 0 1 1 ( 1 + q ) y 1 2 a 1 e s ( 1 + q ) y d y = = ( 1 + q ) a | s | a a + 1 2 e s K a + 1 2 ( | s | ) b ϰ 0 1 y a 1 1 1 + q 2 y a 1 e s ( 1 + q ) y d y
and
I I G 2 = ( 1 + q ) s | s | a 1 2 e s K a 1 2 ( | s | ) 2 a 1 b a × × 0 1 ( 1 + q ) y 1 1 ( 1 + q ) y 1 2 a 1 e s ( 1 + q ) y d y = = ( 1 + q ) a s | s | a 1 2 e c K a 1 2 ( | s | ) b a ( ( 1 + q ) 0 1 y a 1 1 + q 2 y a 1 × × e s ( 1 + q ) y d y 0 1 y a 1 1 1 + q 2 y a 1 e s ( 1 + q ) y d y ) .
Furthermore, we apply (35) to (A14) and (A15) and infer that
I I G 1 = = ( 1 + q ) a | s | a + 1 2 e s K a + 1 2 ( | s | ) b a B a , 1 Φ a , 1 a , a + 1 ; 1 + q 2 , s ( 1 + q )
and
I I G 2 = ( 1 + q ) a s | s | a 1 2 e s K a 1 2 ( | s | ) b a ( ( 1 + q ) B a + 1 , 1 × ×   Φ a + 1 , 1 a , a + 2 ; 1 + q 2 , s ( 1 + q ) B a , 1 × ×   Φ a , 1 a , a + 1 ; 1 + q 2 , s ( 1 + q ) ) .
It follows from (A12), (A16) and (A17) that
I I G = ( 1 + q ) a | s | a 1 2 e s b a 2 π [ B a , 1 ( | s | K a + 1 2 ( | s | ) + s K a 1 2 ( | s | ) ) × ×   Φ a , 1 a , a + 1 ; 1 + q 2 , s ( 1 + q ) s ( 1 + q ) B a + 1 , 1 × ×   K a 1 2 ( | s | ) Φ a + 1 , 1 a , a + 2 ; 1 + q 2 , s ( 1 + q ) ] .
Finally, it results from (A11), (A18) and (10) that
C I G = S 0 b a Γ ( a ) ( I ( a , b , 0 , c 2 , c 5 ) K exp T 0 T r ˜ s d s I ( a , b , 0 , c 2 , c 5 ) ) .

References

  1. Aas, Kjersti, and Ingrid H. Haff. 2006. The generalized hyperbolic skew Student’s t-distribution. Journal of Financial Econometrics 4: 275–309. [Google Scholar] [CrossRef] [Green Version]
  2. Alghalith, Moawia, Christos Floros, and Konstantonos Gkillas. 2020. Estimating stochastic volatility under the assumption of stochastic volatility of volatility. Risks 8: 35. [Google Scholar] [CrossRef] [Green Version]
  3. Andersen, Torben G., Tim Bollerslev, Francis X. Diebold, and Paul Labys. 2001. The distribution of realized exchange rate volatility. Journal of the American Statistical Association 96: 42–55. [Google Scholar] [CrossRef]
  4. Andersen, Torben G., Tim Bollerslev, and Dobrislav Dobrev. 2007. No-arbitrage semi-martingale restrictions for continuous-time volatility models subject to leverage effects, jumps and i.i.d. noise: Theory and testable distribution implications. Journal of Econometrics 138: 125–80. [Google Scholar] [CrossRef] [Green Version]
  5. Ano, Katsunori, and Roman V. Ivanov. 2016. On exact pricing of FX options in multivariate time-changed Lévy models. Review Derivatives Research 19: 201–16. [Google Scholar]
  6. Bateman, Harry, and Arthur Erdélyi. 1953. Higher Transcendental Functions. New York: McGraw-Hill, pp. 222–44. [Google Scholar]
  7. Bates, David S. 1996. Jumps and stochastic volatility: Exchange rate processes implicity in Deutsche mark options. The Review of Financial Studies 9: 69–107. [Google Scholar] [CrossRef]
  8. Bibby, Bo Martin, and Michael Sørensen. 2003. Hyperbolic processes in finance. In Handbook of Heavy Tailed Distributions in Finance. Edited by Svetlozar Todorov Rachev. Amsterdam: Elsevier, pp. 211–48. [Google Scholar]
  9. Black, Fischer, and Myron Scholes. 1973. The pricing of options and corporate liabilities. Journal of Political Economy 81: 637–54. [Google Scholar] [CrossRef] [Green Version]
  10. Brigo, Damiano, and Fabio Mercurio. 2002. Lognormal-mixture dynamics and calibration to market volatility smiles. International Journal of Theoretical and Applied Finance 5: 427–46. [Google Scholar] [CrossRef] [Green Version]
  11. Brigo, Damiano, Fabio Mercurio, and Giulio Sartorelli. 2003. Alternative asset-price dynamics and volatility smile. Quantitative Finance 3: 173–83. [Google Scholar] [CrossRef] [Green Version]
  12. Cao, Jiling, Guanghua Lian, and Teh R. N. Roslan. 2016. Pricing variance swaps under stochastic volatility and stochastic interest rate. Applied Mathematics and Computation 277: 72–81. [Google Scholar] [CrossRef]
  13. Chib, Siddhartha, Federico Nardari, and Neil Shephard. 2002. Markov chain Monte Carlo methods for stochastic volatility models. Journal of Econometrics 108: 281–316. [Google Scholar] [CrossRef]
  14. Cuchiero, Christa, Wahid Khosrawi, and Josef Teichmann. 2020. A generative adversarial network approach to calibration of local stochastic volatility models. Risks 8: 101. [Google Scholar] [CrossRef]
  15. Daal, Elton A., and Dilip B. Madan. 2005. An empirical examination of the variance-gamma model for foreign currency options. Journal of Business 78: 2121–52. [Google Scholar] [CrossRef]
  16. Derman, Emanuel, and Iraj Kani. 1994. Riding on a smile. Risk Magazine 2: 32–39. [Google Scholar]
  17. Dupire, Bruno. 1994. Pricing with a smile. Risk Magazine 1: 18–20. [Google Scholar]
  18. Eberlein, Ernst, Antonis Papapantoleon, and Albert N. Shiryaev. 2008. On the duality principle in option pricing: Semimartingale setting. Finance and Stochastics 12: 265–92. [Google Scholar] [CrossRef]
  19. Eberlein, Ernst, Antonis Papapantoleon, and Albert N. Shiryaev. 2009. Esscher transform and the duality principle for multidimensional semimartingales. Annals of Applied Probability 19: 1944–71. [Google Scholar] [CrossRef]
  20. Eraker, Bjorn, Michael Johannes, and Nicholas Polson. 2003. The impact of jumps in returns and volatility. Journal of Finance 53: 1269–300. [Google Scholar] [CrossRef]
  21. Erdélyi, Arthur, Wilhelm Magnus, Fritz Oberhettinger, and Francesco G. Tricomi. 1954. Tables of Integral Transforms. New York: McGraw-Hill, pp. 129–223. [Google Scholar]
  22. Finlay, Richard, and Eugene Seneta. 2006. Stationary-increment student and variance-gamma processes. Journal of Applied Probability 43: 441–53. [Google Scholar] [CrossRef]
  23. Finlay, Richard, and Eugene Seneta. 2008. Stationary-increment variance-gamma and t Models: Simulation and parameter estimation. International Statistical Review 76: 167–86. [Google Scholar] [CrossRef]
  24. Fung, Thomas, and Eugene Seneta. 2010. Modelling and estimation for bivariate financial returns. International Statistical Review 78: 117–33. [Google Scholar] [CrossRef]
  25. Göncü, Ahmet, Mehmet O. Karahan, and Tolga U. Kuzubas. 2016. A comparative goodness-of-fit analysis of distributions of some Lévy processes and Heston model to stock index returns. North American Journal of Economics and Finance 36: 69–83. [Google Scholar] [CrossRef] [Green Version]
  26. Gradshteyn, Israil S., and Iosif M. Ryzhik. 2007. Table of Integrals, Series and Products, 7th ed. London: Academic Press, pp. 346–71. [Google Scholar]
  27. Grzelak, Lech A., and Cornelius W. Oosterlee. 2011. On the Heston model with stochastic interest rates. SIAM Journal on Financial Mathematics 2: 255–86. [Google Scholar] [CrossRef] [Green Version]
  28. Herdegen, Martin, and Martin Schweizer. 2018. Semi-efficient valuations and put-call parity. Mathematical Finance 28: 1061–106. [Google Scholar] [CrossRef] [Green Version]
  29. Hong, Seok Y., Ingmar Nolte, Stephen J. Taylor, and Xiaolu Zhao. 2023. Volatility estimation and forecasts based on price durations. Journal of Financial Econometrics 21: 106–44. [Google Scholar] [CrossRef]
  30. Ivanov, Roman V. 2018. A credit-risk valuation under the variance-gamma asset return. Risks 6: 58. [Google Scholar] [CrossRef] [Green Version]
  31. Ivanov, Roman V. 2022. The risk measurement under the variance-gamma process with drift switching. Journal of Risk and Financial Management 15: 22. [Google Scholar] [CrossRef]
  32. Jacod, Jean, and Albert N. Shiryaev. 1987. Limit Theorems for Stochastic Processes. Berlin: Springer, pp. 249–69. [Google Scholar]
  33. Kallsen, Jasper, and Albert N. Shiryaev. 2002. The cumulant process and Esscher’s change of measure. Finance and Stochastics 6: 397–428. [Google Scholar] [CrossRef] [Green Version]
  34. Kim, Sangjoon, Neil Shephard, and Siddhartha Chib. 1998. Stochastic volatility: Likelihood inference and comparison with ARCH Models. Review of Economic Studies 65: 361–93. [Google Scholar] [CrossRef]
  35. Levendis, Alexis, and Eben Maré. 2022. Efficient pricing of spread options with stochastic rates and stochastic volatility. Journal of Risk and Financial Management 15: 504. [Google Scholar] [CrossRef]
  36. Linders, Daniel, and Ben Stassen. 2016. The multivariate variance gamma model: Basket option pricing and calibration. Quantitative Finance 16: 555–72. [Google Scholar] [CrossRef]
  37. Liu, Jia. 2021. A Bayesian semiparametric realized stochastic volatility model. Journal of Risk and Financial Management 14: 617. [Google Scholar] [CrossRef]
  38. Liu, Shuaiqiang, Cornelius W. Oosterlee, and Sander M. Bohte. 2019. Pricing options and computing implied volatilities using neural networks. Risks 7: 16. [Google Scholar] [CrossRef] [Green Version]
  39. Liu, Tao, Malik Z. Ullah, Stanford Shateyi, Chao Liu, and Yanxiong Yang. 2023. An efficient localized RBF-FD method to simulate the Heston-Hull-White PDE in finance. Mathematics 11: 833. [Google Scholar] [CrossRef]
  40. Luciano, Elisa, and Wim Schoutens. 2016. A multivariate jump-driven financial asset model. Quantitative Finance 6: 385–402. [Google Scholar] [CrossRef] [Green Version]
  41. Madan, Dilip B., Peter Carr, and Eric C. Chang. 1998. The variance gamma process and option pricing. Review of Finance 2: 79–105. [Google Scholar] [CrossRef] [Green Version]
  42. Mahieu, Ronald J., and Peter C. Schotman. 1998. An empirical application of stochastic volatility models. Journal of Applied Econometrics 13: 333–60. [Google Scholar] [CrossRef]
  43. Mao, Chen, Guanqi Liu, and Yuwen Wang. 2022. A closed-form pricing formula for log-return variance swaps under stochastic volatility and stochastic interest Rate. Mathematics 10: 5. [Google Scholar] [CrossRef]
  44. McNeil, Alexander J., Rüdiger Frey, and Paul Embrechts. 2005. Quantitative Risk Management. Princeton: Princeton University Press, p. 75. [Google Scholar]
  45. Men, Zhongxian, Tony S. Wirjanto, and Adam W. Kolkiewicz. 2021. Multiscale stochastic volatility model with heavy tails and leverage effects. Journal of Risk and Financial Management 14: 225. [Google Scholar] [CrossRef]
  46. Merton, Robert C. 1973. Theory of rational option pricing. Bell Journal of Economics and Management Science 4: 141–83. [Google Scholar] [CrossRef] [Green Version]
  47. Moosbrucker, Thomas. 2006. Explaining the correlation smile using variance gamma distributions. Journal of Fixed Income 16: 71–87. [Google Scholar] [CrossRef]
  48. Mozumder, Sharif, Ghulam Sorwar, and Kevin Dowd. 2015. Revisiting variance gamma pricing: An application to s&p500 index options. International Journal of Financial Engineering 2: 1550022. [Google Scholar]
  49. Musiela, Marek, and Marek Rutkowski. 2005. Martingale Methods in Financial Modelling, 2nd ed. Berlin: Springer, pp. 195–96, 234–36. [Google Scholar]
  50. Nakajima, Jouchi. 2020. Skew selection for factor stochastic volatility models. Journal of Applied Statisitics 47: 582–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Nakajima, Jouchi, and Yasuhiro Omori. 2012. Stochastic volatility model with leverage and asymmetrically heavy-tailed error using GH skew Student’s t-distribution. Computational Statistics and Data Analysis 56: 3690–704. [Google Scholar] [CrossRef]
  52. Nakakita, Makoto, and Teruo Nakatsuma. 2021. Bayesian analysis of intraday stochastic volatility models of high-frequency stock returns with skew heavy-tailed Errors. Journal of Risk and Financial Management 14: 145. [Google Scholar] [CrossRef]
  53. Nzokem, Aubain H. 2023. Pricing European options under stochastic volatility models: Case of five-parameter variance-gamma process. Journal of Risk and Financial Management 16: 55. [Google Scholar] [CrossRef]
  54. Oksendal, Bernt. 2003. Stochastic Differential Equations, 5th ed. New York: Springer, pp. 249–87. [Google Scholar]
  55. Rathgeber, Andreas, Johannes Stadler, and Stefan Stöckl. 2016. Modeling share returns—An empirical study on the variance gamma model. Journal of Economics and Finance 40: 653–82. [Google Scholar] [CrossRef]
  56. Schoutens, Wim. 2003. Lévy Processes in Finance. Chichester: John Wiley & Sons, pp. 52–53. [Google Scholar]
  57. Schweizer, Martin. 2002. On Bermudan options. In Advances in Finance and Stochastics: Essays in Honour of Dieter Sondermann. Edited by Klaus Sandmann and Philip J. Schönbucher. Berlin: Springer, pp. 257–69. [Google Scholar]
  58. Seneta, Euegene. 2004. Fitting the variance-gamma model to financial data. In Stochastic Methods and Their Applications. Edited by Joe Gani and Euegene Seneta. Sheffield: Applied Probability Trust, vol. 41, pp. 177–87. [Google Scholar]
  59. Shackleton, Mark B., Stephen J. Taylor, and Peng Yu. 2010. A multi-horizon comparison of density forecasts for the S& P 500 using index returns and option prices. Journal of Banking and Finance 34: 2678–93. [Google Scholar]
  60. Shiryaev, Albert N. 1999. Essentials of Stochastic Finance. Singapore: World Scientific, pp. 739–48. [Google Scholar]
  61. Srivastava, Hari M., and Per W. Karlsson. 1985. Multiple Gaussian Hypergeometric Series. New York: Wiley, pp. 15–33. [Google Scholar]
  62. Takahashi, Makoto, Toshiaki Watanabe, and Yasuhori Omori. 2021. Forecasting daily volatility of stock price index using daily returns and realized volatility. Econometrics and Statistics, in press. [Google Scholar] [CrossRef]
  63. Taylor, Stephen J., Joseph Tzeng, and Martin Widdicks. 2018. Information about price and volatility jumps inferred from options prices. Journal of Futures Markets 38: 1206–26. [Google Scholar] [CrossRef]
  64. Wallmeier, Martin, and Martin Diethelm. 2012. Multivariate downside risk: Normal versus variance gamma. Journal of Futures Markets 32: 431–58. [Google Scholar] [CrossRef] [Green Version]
  65. Witkovský, Viktor. 2001. Computing the distribution of a linear combination of inverted gamma variables. Kybernetika 37: 79–90. [Google Scholar]
Table 1. The comparison of the standard European call option prices.
Table 1. The comparison of the standard European call option prices.
C K 0.05 0.1 0.5 0.75 1 1.25 1.5 210
C B S 0.950 0.901 0.595 0.473 0.383 0.316 0.264 0.191 0.010
C G 0.951 0.903 0.583 0.438 0.334 0.267 0.222 0.167 0.033
C I G 0.951 0.902 0.586 0.452 0.356 0.288 0.237 0.171 0.025
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ivanov, R.V. On the Stochastic Volatility in the Generalized Black-Scholes-Merton Model. Risks 2023, 11, 111. https://doi.org/10.3390/risks11060111

AMA Style

Ivanov RV. On the Stochastic Volatility in the Generalized Black-Scholes-Merton Model. Risks. 2023; 11(6):111. https://doi.org/10.3390/risks11060111

Chicago/Turabian Style

Ivanov, Roman V. 2023. "On the Stochastic Volatility in the Generalized Black-Scholes-Merton Model" Risks 11, no. 6: 111. https://doi.org/10.3390/risks11060111

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop