Next Article in Journal
From Traditional-Ritual Activities to Financial Report: Integrating Local Wisdom in Bantengan Financial Bookkeeping
Previous Article in Journal
Evaluation of the Resilience of Real Estate and Property Stocks to Inflation and Interest Rate Uncertainty: Implementation of Two Asset Pricing Models
Previous Article in Special Issue
A Double Optimum New Solution Method Based on EVA and Knapsack
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fitting the Seven-Parameter Generalized Tempered Stable Distribution to Financial Data

1
Department of Mathematics & Statistics, York University, Toronto, ON M3J 1P3, Canada
2
Department of Statistics and Operations Research, University of Limpopo, Sovenga 0727, South Africa
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
J. Risk Financial Manag. 2024, 17(12), 531; https://doi.org/10.3390/jrfm17120531
Submission received: 15 October 2024 / Revised: 16 November 2024 / Accepted: 18 November 2024 / Published: 22 November 2024
(This article belongs to the Special Issue Featured Papers in Mathematics and Finance)

Abstract

:
This paper proposes and implements a methodology to fit a seven-parameter Generalized Tempered Stable (GTS) distribution to financial data. The nonexistence of the mathematical expression of the GTS probability density function makes maximum-likelihood estimation (MLE) inadequate for providing parameter estimations. Based on the function characteristic and the fractional Fourier transform (FRFT), we provide a comprehensive approach to circumvent the problem and yield a good parameter estimation of the GTS probability. The methodology was applied to fit two heavy-tailed data (Bitcoin and Ethereum returns) and two peaked data (S&P 500 and SPY ETF returns). For each historical data, the estimation results show that six-parameter estimations are statistically significant except for the local parameter, μ . The goodness of fit was assessed through Kolmogorov–Smirnov, Anderson–Darling, and Pearson’s chi-squared statistics. While the two-parameter geometric Brownian motion (GBM) hypothesis is always rejected, the GTS distribution fits significantly with a very high p-value and outperforms the Kobol, Carr–Geman–Madan–Yor, and bilateral Gamma distributions.

1. Introduction

Modeling high-frequency asset return with the normal distribution is the underlying assumption in many financial tools, such as the Black–Scholes–Merton option pricing model and the risk metric variance–covariance technique for value at risk (VAR). However, substantial empirical evidence rejects the normal distribution for various asset classes and financial markets. The symmetric and rapidly decreasing tail properties of the normal distribution cannot describe the skewed and fat-tailed properties of the asset return distribution.
The α -stable distribution has been proposed (Nolan 2020; Sato 1999) as an alternative to the normal distribution for modeling asset return and many types of physical and economic systems. The theoretical and empirical argument is that the stable distribution generalizes the Central Limit Theorem regardless of the variance nature (finite or infinite) (Nzokem 2024; Rachev et al. 2011). There are two major drawbacks (Borak et al. 2005; Nolan 2020): firstly, the lack of closed formulas for densities and distribution functions, except for the normal distribution ( α = 2 ), Cauchy distribution ( α = 1 ), and Lévy distribution ( α = 1 2 ) (Tsallis 1997); secondly, most of the moments of the stable distribution are infinite. An infinite variance of asset return leads to an infinite price for derivative instruments such as options.
The Generalized Tempered Stable (GTS) distribution was developed to overcome the shortcomings of the two distributions, and the tails of the GTS distribution are heavier than the normal distribution but thinner than the stable distribution (Grabchak and Samorodnitsky 2010; Kim et al. 2009). The general form of the GTS distribution can be defined by the following Lévy measure ( V ( d x ) ) (1):
V ( d x ) = α + e λ + x x 1 + β + 1 x > 0 + α e λ | x | | x | 1 + β 1 x < 0 d x
where 0 β + 1 , 0 β 1 , α + 0 , α 0 , λ + 0 , and λ 0 . More details on the Tempered Stable distribution are provided in (Küchler and Tappe 2013; Rachev et al. 2011).
The rich class of the GTS distribution (1) has a myriad of applications ranging from financial to mathematical physics and economic systems. However, few studies (Fallahgoul and Loeper 2021; Massing 2024; Nzokem and Montshiwa 2022) have covered the methods and techniques to estimate the parameters of the GTS distribution. This study aims to contribute to the literature by providing a methodology for fitting the seven-parameter GTS distribution. As illustrations, the study used four historical prices: two heavy-tailed data (Bitcoin and Ethereum returns) and two peaked data (S&P 500 and SPY ETF returns). The GTS distribution is fitted to the underlying distribution of each data index and goodness-of-fit analysis is carried out. The main disadvantage of the GTS distribution is the lack of the closed forms of the density, cumulative, and derivative functions. We use a computational algorithm, called the enhanced fast FRFT scheme (Nzokem 2023a), to circumvent the problem.
The rest of the paper is organized as follows: Section 2 provides some theoretical framework of the GTS distribution. Section 3 presents the multivariate maximum-likelihood (ML) method and the analytic version of the two-parameter normal distribution. Section 4 presents the results of the GTS parameter estimations, along with the associated statistical tests for the heavy-tailed Bitcoin and Ethereum returns. Section 5 fits the GTS distribution to the traditional indices S&P 500 and SPY ETF returns, while Section 6 presents the results of the goodness-of-fit test. Section 7 provides the concluding remarks.

2. Generalized Tempered Stable (GTS) Distribution

The Lévy measure of the GTS distribution ( V ( d x ) ) is defined in (2) as a product of a tempering function q ( x ) and a Lévy measure of the α -stable distribution V s t a b l e ( d x ) :
q ( x ) = e λ + x 1 x > 0 + e λ | x | 1 x < 0 V s t a b l e ( d x ) = α + 1 x 1 + β + 1 x > 0 + α 1 | x | 1 + β 1 x < 0 d x V ( d x ) = q ( x ) V s t a b l e ( d x ) = α + e λ + x x 1 + β + 1 x > 0 + α e λ | x | | x | 1 + β 1 x < 0 d x
where 0 β + 1 , 0 β 1 , α + 0 , α 0 , λ + 0 and λ 0 .
The six parameters that appear have important interpretations. β + and β are the indexes of stability bounded below by 0 and above by 2 (Borak et al. 2005). They capture the peakedness of the distribution similarly to the β -stable distribution, but the distribution tails are tempered. If β increases (decreases), then the peakedness decreases (increases). α + and α are the scale parameters, also called the process intensity (Boyarchenko and Levendorskii 2002); they determine the arrival rate of jumps for a given size. λ + and λ control the decay rate on the positive and negative tails. Additionally, λ + and λ are also skewness parameters. If λ + > λ ( λ + < λ ), then the distribution is skewed to the left (right), and if λ + = λ , then it is symmetric (Fallahgoul et al. 2019; Rachev et al. 2011). α and λ are related to the degree of peakedness and thickness of the distribution. If α increases (decreases), the peakedness and the thickness decrease (increase). Similarly, if λ increases (decreases), then the peakedness increases (decreases) and the thickness decreases (increases) (Bianchi et al. 2019). For more details on the tempering function and Lévy measure of the tempered stable distribution, refer to (Küchler and Tappe 2013; Rachev et al. 2011).
The activity process of the GTS distribution can be studied from the integral (3) of the Lévy measure (2):
+ V ( d x ) = + if 0 β + < 1 0 β < 1 α + λ + β + Γ ( β + ) + α λ β Γ ( β ) if β + < 0 β < 0 .
As shown in (3), if β + < 0 and β < 0 , G T S ( β + , β , α + , α , λ + , λ ) is of a finite activity process and can be written as a compound Poisson (Barndorff-Nielsen and Shephard 2002). When 0 β + < 1 and 0 β < 1 , this Lévy density ( V ( d x ) ) is not integrable as it goes off to infinity too rapidly as x goes to zero (Barndorff-Nielsen and Shephard 2002), which means in practice that there will be a large number of very small jumps. As shown in (3), G T S ( β + , β , α + , α , λ + , λ ) is an infinite activity process with infinite jumps in any given time interval.
In addition to the infinite activities process, the variation in the process can be studied through the following integral:
1 1 | x | V ( d x ) = 1 0 | x | V ( d x ) + 0 1 | x | V ( d x ) = α λ β 1 γ ( 1 β , λ ) + α + λ + β + 1 γ ( 1 β + , λ + )
where γ ( s , x ) = 0 x y s 1 e y d y is the lower incomplete gamma function.
And we have:
1 1 | x | V ( d x ) < + if 0 < β 1 & 0 < β + 1 .
As shown in (4), G T S ( β + , β , α + , α , λ + , λ ) generates a finite variance process, which is contrary to the Brownian motion process. G T S ( β + , β , α + , α , λ + , λ ) generates a type B Lévy process (Ken-Iti 2001), which is a purely non-Gaussian infinite activity Lévy process of finite variation whose sample paths have an infinite number of small jumps and a finite number of large jumps in any finite time interval.
The GTS distribution can be denoted by X G T S ( β + , β , α + , α , λ + , λ ) and X = X + X with X + 0 , X 0 . X + T S ( β + , α + , λ + ) and X T S ( β , α , λ ) . By adding the location parameter, the GTS distribution becomes G T S ( μ , β + , β , α + , α , λ + , λ ) , and we have (5):
Y = μ + X = μ + X + X , Y G T S ( μ , β + , β , α + , α , λ + , λ ) .

2.1. GTS Distribution and Characteristic Exponent

Theorem 1. 
Consider a variable Y G T S ( μ , β + , β , α + , α , λ + , λ ) . The characteristic exponent can be written as:
Ψ ( ξ ) = μ ξ i + α + Γ ( β + ) ( λ + i ξ ) β + λ + β + + α Γ ( β ) ( λ + i ξ ) β λ β .
Proof. 
V ( d x ) in (2) is a Lévy measure. The following relation is satisfied from (4):
+ M i n ( 1 , | x | ) V ( d x ) < + .
More details on the proof are provided in (Nzokem and Maposa 2024).
The Lévy–Khintchine representation (Barndorff-Nielsen and Shephard 2002) for non-negative Lévy process is applied on Y. Y = μ + X = μ + X + X and we have:
Ψ ( ξ ) = L o g E e i Y ξ = i μ ξ + L o g E e i X + ξ + L o g E e i X ξ = i μ ξ + 0 + e i y ξ 1 α + e λ + y y 1 + β + d y + 0 + e i y ξ 1 α e λ y y 1 + β d y ,
0 + e i y ξ 1 α + e λ + y y 1 + β + d y = α + λ + β + Γ ( β + ) k = 1 + Γ ( k β + ) Γ ( β + ) k ! ( i ξ λ + ) k = α + λ + β + Γ ( β + ) k = 1 + β + k ( i ξ λ + ) k = α + Γ ( β + ) ( λ + i ξ ) β + λ + β + .
Similarly, we have:
0 + e i y ξ 1 α e λ y y 1 + β d y = α Γ ( β ) ( λ + i ξ ) β λ β .
The expression in (7) becomes:
Ψ ( ξ ) = i μ ξ + α + Γ ( β + ) ( λ + i ξ ) β + λ + β + + α Γ ( β ) ( λ + i ξ ) β λ β .
Theorem 2. 
Consider a variable Y G T S ( μ , β + , β , α + , α , λ + , λ ) .
If ( β , β + ) ( 0 , 0 ) , GTS becomes a bilateral Gamma distribution with the following characteristic exponent:
Ψ ( ξ ) = μ ξ i α + log 1 1 λ + i ξ α log 1 + 1 λ i ξ .
In addition to ( β , β + ) ( 0 , 0 ) , if α = α + = α , GTS becomes a Variance-Gamma (VG) distribution with parameter ( μ , δ , σ , α , θ )
δ = λ λ + σ = 1 α = α = α + θ = 1 λ λ +
and the following characteristic exponent:
Ψ ( ξ ) = μ ξ i α log 1 λ λ + λ + λ i ξ + 1 λ + λ ξ 2 .
Proof. 
Γ ( β + ) = Γ ( 1 β + ) β + lim β + 0 Γ ( β + ) ( λ + i ξ ) β + λ + β + = log 1 1 λ + i ξ .
Similarly, (12) works for β 0 , and we have the characteristic exponent (10).
In addition, if α = α + = α , from (10), the characteristic exponent becomes:
Ψ ( ξ ) = μ ξ i α log 1 λ λ + λ + λ i ξ + 1 λ + λ ξ 2 ,
which is a Variance-Gamma (VG) distribution with parameter ( μ , λ λ + , 1 , α , 1 λ λ + ) . For more details on the VG model, refer to (Madan et al. 1998; Nzokem 2023c). □
Theorem 3 
(Cumulants κ k ). Consider a variable Y G T S ( μ , β + , β , α + , α , λ + , λ ) . The cumulants κ k of the GTS distribution are defined as follows:
κ 0 = 0 κ 1 = μ + α + Γ ( 1 β + ) λ + 1 β + α Γ ( 1 β ) λ 1 β κ k = α + Γ ( k β + ) λ + k β + + ( 1 ) k α Γ ( k β ) λ k β k N \ { 0 , 1 } .
Proof. 
We reconsider the characteristic exponent Ψ ( ξ ) in (7):
Ψ ( ξ ) = i μ ξ + 0 + e i y ξ 1 α + e λ + y y 1 + β + d y + 0 + e i y ξ 1 α e λ y y 1 + β d y = i μ ξ + α + k = 1 + Γ ( k β + ) λ + k β + ( i ξ ) k k ! + α k = 1 + Γ ( k β ) λ k β ( i ξ ) k k ! = i μ ξ + k = 1 + 1 k ! α + Γ ( k β + ) λ + k β + + α Γ ( k β ) λ k β ( 1 ) k ( i ξ ) k = k = 0 + κ k k ! ( i ξ ) k .
Hence, the k-th order cumulant κ k is given by comparing the coefficients of both polynomial functions in i ξ . For more details on the relationship between the characteristic exponent and cumulant functions, refer to (Feller 1971; Kendall 1945). □

2.2. GTS Distribution and Lévy Process

Corollary 1. 
Let Y = Y t be a Lévy process on R + generated by G T S ( μ , β + , β , α + , α , λ + , λ ) , and then
Y t G T S ( t μ , β + , β , t α + , t α , λ + , λ ) t R + .
Proof. 
Let Ψ ( ξ , t ) be the characteristic exponent of the Lévy process Y = Y t . By applying the infinitely divisible property, we have:
Ψ ( ξ , t ) = L o g E e i Y t ξ = t L o g E e i X ξ = t μ ξ i + t α + Γ ( β + ) ( λ + i ξ ) β + λ + β + + t α Γ ( β ) ( λ + i ξ ) β λ β
and we deduce that Y t G T S ( t μ , β + , β , t α + , t α , λ + , λ ) . □
Theorem 4 
(Asymptotic distribution of Generalized Tempered Stable distribution process). Let Y = Y t be a Lévy process on R generated by G T S ( μ , β + , β , α + , α , λ + , λ ) . Then, Y t converges in distribution to a Lévy process driving by a normal distribution with mean κ 1 and variance κ 2
Y t d N ( t κ 1 , t κ 2 ) as t +
where
κ 1 = μ + α + Γ ( 1 β + ) λ + 1 β + α Γ ( 1 β ) λ 1 β κ 2 = α + Γ ( 2 β + ) λ + 2 β + + α Γ ( 2 β ) λ 2 β .
Proof. 
The proof relies on the cumulant-generating function. As in (14), the characteristic exponent ( Ψ ( ξ ) ) can be written as follows:
Ψ ( ξ ) = L o g E e i Y ξ = j = 0 + κ j ( i ξ ) j j ! .
Let ϕ ( ξ , t ) be the characteristic function of the stochastic process Y t t κ 1 t κ 2 and we have:
ϕ ( ξ , t ) = E e i Y t t κ 1 t κ 2 ξ = e i t κ 1 t κ 2 ξ E e i ξ t κ 2 Y t = e i t κ 1 t κ 2 ξ e t Ψ ( ξ t κ 2 ) = e i t κ 1 t κ 2 ξ e j = 0 + t κ j j ! ( i ξ t κ 2 ) j = e ξ 2 2 + j = 3 + t κ j j ! ( i ξ t κ 2 ) j ,
lim t + j = 3 + t κ j j ! ( i ξ t κ 2 ) j = 0 lim t + ϕ ( ξ , t ) = lim t + e ξ 2 2 + j = 3 + t κ j j ! ( i ξ t κ 2 ) j = e 1 2 ξ 2 .

3. Multivariate Maximum-Likelihood Method

3.1. Maximum-Likelihood Method: Numerical Approach

From a probability density function f ( x , V ) with parameter V = ( μ , β + , β , α + , α , λ + , λ ) and sample data x = x j 1 j m , we define the likelihood function and its first and second derivatives as follows:
L m ( x , V ) = j = 1 m f ( x j , V ) , l m ( x , V ) = j = 1 m l o g ( f ( x j , V ) ) d l m ( x , V ) d V j = i = 1 m d f ( x i , V ) d V j f ( x i , V ) d 2 l m ( x , V ) d V k d V j = i = 1 m d 2 f ( x i , V ) d V k d V j f ( x i , V ) d f ( x i , V ) d V k f ( x i , V ) d f ( x i , V ) d V j f ( x i , V ) .
To perform the maximum of the likelihood function ( L m ( x , V ) ), we need the gradient of the likelihood function ( d l m ( x , V ) d V ), also known as the score function, and the Hessian matrix ( d 2 l m ( x , V ) d V d V ), which is the variance–covariance matrix generated by the likelihood function.
Given the parameters V = ( μ , β + , β , α + , α , λ + , λ ) and the sample data set X, we have the following quantities (21) from the previous development:
I m ( X , V ) = d l m ( x , V ) d V j 1 j p , I m ( X , V ) = d 2 l m ( x , V ) d V k d V j 1 k p 1 j p .
We use a computational algorithm built as a composite of a standard FRFT to compute the likelihood function and its derivatives (20) in the optimization process. More details on applying the composite of FRFTs for parameter estimations are provided in (Nzokem 2021b, 2021c; Nzokem and Montshiwa 2022, 2023); for other computations (such as probability density and cumulative functions), see (Cherubini et al. 2010; Eberlein 2014; Eberlein et al. 2010; Nzokem 2023b; Nzokem and Maposa 2024).
The computational algorithm yields a local solution, V, and a negative semi-definite matrix, I m ( x , V ) , when the following two conditions are satisfied:
I m ( x , V ) = 0 , U T I m ( X , V ) U 0 , U R p .
The solutions, V, in (22) are provided by the Newton–Raphson iteration algorithm Formula (23):
V n + 1 = V n I m ( x , V n ) 1 I m ( x , V n ) .
More details on the maximum-likelihood and Newton–Raphson iteration procedures are provided in (Giudici et al. 2013).

3.2. Asymptotic Distribution of the Maximum-Likelihood Estimator (MLE)

Theorem 5 
(Cramer-Rao). Let T = T ( X 1 , , X m ) be a statistic and write E [ T ] = k ( θ ) . Then, under suitable (smoothness) assumptions,
V a r [ T ] ( d E [ T ] d θ ) 2 m I ( θ ) .
For the proof of Theorem 5, refer to (Casella and Berger 2024; Van den Bos 2007).
Theorem 6 
(Consistency Estimator). Let X 1 , , X m be independent and identically distributed (i.i.d) random variables with density f ( x | θ ) satisfying some regularity conditions (Lehmann 1999). Let θ be the true parameter; then, there exists a sequence θ ^ m = θ m ( X 1 , , X m ) of local maxima of the likelihood function L m ( θ ) which is consistent, that is, which satisfies
θ ^ m a . s . θ as m + .
More details on the proof of Theorem 6 are provided in (Casella and Berger 2024; Lehmann 1999).
Theorem 7 
(Asymptotic Efficiency and Normality). Let X 1 , , X m be independent and identically distributed (i.i.d) random variables with density f ( x | θ ) satisfying some regularity conditions in (Lehmann 1999). There exists a solution θ ^ m = θ m ( X 1 , , X m ) of the likelihood equations which is consistent, and any such solution satisfies:
θ ^ m θ d N 0 , I m 1 ( θ ) as m + ,
where θ = ( θ 1 , , θ k ) is the actual parameter and I m ( θ ) is the Fisher information matrix.
More details on the proof of Theorem 7 are provided in (Hall and Oakes 2023; Lehmann 1999; Olive 2014).
Theorem 8 
(Likelihood Ratio Test). Suppose the assumptions of Theorem 7 hold and that ( θ ^ 1 n , , θ ^ k n ) are consistent roots of the likelihood equations for θ = ( θ 1 , , θ k ) . In addition, suppose that the corresponding assumptions hold for the parameter vector ( θ r + 1 , , θ k ) when r < k and that θ ^ ^ r + 1 , n , , θ ^ ^ k n are consistent roots of the likelihood equations for ( θ r + 1 , , θ k ) under the null hypothesis. We consider the likelihood ratio statistic
l m ( x , θ ^ ) l m ( x , θ ^ ^ )
where θ ^ ^ = ( θ 1 , , θ r , θ ^ ^ r + 1 , n , , θ ^ ^ k n ) . Then under the null hypothesis H 0 , if
Δ n = l m ( x , θ ^ ) l m ( x , θ ^ ^ ) ,
the statistic 2 Δ n has a limiting χ r 2 distribution.
More details on the proof of Theorem 8 are provided in (Lehmann 1999; Vuong 1989).

3.3. Asymptotic Test and Confidence Interval

The above results allow us to construct an asymptotically efficient estimator θ ^ m = ( θ ^ 1 m , , θ ^ k m ) of θ = ( θ 1 , , θ k ) such that
( θ ^ 1 m θ 1 , , θ ^ k m θ k )
has a joint multivariate limit distribution with mean ( 0 , , 0 ) and covariance matrix I m 1 ( θ ) = ( J i j ) . In particular, we have:
θ ^ j m θ j d N ( 0 , J j j ) as m + .
One approach to constructing an asymptotically valid confidence interval for the parameters is via the asymptotic distribution of the ML estimator (27). An approximate ( 1 α 2 ) confidence interval for θ ^ j m can be written as follows:
θ ^ j m ± z ( α 2 ) * J j j as m + ,
where z ( α 2 ) is the α 2 quantile of the standard normal distribution.

3.4. Applications of the Log-Likelihood Estimator to the Normal Distribution

We suppose the sample data x = x j 1 j m are independent observations and have a normal distribution (Mensah et al. 2023) with parameter V ( μ , σ 2 ) , that is, y N ( μ , σ 2 ) ; then, the density is
f ( y | V ) = ( 2 π σ 2 ) 1 2 e x p ( y μ ) 2 2 σ 2 .
The log-likelihood function in (20) becomes
l m ( x | V ) = j = 1 m l o g ( f ( x j | V ) ) = m 2 l o g ( 2 π σ 2 ) 1 2 σ 2 j = 1 m ( x j μ ) 2 .
The first-order derivatives of the log-likelihood function with respect to μ and σ 2 in (20) become
I m ( X , V ) = d l m ( x , V ) d μ d l m ( x , V ) d σ 2 = 1 σ 2 j = 1 m ( x j μ ) 1 2 σ 4 j = 1 m ( x j μ ) 2 m 2 σ 2 .
By setting I m ( X , V ) = 0 , we have
μ ^ = 1 m j = 1 m x j σ ^ 2 = 1 m j = 1 m ( x j μ ^ ) 2 .
The second-order derivative of the log-likelihood function with respect to μ and σ 2 in (20) becomes
I m ( X , V ) = d 2 l m ( x , V ) d μ 2 d l m ( x , V ) d μ d σ 2 d l m ( x , V ) d σ 2 d μ d 2 l m ( x , V ) ( d σ 2 ) 2 = m σ 2 j = 1 m ( x j μ ) σ 4 j = 1 m ( x j μ ) 2 σ 4 1 σ 6 j = 1 m ( x j μ ) 2 + m 2 σ 4 .
Refer to (Casella and Berger 2024) for more details.
We have the Fisher information matrix and the inverse:
I m ( V ) = E I m ( X , V ) = m σ 2 0 0 m 2 σ 4 , I m 1 ( V ) = σ 2 m 0 0 2 σ 4 m .
Corollary 2. 
The limiting distribution of the MLE is given by:
μ ^ σ ^ 2 d N μ σ 2 , σ 2 m 0 0 2 σ 4 m , as m + .
The proof of Corollary 2 comes from Theorem 7, Equation (26).

4. Fitting Tempered Stable Distribution to Cryptocurrencies: Bitcoin (BTC) and Ethereum

4.1. Data Summaries

Bitcoin was the first cryptocurrency created in 2009 by Satoshi Nakamoto. The idea behind Bitcoin was to create a peer-to-peer electronic payment system that allows online payments to be sent directly from one party to another without going through a financial institution (Nakamoto 2008). Since its inception, Bitcoin has grown in popularity and adoption and is now viewed as a viable legal tender in some countries. Bitcoin is currently used more as an investment tool, a risk-diversified tool, and less as a medium of exchange, a store of value, or a unit of account (Nzokem and Maposa 2024).
Bitcoin (BTC) and Ethereum (ETH) prices were extracted from CoinMarketCap. The period spans from 28 April 2013 to 4 July 2024 for Bitcoin and from 7 August 2015 to 4 July 2024 for Ethereum.
The daily price dynamics are provided in Figure 1. The prices have an increasing trend, even after having major significant increases and decreases over the studied period. Figure 1a,b show that Bitcoin outperforms Ethereum, which is the second-largest cryptocurrency by market capitalization after Bitcoin.
Let m be the number of observations and S j be the daily observed price on the day t j with j = 1 , , m . The daily return ( y j ) is computed as follows:
y j = log ( S j / S j 1 ) j = 2 , , m .
As shown in Figure 2a,b, the daily return reaches the lowest level ( 46 % for Bitcoin and 55 % for Ethereum) in the first quarter of 2020 amid the coronavirus pandemic and massive disruptions in the global economy. Nine values were identified as outliers and removed from the dataset to avoid a negative impact on the GTS model estimation and the empirical statistics.

4.2. Multidimensional Estimation Results for Cryptocurrencies

The results of the GTS parameter estimation are summarized in Table 1 for Bitcoin and Table 2 for Ethereum data. The brackets are the asymptotic standard errors computed using the inverse of the Hessian matrix built in (20). The ML estimate of μ is negative for both Bitcoin and Ethereum, while others are positive, as expected in the literature. The asymptotic standard error for μ is quite large and suggests that μ is not statistically significant at 5%.
The log-likelihood, Akaike’s information Criteria (AIC), and Bayesian information criteria (BIK) statistics show that the GTS distribution with seven parameters performs better than the two-parameter normal distribution (GBM). A comprehensive and detailed examination of the statistical significance of the results will be carried out in Section 6.
Table 1 summarizes the estimation results for Bitcoin returns. The skewness parameters ( λ + , λ ) are statistically significant at 5%. The difference is positive and statistically significant, which proves that the Bitcoin return is asymmetric and skewed to the left. The process intensity parameters ( α + , α ) are statistically significant at 5%. Similarly, the difference is positive and statistically significant, showing that the Bitcoin is more likely to produce positive returns than negative ones. The index of stability parameters ( β + , β ) are both statistically significant at 5%. However, the difference is positive but not statistically significant.
The GTS distribution with β = β + = β , called the Kobol distribution, was fitted to the Bitcoin data as well, and the estimation results are presented in Appendix B.1. As shown in Table A6, all the parameters are statistically significant at 5%, and have the expected positive sign. However, the likelihood ratio test in Table 6 shows that the GTS distribution in Table 1 is not significantly different from the Kobol distribution as the p-value ( 69.6 % ) is large. Refer to (Boyarchenko and Levendorskii 2002) for more details on the Kobol distribution.
As shown in Table 2, the parameters for Ethereum returns data are statistically significant at 5%, except μ and β . The difference ( λ + λ ) in skewness parameters is negative and not statistically significant, showing that the Ethereum return is asymmetric and skewed to the right. Similarly, the difference ( α + α ) in the intensity parameters is positive and not statistically significant, as shown the confidence interval. Contrary to the Bitcoin return, the Ethereum return has a larger process intensity, which provides evidence that Ethereum has a lower level of peakedness and a higher level of thickness.
We consider the following constraints λ = λ + = λ and β = β + = β , which are the Carr–Geman–Madan–Yor (CGMY) distribution, also called the Classical Tempered Stable Distribution. The CGMY distribution was fitted as well, and the estimation results are presented in Appendix B.2. As shown in Table A8, all the parameters are statistically significant at 5%, and have the expected positive sign. However, the likelihood ratio test in Table 6 shows, with a high p-value ( 35.3 % ), that the GTS distribution is not significantly different from the CGMY distribution, and the null hypothesis cannot be rejected. Refer to (Carr et al. 2003; Rachev et al. 2011) for more details on the CGMY distribution.
Table 1 and Table 2 summarize the last row of Table A1 and Table A2, respectively, in Appendix A.1, which describes the convergence process of the GTS parameter for Bitcoin and Ethereum data. The convergence process was obtained using the Newton–Raphson iteration algorithm (23). Each row has eleven columns made of the iteration number, the seven parameters μ , β + , β , α + , α , λ + , λ , and three statistical indicators, the log-likelihood ( L o g ( M L ) ), the norm of the partial derivatives ( | | d L o g ( M L ) d V | | ), and the maximum value of the eigenvalues ( M a x E i g e n V a l u e ). The statistical indicators aim at checking if the two necessary and sufficient conditions described in (22) are all met. L o g ( M L ) displays the value of the Naperian logarithm of the likelihood function L ( x , V ) , as described in (20); | | d L o g ( M L ) d V | | displays the value of the norm of the first derivatives ( d l ( x , V ) d V j ) described in (21); and M a x E i g e n V a l u e displays the maximum value of the seven eigenvalues generated by the Hessian matrix ( d 2 l ( x , V ) d V k d V j ), as described in (21).
Similarly, Table A7 and Table A9 describe the convergence process of the Kobol distribution parameter for Bitcoin returns and the CGMY distribution parameter for Ethereum returns.
GTS parameter estimations in Table 1 and Table 2 are used to evaluate the impact of each parameter on the GTS probability density function. As shown in Figure 3 and Figure 4, the effect of the GTS parameters on the probability density function has the same patterns on Bitcoin and Ethereum returns. However, the magnitudes are different. As shown in Figure 3a,b, β ( α ) has a higher effect on the probability density function (pdf) than β + ( α + ). However, λ and λ + in both graphs seem symmetric and have the same impact.

4.3. Evaluation of the Method of Moments

The method of moments estimates the parameters of the GTS distribution by equating empirical moments and the theoretical moments of the GTS distribution. We empirically estimate the kth moments ( m k = E ( x k ) ), based on sample data x = x j 1 j m as follows:
m ^ k = 1 M j = 1 m x j k for k = 1 , , 7 .
On the other side, the cumulants ( κ k ) in Theorem 4 can be related to the moment of the GTS distribution by the following relationship (Poloskov 2021; Rota and Shen 2000; Smith 1995):
m k = E ( x k ) = j = 1 k 1 k 1 j 1 κ j m k j + κ k for k = 1 , , 7 .
The method of moments estimator for V = ( μ , β + , β , α + , α , λ + , λ ) is defined as the solution to the following system of equations:
m ^ k = m k for k = 1 , , 7 .
The system of Equation (42) is often not analytically solvable. For the conditions of existence and uniqueness of the solution, refer to (Küchler and Tappe 2013).
The maximum-likelihood GTS parameter estimations in Table 1 and Table 2 are used to evaluate the system of equations in (42). As shown in Table 3, the solution of the maximum-likelihood method satisfies to a certain extent the equations for the first four moments: m ^ 1 , m ^ 2 , m ^ 3 , m ^ 4 in the system (42). The seventh-moment equation has the highest relative error: 89.9% for Bitcoin (BTC) and 68.3% for Ethereum. Therefore, the maximum likelihood GTS parameter estimation is not the same as the GTS parameter estimation from the method of moments.
In addition to the method of moments estimations, the lower relative errors in Table 3 show that empirical and theoretical standard deviation ( σ ), skewness, and kurtosis seem to be consistent for Bitcoin and Ethereum. The empirical and theoretical statistics show that the average Ethereum daily return is greater and more volatile than the Bitcoin daily returns. Both assets are thicker than the normal distribution. However, the daily return of Bitcoin is skewed to the left, whereas the daily return of Ethereum is skewed to the right.

5. Fitting Tempered Stable Distribution to Traditional Indices: S&P 500 and SPY EFT

5.1. Data Summaries

The Standard & Poor’s 500 Composite Stock Price Index, also known as the S&P 500, is a stock index that tracks the share prices of 500 of the largest public companies with stocks listed on the New York Stock Exchange (NYSE) and the Nasdaq in the United States. It was introduced in 1957 and is often treated as a proxy for describing the overall health of the stock market or the United States (US) economy. The SPDR S&P 500 ETF (SPY), also known as the SPY ETF, is an Exchange-Traded Fund (ETF) that tracks the performance of the S&P 500. SPY ETF provides a mutual fund’s diversification, the stock’s flexibility, and lower trading fees. The data were extracted from Yahoo Finance. The historical prices span from 04 January 2010 to 22 July 2024 and were adjusted for splits and dividends.
The daily price dynamics are provided in Figure 5. Prices have an increasing trend, even after being temporally disrupted in the first quarter of 2020 by the coronavirus pandemic. The S&P 500 is priced in thousands of US dollars, whereas the SPY ETF is in hundreds of US dollars. The SPY ETF is cheaper and provides all the attributes of the S&P 500 index, as shown in Figure 5a,b.
Let the number of observations be m and the daily observed price be S j on day t j with j = 1 , , m ; t 1 is the first observation date (4 January 2010) and t m is the last observation date (22 July 2024). The daily return, y j , is computed as in (43):
y j = log ( S j / S j 1 ) j = 2 , , m .
The SPY ETF aims to mirror the performance of the S&P 500. Figure 6a,b look similar, which is consistent with the goal of the SPY ETF. As shown in Figure 6a,b, the daily return reaches the lowest level ( 12.7 % for the S&P 500 and 11.5 % for the SPY ETF) in the first quarter of 2020 amid the coronavirus pandemic and massive disruptions in the global economy. Nine values were identified as outliers and removed from the dataset to avoid a negative impact on GTS model estimation and the empirical statistics.

5.2. Multidimensional Estimation Results for Traditional Indices

The estimation results are provided in Table 4 for S&P 500 return data and Table 5 for SPY EFT return data. As previously, the log-likelihood, AIC, and BIK statistics suggest that the GTS distribution with seven parameters performs better than the two-parameter normal distribution (GBM).
As shown in both Table 4 and Table 5, the ML estimate of μ is negative, while the others are positive, as expected in the literature. The asymptotic standard error for μ , β + and β are pretty large and result in μ , β + and β not being significantly different from zero.
However, other parameters have larger t-statistics ( | z | > 2 ) and are statistically significant at 5%. Except for the index of stability parameters ( β + , β ), the estimation results for the S&P 500 and SPY ETF indexes show that the difference in skewness parameters ( λ + , λ ) and intensity parameters ( α + , α ) are positive but are not statistically significant.
The hypothesis with β + = β = 0 was considered by fitting the S&P 500 and SPY ETF indexes to the bilateral Gamma distribution. The estimation results are summarized in Appendices Appendix C.1 and Appendix C.2. As shown in Table A10 and Table A12, the skewness parameters ( λ + , λ ) are positive and statistically significant, and the difference ( λ + λ ) is also positive and statistically significant, which proves that the S&P 500 and SPY ETF returns are skewed to the left. We have the same statistical features for the intensity parameters ( α + , α ), and both indexes are more likely to produce positive returns than negative returns. Refer to (Küchler and Tappe 2008; Nzokem 2021a) for more details on the bilateral Gamma distribution.
The likelihood ratio test in Table 6 shows that, even with non-statistically significant parameters, the GTS distribution fits significantly better than the bilateral Gamma distribution for both the S&P 500 and SPY ETF indexes. Contrary to the AIC statistics, the BIK statistics do not provide the same information. A comprehensive and detailed examination of the statistical significance of the results is carried out in Section 6.
Table 4 and Table 5 summarize the last row of Table A3 and Table A4, respectively, in Appendix A.1, which describes the convergence process of the GTS parameter for S&P 500 index, and SPY ETF return data. The convergence process was obtained using the Newton–Raphson iteration algorithm (23). Each row has eleven columns made of the iteration number; the seven parameters μ , β + , β , α + , α , λ + , λ ; and three statistical indicators, the log-likelihood ( L o g ( M L ) ), the norm of the partial derivatives ( | | d L o g ( M L ) d V | | ), and the maximum value of the eigenvalues ( M a x E i g e n V a l u e ). The statistical indicators aim at checking if the two necessary and sufficient conditions described in (22) are all met. L o g ( M L ) displays the value of the Naperian logarithm of the likelihood function L ( x , V ) , as described in (20); | | d L o g ( M L ) d V | | displays the value of the norm of the first derivatives ( d l ( x , V ) d V j ) described in Equation (21); and M a x E i g e n V a l u e displays the maximum value of the seven eigenvalues generated by the Hessian matrix ( d 2 l ( x , V ) d V k d V j ), as described in (21).
Similarly, Table A11 and Table A13 describe the convergence process of the bilateral Gamma distribution parameter for S&P 500 index and SPY ETF return data.
The GTS parameter estimations in Table 4 and Table 5 were used to evaluate the impact of the parameters on the GTS probability density function. As shown in Figure 7 and Figure 8, the effect of the GTS parameters on the probability density function generated by the S&P 500 and SPY ETF have the same patterns. However, the magnitudes are different. As shown in Figure 7a,b on the S&P 500 return data, β + ( α + ) has a higher effect on the probability density function than β ( α ). However, λ and λ + in Figure 7c are symmetric and have the same impact.

5.3. Evaluation of the Methods of Moments

The maximum-likelihood GTS parameter estimations in Table 4 and Table 5 are used to evaluate the system of equations in (42). As shown in Table 7, the solution of the maximum-likelihood method satisfies to a certain extent the equations (42) for the following first four moments: m ^ 1 , m ^ 2 , m ^ 4 , m ^ 5 . As for Bitcoin and Ethereum, the seventh-moment equation has the highest relative error: 53.3% for S&P 500 index and −85.9% for SPY ETF. Therefore, the maximum-likelihood GTS parameter estimation is not the GTS parameter estimation from the method of moments.
In addition to the moment estimations in Table 7, the empirical and theoretical standard deviation ( σ ), skewness, and kurtosis are consistent with lower relative errors for both S&P 500 and SPY ETF. The empirical and theoretical statistics show that both assets are skewed to the left and also thicker than the normal distribution.

6. Goodness-of-Fit Analysis

6.1. Kolmogorov–Smirnov (KS) Analysis

Given the sample of daily return { y 1 , y 2 y m } of size m and the empirical cumulative distribution function, F m ( x ) , for each index, the Kolmogorov–Smirnov (KS) test is performed under the null hypothesis, H 0 , that the sample { y 1 , y 2 y m } comes from the GTS distribution, F ( x ) . The cumulative distribution function of the theoretical distribution, F ( x ) , needs to be computed. The density function, f ( x ) , does not have a closed form, the same for the cumulative function, F ( x ) , in (45). However, we know the closed form of the Fourier of the density function, F [ f ] , and the relationship in (46) provides the Fourier of the cumulative distribution function, F [ F ] . The GTS distribution function, F ( x ) , was computed from the inverse of the Fourier of the cumulative distribution, F [ F ] , in (47):
Y G T S ( μ , β + , β , α + , α , λ + , f λ )
F ( x ) = x f ( t ) d t f is the density function of Y
F [ F ] ( x ) = F [ f ] ( x ) i x + π F [ f ] ( 0 ) δ ( x )
F ( x ) = 1 2 π + F [ f ] ( y ) i y e i x y d y + 1 2
See Appendix A in (Nzokem 2021a) for (46) proof.
The two-sided KS goodness-of-fit statistic ( D m ) is defined as follows:
D m = sup x | F ( x ) F m ( x ) | ,
where m is the sample size, F m ( x ) denotes the empirical cumulative distribution of { y 1 , y 2 y m } .
The distribution of Kolmogorov’s goodness-of-fit measure D m has been studied extensively in the literature. It was shown (Massey 1951) that the D m distribution is independent of the theoretical distribution, F ( x ) , under the null hypothesis, H 0 . The discrete, mixed, and discontinuous distributions case has also been studied (Dimitrova et al. 2020). Under the null hypothesis, H 0 , that the sample { y 1 , y 2 y m } of size m comes from the hypothesized continuous distribution, it was shown (An 1933) that the asymptotic statistic n D n converges to the Kolmogorov distribution.
The limiting form for the distribution function of Kolmogorov’s goodness-of-fit measure D m is
lim m + P r ( m D m x ) = 1 2 k = 1 + ( 1 ) k 1 e 2 k 2 x 2 = 2 π x k = 1 + e ( 2 k 1 ) 2 π 2 8 x 2 .
The first representation was given in (An 1933), and the second came from a standard relation for theta functions (Marsaglia et al. 2003).
As shown in Figure 9, the asymptotic statistic, n D n , is a positively skewed distribution with a mean and a standard deviation (Marsaglia et al. 2003) as follows.
μ = π 2 l o g ( 2 ) 0.8687 , σ = π 2 12 μ 2 0.2603 .
At a 5 % risk level, the risk threshold is d = 1.3581 and represents the area in the shaded area under the probability density function.
The p-value of the test statistic, D m , is computed based on (49) as follows:
p _ v a l u e = P r ( D m > D ^ m | H 0 ) = 1 P r ( m D m m D ^ m ) .
A p-value is defined as the probability that values are even more extreme or more in the tail than our test statistic. A small p-value leads to a rejection of the null hypothesis, H 0 , because the test statistic, D m , is already extreme. We reject the hypothesis if the p-value is less than our level of significance, which we take to be equal to 0.05.
D ^ m is a realization value of the KS estimator D m computed from the sample { y 1 , y 2 y m } . D ^ m is estimated (Krysicki et al. 1999) as follows:
D ^ m = M a x ( sup 0 j P | F ( x j ) F m ( x j ) | , sup 1 j P | F ( x j ) F m ( x j 1 ) | ) .
The following computations were performed for Bitcoin (BTC) data, and the quantity D ^ m was obtained:
sup 0 j P | F ( x j ) F m ( x j ) | = 0.01300 sup 1 j P | F ( x j ) F m ( x j 1 ) | = 0.00538 D ^ m = 0.01300 p _ v a l u e = p r o b ( m D m > 0.6903 | H 0 ) = 49.48 % .
For each index, KS statistics ( D ^ m ) and associated p-values were computed and summarized in Table 8, along with the index sample size, m.
The asymptotic statistics, n D n , produced from the two-parameter geometric Brownian motion (GBM) hypothesis, have high values and show that the GBM hypothesis is always rejected. On the other hand, the high p-values generated by the asymptotic statistics suggest insufficient evidence to reject the assumption that the data were randomly sampled from a GTS. The same observations work for the GTS variants: the Kobol, CGMY, and bilateral Gamma distributions. In addition, as shown the p-value indicator in Table 8, the GTS distribution outperforms the bilateral Gamma distribution for the S&P 500 and SPY ETF indexes. However, the Kobol and CGMY distributions, respectively, for Bitcoin and Ethereum have almost the same performance as the GTS distribution.

6.2. Anderson–Darling Test Analysis

The Anderson–Darling test (Anderson 2008) is a goodness-of-fit test that allows the control of the hypothesis that the distribution of a random variable observed in a sample follows a certain theoretical distribution. The Anderson–Darling statistic belongs to the class of quadratic EDF statistics (Stephens 1974) based on the empirical distribution function. The quadratic EDF statistics measure the distance between the hypothesized distribution ( F ( x ) ) and empirical distribution. It is defined as
m + F m ( x ) F ( x ) 2 w ( x ) d F x ,
where m is the number of elements in the sample, w ( x ) is a weighting function, and F m ( x ) is the empirical distribution function defined on the sample of size m.
When the weighting function is w ( x ) = 1 , the statistic (54) is the Cramér–-Von Mises statistic, while the Anderson–Darling statistic is obtained by choosing the weighting function w ( x ) = F ( x ) 1 F ( x ) . Compared with the Cramér–Von Mises statistic, the Anderson–Darling statistic places more weight on the tails of the distribution.
The Anderson–Darling statistic is
A m 2 = m + F m ( x ) F ( x ) F ( x ) 1 F ( x ) d F ( x ) .
It can be shown that the asymptotic distribution of the Anderson–Darling statistic, A m 2 , is independent of the theoretical distribution under the null hypothesis. The asymptotic distribution (Lewis 1961; Marsaglia and Marsaglia 2004) is defined as follows:
G ( x ) = lim m P r A m 2 < x = j = 0 + a j ( x b j ) 1 2 e x p ( b j x ) 0 + f j ( y ) e x p ( y 2 ) d y f j ( y ) = e x p 1 8 x b j y 2 x + b j , a j = ( 1 ) j ( 2 ) 1 2 ( 4 j + 1 ) Γ ( j + 1 2 ) j ! b j = 1 2 ( 4 j + 1 ) 2 π 2 .
As shown in Figure 10, the asymptotic distribution of the Anderson–Darling statistic ( A m 2 ) is a positively skewed distribution with a mean and a standard deviation (Anderson 2011) as follows
μ = 1 , σ = 2 3 ( π 2 9 ) 0.761 .
At a 5 % risk level, the risk threshold is d = 2.4941 and represents the area in the shaded area under the probability density function.
The p-value of the test statistic, A m 2 , is defined as follows:
p - v a l u e = p r o b ( A m 2 > A ^ m 2 | H 0 ) = 1 G ( A ^ m 2 ) .
In order to compute the Anderson–Darling statistic, A m 2 , in (55), the sample of daily return { y 1 , y 2 y m } of size m is arranged in ascending order: y ( 1 ) < y ( 2 ) < < y ( m ) . The Anderson–Darling statistic (Lewis 1961) then becomes
A m 2 = m 1 m j = 1 m ( 2 j 1 ) l o g ( F ( y ( j ) ) ) + ( 2 ( n j ) + 1 ) l o g ( F ( y ( j ) ) ) .
For each index, the Anderson–Darling statistic (59) is computed, along with the p-value statistic. Table 9 shows the KS statistics ( A m 2 ) and p-values for the GTS, GBM, and GTS variant distributions. While the two-parameter GBM hypothesis is always rejected, the GTS hypothesis is accepted and yields a very high p-value.
In addition, as shown by the p-value indicator in Table 9, the GTS distribution outperforms the bilateral Gamma distribution for the S&P 500 and SPY ETF indexes. However, the Kobol and CGMY distributions for Bitcoin and Ethereum, respectively, have almost the same performance as the GTS distribution.

6.3. Pearson’s Chi-Squared Test Analysis

Pearson’s chi-squared test (Schoutens 2003) counts the number of sample points falling into certain intervals and compares them with the expected number under the null hypothesis. Under the null hypothesis, H 0 , a random sample { y 1 , y 2 y m } comes from the GTS distribution, which has seven parameters estimated in Section 5. Suppose that m observations in the sample from a population are classified into K mutually exclusive classes with respective observed numbers of observations N j (for j = 1 , 2 , , K ), and a null hypothesis gives the probability Π j = F ( x j ) F ( x j 1 ) (47) that an observation falls into the jth class.
The following Pearson statistic calculates the value of the chi-squared goodness-of-fit test:
χ 2 ( K 1 p ) = j = 1 K N j m Π j 2 m Π j .
Under the null hypothesis assumption, as m goes to + , the limiting distribution (Schoutens 2003) of the Pearson statistic (60) follows the χ 2 ( K 1 p ) distribution with K 1 p degrees of freedom, and p is the number of estimated parameters.
Table 10 shows the Pearson chi-squared statistics ( χ ^ 2 ( K 1 p ) ), p-values, and class number for the GTS, GBM, and GTS variant distributions. While the two-parameter GBM hypothesis is always rejected, the GTS hypothesis is accepted and yields a high p-value.
In addition, as shown by the p-value indicator in Table 10, the GTS distribution outperforms the bilateral Gamma distribution for the S&P 500 and SPY ETF indexes. However, the Kobol and CGMY distributions for Bitcoin and Ethereum, respectively, have almost the same performance as the GTS distribution. For more details on the estimation of the Pearson statistic inputs under the GTS distribution, refer to Table A5 in Appendix A.2.

7. Conclusions

This study provides a methodology for fitting the rich class of the seven-parameter GTS distribution to financial data. Four historical prices were considered in the methodology application: two heavy-tailed data (Bitcoin and Ethereum returns) and two peaked data (S&P 500 and SPY ETF returns). The study used each historical data to fit the seven-parameter GTS distribution to the underlying data return distribution. The advanced fast FRFT scheme, based on the classic fast FRFT algorithm and the 11-point composite Newton–Cotes rule, was used to perform the maximum-likelihood estimation of seven parameters of the GTS distribution. The maximum likelihood estimate results show that, for each index, the location parameter, μ , is negative, while the others are positive, as expected in the literature. The statistical significance of the parameters was analyzed. The non-statistical significance of the index of stability parameters ( β + , β ) has led to the fitting of the Kobol, CGMY, and bilateral Gamma distributions. The goodness of fit was assessed through Kolmogorov–Smirnov, Anderson–Darling, and Pearson’s chi-squared statistics. While the two-parameter GBM hypothesis is always rejected, the goodness-of-fit analysis shows that the GTS distribution fits significantly the four historical data with a very high p-value.
As a main limitation of the study, the applied methodology is computation-intensive, and the researchers need good skills in computer programming. In future work, the estimated parameter of the GTS distribution will be used in the Ornstein–Uhlenbeck-type process to simulate the daily cumulative returns of financial assets.

Author Contributions

Conceptualization, A.N. and D.M.; Methodology, A.N. and D.M.; Visualization, A.N. and D.M.; Resources, A.N. and D.M.; Writing—Original Draft and Editing, A.N. and D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available in a publicly accessible repository: (for Bitcoin Prices: (accessed on 17 June 2024)) https://coinmarketcap.com/; (for S&P 500 index Prices: (accessed on 22 June 2024)) https://ca.finance.yahoo.com/.

Acknowledgments

The authors would like to thank the University of Limpopo for supporting the publication of this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Appendix A.1. Iterative Maximum-Likelihood Estimation (MLE) Procedure

Table A1. Convergence of the GTS parameter for Bitcoin return data.
Table A1. Convergence of the GTS parameter for Bitcoin return data.
Iterations μ β + β α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.73692460.46137830.26717870.81001730.51734700.21562890.1919378−10609.058282.67656663.6240151
2−0.79770190.46543900.21693920.78468170.49053320.21643950.2049523−10607.25326.7522215−1.6194299
3−0.44558410.38847210.32138670.77581500.51933950.23401870.1883953−10607.00150.13552913.0916011
4−0.76344450.45218780.22177020.79351290.49593710.22182530.2055181−10607.2104.8235882−2.6390063
5−0.49067460.41465310.34041760.77227290.52221100.22692020.1846457−10607.05967.66463389.0971871
6−0.55158340.44348270.33359050.77244840.51906190.21975660.1853686−10607.02217.4476962−0.4021102
7−0.49145860.43277140.35030120.76868830.52353610.22164500.1826269−10606.99116.2838831−0.1781480
8−0.29009080.38853500.39561860.75633570.53702600.23007720.1754994−10606.86412.0116477−2.4090216
9−0.27526980.38326600.39697040.75554560.53773670.23122240.1753571−10606.84711.4457840−2.5487401
10−0.26093390.37804000.39824560.75478120.53842090.23236320.1752258−10606.83210.8628213−2.6874876
11−0.20854090.35769270.40257620.75199660.54088640.23685440.1748113−10606.7828.3600783−3.4356438
12−0.19701090.35285750.40340020.75139230.54141380.23793620.1747455−10606.7727.6818408−3.6954428
13−0.17617330.34364160.40468180.75031910.54234140.24001740.1746675−10606.7566.2516380−4.2766527
14−0.16684210.33927940.40515220.74984920.54274380.24101200.1746529−10606.7505.5002876−4.5807361
15−0.15818600.33508540.40552560.74942090.54310900.24197400.1746517−10606.7454.7262048−4.8824015
16−0.15016000.33106120.40581660.74903110.54344040.24290240.1746615−10606.7413.9306487−5.1742197
17−0.12093760.31593010.40669450.74763930.54462240.24641220.1747326−10606.7342.8592342−6.2251311
18−0.12164870.31557070.40644380.74771790.54456080.24652470.1747753−10606.7340.0014787−6.2014232
19−0.12157140.31554830.40646350.74771420.54456520.24652960.1747719−10606.7341.82 × 10 06 −6.2026532
20−0.12157140.31554830.40646350.74771420.54456520.24652960.1747719−10606.7349.80 × 10 10 −6.2026530
Table A2. Convergence of the GTS parameter for Ethereum return data.
Table A2. Convergence of the GTS parameter for Ethereum return data.
Iterations μ β + β α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.12157140.31554830.40646350.74771420.54456520.24652960.1747719−9745.1712673.428257206.013602
2−0.17248350.33195050.40910220.73641290.54799340.22278700.1684568−9700.7152388.609394180.884105
3−0.20414180.33847420.41189290.73387940.55310830.20832030.1632896−9669.9862139.267660157.699659
4−0.40061570.35300350.43934740.75137840.61724250.11357430.1221930−9586.1151471.57047532.410140
5−0.64855510.44938170.44045080.92478870.72100310.14129490.1482307−9556.026380.60573756.584055
6−0.62905250.43714020.43595160.97807840.78247770.15823400.1608694−9553.00524.905322−0.719221
7−0.55454120.39947780.39181880.96274860.79365710.16524380.1724287−9552.8665.834338−0.847574
8−0.47448370.39139820.40934040.95823660.80228580.16651030.1699928−9552.8622.963350−0.933466
9−0.48255860.39021600.40513650.95807550.80076510.16674000.1706850−9552.8620.214871−0.931142
10−0.48536780.39043690.40448990.95824860.80047990.16671190.1707853−9552.8620.004754−0.931872
11−0.48538000.39043620.40448460.95824870.80047790.16671210.1707862−9552.8622.96 × 10 07 −0.931836
12−0.48538000.39043620.40448460.95824870.80047790.16671210.1707862−9552.8621.18 × 10 10 −0.931836
13−0.48538000.39043620.40448460.95824870.80047790.16671210.1707862−9552.8621.27 × 10 11 −0.931836
Table A3. Convergence of the GTS parameter for S&P 500 return data.
Table A3. Convergence of the GTS parameter for S&P 500 return data.
Iterations μ β + β α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.26064260.340879790.022211410.787757290.597110611.288555131.01435308−4921.0858147.214541−0.476265
2−0.27478870.378485670.025178460.725382480.5946281.221079351.01081205−4920.9765107.910271−12.169518
3−0.28527430.345627420.016289720.783533610.580246581.274235440.9888729−4920.623623.7087311.9588258
4−0.29712540.379858150.053925930.740684720.559721791.227379860.96278568−4920.54934.214433560.29705471
5−0.34150820.426006750.04322390.697834970.561063651.182864940.966753−4920.572237.0642417−1.7903876
6−0.29958170.403151290.122365070.71682740.5221721.203833510.9117451−4920.5743.07232514−0.7101089
7−0.29446230.39772570.122187510.721743510.522600321.208997140.9121201−4920.57012.63567879−1.0187469
8−0.27674290.375610630.115610970.74276150.526967991.230673840.91742165−4920.55111.83311761−2.1103436
9−0.2742040.371779390.113558830.746597630.528143351.23455240.91893546−4920.54771.75839181−2.177405
10−0.25598120.341479260.096433120.778155810.537842211.265942490.93144448−4920.53081.33811298−2.6954121
11−0.24969770.329540130.089280690.791254940.541860441.278686420.93662846−4920.52910.79520373−2.8166517
12−0.24942370.328664950.088694450.792381610.542215611.279700940.93708759−4920.52910.00166731−2.6765739
13−0.24940720.328624620.088645690.792425790.542246321.279742780.93712865−4920.52910.00013552−2.6768326
14−0.24940820.328624280.088640470.792426190.542249441.279743120.93713293−4920.52911.47 × 10 05 −2.6766945
15−0.24940830.328624240.088639920.792426240.542249771.279743150.93713338−4920.52911.57 × 10 06 −2.67668
16−0.24940830.328624240.088639850.792426240.542249811.279743160.93713344−4920.52911.89 × 10 09 −2.6766783
17−0.24940830.328624240.088639850.792426240.542249811.279743160.93713344−4920.52912.09 × 10 10 −2.6766783
Table A4. Convergence of the GTS parameter for SPY EFT return data.
Table A4. Convergence of the GTS parameter for SPY EFT return data.
Iterations μ β + β α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.05186610.11618460.21865481.042692920.527125741.522449910.91168779−4894.227914.7801725−6.5141947
2−0.11024770.184912760.174784720.947566550.528442711.43993150.91415148−4893.827829.8166141−1.9290981
3−0.20942040.293775920.08914460.840291220.560545631.347972710.96500981−4893.355416.99400954.33892902
4−0.19855640.297582080.132300130.831561670.536560791.338330780.93230856−4893.420610.90487441.04588745
5−0.0788830.259399220.396115430.848656730.403655221.355959320.7410936−4895.8806241.02817894.6293224
6−0.07535710.267048570.337541580.841209080.454461641.34528230.80751063−4894.389925.1995505−2.805571
7−0.1966420.316243720.200685430.805091060.503223681.308376120.88967028−4893.888140.25755134.770691
8−0.18982830.30450470.159002910.813804510.526720751.313419120.91775259−4893.46946.29433991−4.6080872
9−0.22752140.329409960.107705350.793402150.550200251.293604490.95260474−4893.30497.34361008−8.1891832
10−0.27262830.349724650.016012220.780615230.597360041.281533041.01757433−4893.219214.0784211−3.6408772
11−0.24998160.326452860.018515240.802173010.600181541.302436721.01792703−4893.21256.27794755−4.6455546
12−0.25759530.338325960.026433210.790083830.594506371.290852151.01101001−4893.2081.23298227−6.8035318
13−0.26070710.340526440.020752520.788170750.598053761.288951611.01555438−4893.20760.07363298−6.71708
14−0.26063680.340888150.022270120.787746930.597070821.288545321.01430383−4893.20760.00156771−6.6908109
15−0.26064320.340879110.022206330.787758130.597113971.288555931.01435731−4893.20760.00010164−6.6915902
16−0.26064260.340879850.022211880.787757210.59711031.288555061.01435268−4893.20768.45 × 10 06 −6.6915177
17−0.26064260.340879790.022211410.787757290.597110611.288555131.01435308−4893.20767.21 × 10 07 −6.6915239

Appendix A.2. Pearson Statistic Inputs

Table A5. Observed versus expected statistics under GTS distribution.
Table A5. Observed versus expected statistics under GTS distribution.
BitcoinEthereumsp500SPY EFT
k x(k) n* Π k n(k) x(k) n* Π k n(k) x(k) n* Π k n(k) x(k) n* Π k n(k)
1−18.9887.5128−20.8617.5316−4.34110.26412−4.4058.32711
2−17.0804.1447−18.3215.5836−3.9355.4565−4.0074.5623
3−15.1726.6037−15.78110.01815−3.5298.4427−3.6087.1167
4−13.26410.6789−13.24118.33120−3.12313.13815−3.21011.14412
5−11.35617.58613−10.70034.42429−2.71720.58820−2.81117.53818
6−9.44829.66132−8.16066.98068−2.31132.54330−2.41327.77527
7−7.54051.65747−5.620137.268134−1.90552.02348−2.01544.35041
8−5.63294.188107−3.080305.591305−1.49984.47989−1.61671.61773
9−3.724184.486168−0.540769.951769−1.093140.406147−1.218117.552122
10−1.816411.5034192.000965.210966−0.687242.564244−0.819198.101186
110.0921195.72511864.540458.955466−0.281455.971456−0.421351.660348
122.0001150.47011597.080219.8732220.126896.809896−0.023725.476730
133.908473.1774699.620111.7601010.532749.1067330.376867.735867
145.816217.78322712.16059.253600.938430.8414260.774541.022522
157.724107.38710214.70032.379321.344234.6922601.173300.464325
169.63255.2725117.24118.099211.750126.8201371.571163.491189
1711.54029.2943919.78110.296122.15668.688571.96988.86282
1813.44815.8611422.3215.93952.56237.387332.36848.49136
1915.3568.729924.8613.46552.96820.460152.76626.60023
2017.2644.8664 5.09143.37411.256123.16514.66913
21 6.4196 14.06714 18.45020

Appendix B

Appendix B.1. Bitcoin BTC: Kobol Distribution (β = β = β+)

V ( d x ) = α + e λ + x x 1 + β 1 x > 0 + α e λ | x | | x | 1 + β 1 x < 0 d x
Ψ ( ξ ) = μ ξ i + Γ ( β ) α + ( ( λ + i ξ ) β λ + β ) + α ( ( λ + i ξ ) β λ β )
Table A6. Kobol maximum-likelihood estimation for Bitcoin return data.
Table A6. Kobol maximum-likelihood estimation for Bitcoin return data.
ModelParameterEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.292833(0.126)−2.322.1 × 10 02 −0.541−0.045
β 0.367074(0.086)4.271.9 × 10 05 0.1990.535
α + 0.755914(0.047)16.024.7 × 10 58 0.6630.848
α 0.535121(0.034)15.689.6 × 10 56 0.4680.602
λ + 0.235266(0.027)8.873.6 × 10 19 0.1830.287
λ 0.181602(0.023)7.949.8 × 10 16 0.1370.226
Log(ML)−10,607
AIC21,226
BIK21,264
Table A7. Convergence of the Kobol parameter for Bitcoin return data.
Table A7. Convergence of the Kobol parameter for Bitcoin return data.
Iterations μ β α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.12157140.31554830.74771420.54456520.24652960.17477186−10614.93879450.055697425.6678081
2−0.2551720.365169580.732151190.531942530.229551860.17909072−10607.0105851.74982347−46.893383
3−0.29122760.370968540.754101080.535294390.233877160.18070591−10606.812361.484798964−53.728563
4−0.29224080.365743330.75595820.535084030.235608190.18189913−10606.810410.258928464−53.391237
5−0.29286410.367143110.755911470.535122390.235248010.18158588−10606.810250.01286122−53.406734
6−0.2928370.367083190.755913820.535121070.235263570.18159941−10606.810250.00174219−53.40643
7−0.29283280.367073730.755914190.535120860.235266030.18160154−10606.810251.18 × 10 05 −53.406384
8−0.29283280.367073790.755914190.535120860.235266020.18160153−10606.810251.60 × 10 06 −53.406384
9−0.29283280.367073790.755914190.535120860.235266010.18160153−10606.810252.18 × 10 07 −53.406384
10−0.29283280.367073790.755914190.535120860.235266010.18160153−10606.810251.09 × 10 08 −53.406384

Appendix B.2. Ethereum: Carr–Geman–Madan–Yor (CGMY) Distributions

V ( d x ) = α e λ + x x 1 + β 1 x > 0 + α e λ | x | | x | 1 + β 1 x < 0 d x
Ψ ( ξ ) = μ ξ i + α Γ ( β ) ( ( λ + i ξ ) β λ + β ) + ( ( λ + i ξ ) β λ β )
Table A8. CGMY maximum-likelihood estimation for Ethereum return data.
Table A8. CGMY maximum-likelihood estimation for Ethereum return data.
ModelParameterEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.147089(0.079)−1.866.3 × 10 02 −0.302−0.008
β 0.398418(0.127)3.121.8 × 10 03 0.1480.649
α 0.887161(0.058)15.221.2 × 10 52 0.7731.001
λ + 0.155369(0.023)6.565.2 × 10 11 0.1090.202
λ 0.185991(0.025)7.292.9 × 10 13 0.1360.236
Log(ML)−9554
AIC19,118
BIK19,149
Table A9. Convergence of the CGMY parameter for Ethereum return data.
Table A9. Convergence of the CGMY parameter for Ethereum return data.
Iterations μ β α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
1−0.485380.390436160.958248750.166712080.17078617−9596.26581653.57149−140.21456
2−0.05451310.401482470.882056740.158753170.18060554−9554.683480.0921993−19.637869
3−0.14796320.390494340.882719980.156314080.18704084−9553.90653.84112398−44.76325
4−0.14658930.403454820.888689270.154503830.18507239−9553.90360.4436942−55.029934
5−0.14724640.396836220.886675970.155640170.18628001−9553.90260.14094418−51.274906
6−0.14702470.399076680.887360360.155255810.18587143−9553.90250.05563606−52.523819
7−0.14711480.398168410.887085690.155412270.18603772−9553.90250.02143506−52.017698
8−0.14708980.398420980.887162340.155368830.18599155−9553.90250.00019543−52.158334
9−0.147090.398418550.887161610.155369240.18599199−9553.90251.16 × 10 05 −52.156981
10−0.147090.398418670.887161640.155369220.18599197−9553.90251.78 × 10 06 −52.157046
11−0.147090.398418690.887161650.155369220.18599197−9553.90252.71 × 10 07 −52.157055
12−0.147090.398418690.887161650.155369220.18599197−9553.90254.14 × 10 08 −52.157057

Appendix C

Appendix C.1. S&P 500 Index: Bilateral Gamma (BG) Distribution (β = β+ = 0)

V ( d x ) = α + e λ + x x 1 x > 0 + α e λ | x | | x | 1 x < 0 d x
Ψ ( ξ ) = μ ξ i α + log 1 1 λ + i ξ α log 1 + 1 λ i ξ
Table A10. BG maximum-likelihood estimation for S&P 500 return data.
Table A10. BG maximum-likelihood estimation for S&P 500 return data.
ModelParameterEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.031467(0.010)−3.072.1 × 10 03 −0.052−0.011
α + 1.092741(0.058)18.982.6 × 10 80 0.9801.206
α 0.701784(0.042)16.802.3 × 10 63 0.6200.784
λ + 1.539690(0.064)22.823.1 × 10 115 1.4071.672
λ 1.110737(0.050)22.076.6 × 10 108 1.0121.209
Log(ML)−4925
AIC9859
BIK9890
Table A11. Convergence of the BG parameter for S&P 500 return data.
Table A11. Convergence of the BG parameter for S&P 500 return data.
Iterations μ α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
100.792426240.542249811.279743160.93713344−4951.14391138.53458−265.251
2−0.00384470.931534130.644612541.411321381.05504868−4931.7583549.025405−171.22804
3−0.01032141.030628460.701988681.494266671.10555794−4926.8412286.215345−126.18156
4−0.01863171.079229120.714219961.533773911.11392285−4925.393135.694287−113.12071
5−0.02794751.094502050.704930921.544181721.10795103−4924.706538.0551545−116.58686
6−0.03139511.093256630.701625811.540387661.10996346−4924.6211.54417452−120.06271
7−0.03146711.092741190.701783651.539690271.11073682−4924.62050.02788236−120.35435
8−0.03146641.092769710.701827881.539711271.11079928−4924.62050.00198685−120.34482
9−0.03146631.092772130.701831581.53971311.11080431−4924.62050.00016039−120.34394
10−0.03146621.092772320.701831881.539713251.11080472−4924.62051.29 × 10 05 −120.34387
11−0.03146621.092772340.70183191.539713261.11080475−4924.62051.04 × 10 06 −120.34387
12−0.03146621.092772340.70183191.539713261.11080476−4924.62058.43 × 10 08 −120.34387
13−0.03146621.092772340.70183191.539713261.11080476−4924.62056.80 × 10 09 −120.34387
14−0.03146621.092772340.70183191.539713261.11080476−4924.62055.63 × 10 10 −120.34387
15−0.03146621.092772340.70183191.539713261.11080476−4924.62055.73 × 10 11 −120.34387

Appendix C.2. SPY ETF: Bilateral Gamma (BG) Distribution (β = β+ = 0)

V ( d x ) = α + e λ + x x 1 x > 0 + α e λ | x | | x | 1 x < 0 d x
Ψ ( ξ ) = μ ξ i α + log 1 1 λ + i ξ α log 1 + 1 λ i ξ
Table A12. BG maximum-likelihood estimation for SPY EFT return data.
Table A12. BG maximum-likelihood estimation for SPY EFT return data.
ModelParameterEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ 0.015048(0.012)1.282.0 × 10 01 −0.0080.038
α + 1.068239(0.067)16.028.6 × 10 58 0.9381.199
α 0.764449(0.044)17.333.0 × 10 67 0.6780.851
λ + 1.525718(0.073)20.981.1 × 10 97 1.3831.668
λ 1.156439(0.052)22.151.1 × 10 108 1.0541.259
Log(ML)−4899
AIC9807
BIK9838
Table A13. Convergence of the BG parameter for SPY EFT return data.
Table A13. Convergence of the BG parameter for SPY EFT return data.
Iterations μ α + α λ + λ Log ( ML ) | | dLog ( ML ) dV | | MaxEigenValue
100.787757290.597110611.288555131.01435308−4918.7331406.35365−252.28104
20.028677730.975622630.675728271.468222490.97596762−4908.5992226.190986−116.11753
30.027274071.051275170.786183061.518465011.17883595−4899.079845.2281041−96.788275
40.008340891.076922510.752260451.53480031.14577232−4898.9955131.637516−107.77617
50.011269621.072423580.75684971.530119131.1494212−4898.75148.0005418−103.03258
60.013864781.069339210.761676681.526883031.15363987−4898.676311.3246873−100.1483
70.014927451.068230470.763975411.525732451.15588409−4898.66930.98802026−99.171136
80.015044641.068211190.764393891.525695281.15636567−4898.66930.02529683−99.040575
90.015047421.068235390.764441631.52571521.15642976−4898.66930.00300624−99.030178
100.015047621.068238810.764447641.525718031.15643796−4898.66930.00038489−99.028926
110.015047641.068239250.764448411.525718391.15643901−4898.66934.92 × 10 05 −99.028765
120.015047651.068239310.764448511.525718441.15643915−4898.66936.30 × 10 06 −99.028745
130.015047651.068239320.764448521.525718451.15643917−4898.66931.32 × 10 08 −99.028742
140.015047651.068239320.764448521.525718451.15643917−4898.66931.69 × 10 09 −99.028742
150.015047651.068239320.764448521.525718451.15643917−4898.66932.23 × 10 10 −99.028742

References

  1. An, Kolmogorov. 1933. Sulla determinazione empirica di una legge didistribuzione. Giorn Dell’inst Ital Degli Att 4: 89–91. [Google Scholar]
  2. Anderson, Theodore W. 2008. Anderson–darling test. In The Concise Encyclopedia of Statistics. Edited by Yadolah Dodge. New York: Springer, pp. 12–14. [Google Scholar] [CrossRef]
  3. Anderson, Theodore W. 2011. Anderson–darling tests of goodness-of-fit. In International Encyclopedia of Statistical Science. Edited by Miodrag Lovric. Berlin and Heidelberg: Springer, pp. 52–54. [Google Scholar] [CrossRef]
  4. Barndorff-Nielsen, Ole E., and Neil Shephard. 2002. Financial Volatility, Lévy Processes and Power Variation. Available online: https://www.olsendata.com/data_products/client_papers/papers/200206-NielsenShephard-FinVolLevyProcessPowerVar.pdf (accessed on 27 August 2024).
  5. Bianchi, Michele Leonardo, Stoyan V. Stoyanov, Gian Luca Tassinari, Frank J. Fabozzi, and Sergio M. Focardi. 2019. Handbook of Heavy-Tailed Distributions in Asset Management and Risk Management. Volume 7 of Financial Economics. Singapore: World Scientific Publishing. [Google Scholar] [CrossRef]
  6. Borak, Szymon, Wolfgang Härdle, and Rafał Weron. 2005. Stable distributions. In Statistical Tools for Finance and Insurance. Berlin and Heidelberg: Springer, pp. 21–44. [Google Scholar] [CrossRef]
  7. Boyarchenko, Svetlana, and Sergei Z. Levendorskii. 2002. Non-Gaussian Merton-Black-Scholes Theory. Singapore: World Scientific Publishing, vol. 9. [Google Scholar]
  8. Carr, Peter, Hélyette Geman, Dilip B. Madan, and Marc Yor. 2003. Stochastic volatility for lévy processes. Mathematical Finance 13: 345–82. [Google Scholar] [CrossRef]
  9. Casella, George, and Roger Berger. 2024. Statistical Inference. Chapman & Hall/CRC Texts in Statistical Science. New York: CRC Press. [Google Scholar]
  10. Cherubini, Umberto, Giovanni Della Lunga, Sabrina Mulinacci, and Pietro Rossi. 2010. Fourier Transform Methods in Finance. Hoboken: John Wiley & Sons. [Google Scholar]
  11. Dimitrova, Dimitrina S., Vladimir K. Kaishev, and Senren Tan. 2020. Computing the kolmogorov-smirnov distribution when the underlying cdf is purely discrete, mixed, or continuous. Journal of Statistical Software 95: 1–42. [Google Scholar] [CrossRef]
  12. Eberlein, Ernst. 2014. Fourier-based valuation methods in mathematical finance. In Quantitative Energy Finance: Modeling, Pricing, and Hedging in Energy and Commodity Markets. Edited by Fred Espen Benth, Valery A. Kholodnyi and Peter Laurence. New York: Springer, pp. 85–114. [Google Scholar] [CrossRef]
  13. Eberlein, Ernst, Kathrin Glau, and Antonis Papapantoleon. 2010. Analysis of fourier transform valuation formulas and applications. Applied Mathematical Finance 17: 211–40. [Google Scholar] [CrossRef]
  14. Fallahgoul, Hasan, and Gregoire Loeper. 2021. Modelling tail risk with tempered stable distributions: An overview. Annals of Operations Research 299: 1253–80. [Google Scholar]
  15. Fallahgoul, Hasan A., David Veredas, and Frank J. Fabozzi. 2019. Quantile-based inference for tempered stable distributions. Computational Economics 53: 51–83. [Google Scholar] [CrossRef]
  16. Feller, William. 1971. An Introduction to Probability Theory and its Applications, 2nd ed. New York: John Wiley & Sons, vol. 2. [Google Scholar]
  17. Giudici, Paolo, Geof H. Givens, and Bani K. Mallick. 2013. Wiley Series in Computational Statistics. Hoboken: Wiley Online Library. [Google Scholar]
  18. Grabchak, Michael, and Gennady Samorodnitsky. 2010. Do financial returns have finite or infinite variance? A paradox and an explanation. Quantitative Finance 10: 883–93. [Google Scholar] [CrossRef]
  19. Hall, W. Jackson, and David Oakes. 2023. A Course in the Large Sample Theory of Statistical Inference. New York: CRC Press. [Google Scholar]
  20. Ken-Iti, Sato. 2001. Basic results on lévy processes. In Lévy Processes: Theory and Applications. Edited by Ole E. Barndorff-Nielsen, Thomas Mikosch and Sidney I. Resnick. New York: Springer Science & Business Media, pp. 1–37. [Google Scholar] [CrossRef]
  21. Kendall, Maurice George. 1945. The Advanced Theory of Statistics, 2nd ed. London: Charles Griffin & Co. Ltd., vol. 1. [Google Scholar]
  22. Kim, Young Shin, Svetlozar T. Rachev, Michele Leonardo Bianchi, and Frank J. Fabozzi. 2009. A new tempered stable distribution and its application to finance. In Risk Assessment: Decisions in Banking and Finance. Edited by Georg Bol, Svetlozar T. Rachev and Reinhold Würth. Berlin and Heidelberg: Springer, pp. 77–109. [Google Scholar]
  23. Krysicki, W., J Bartos, W. Dyczka, K. Królikowska, and M. Wasilewski. 1999. Rachunek prawdopodobieństwa i statystyka matematyczna w zadaniach. Cz. II. Statystyka matematyczna. Warszawa: PWN. [Google Scholar]
  24. Küchler, Uwe, and Stefan Tappe. 2008. Bilateral gamma distributions and processes in financial mathematics. Stochastic Processes and their Applications 118: 261–83. [Google Scholar] [CrossRef]
  25. Küchler, Uwe, and Stefan Tappe. 2013. Tempered stable distributions and processes. Stochastic Processes and Their Applications 123: 4256–93. [Google Scholar] [CrossRef]
  26. Lehmann, Erich Leo. 1999. Elements of Large-Sample Theory. New York: Springer. [Google Scholar]
  27. Lewis, Peter A. W. 1961. Distribution of the anderson-darling statistic. The Annals of Mathematical Statistics 32: 1118–24. [Google Scholar] [CrossRef]
  28. Madan, Dilip B., Peter P. Carr, and Eric C. Chang. 1998. The variance gamma process and option pricing. Review of Finance 2: 79–105. [Google Scholar] [CrossRef]
  29. Marsaglia, George, and John Marsaglia. 2004. Evaluating the anderson-darling distribution. Journal of Statistical Software 9: 1–5. [Google Scholar] [CrossRef]
  30. Marsaglia, George, Wai Wan Tsang, and Jingbo Wang. 2003. Evaluating kolmogorov’s distribution. Journal of Statistical Software 8: 1–4. [Google Scholar] [CrossRef]
  31. Massey, Frank J., Jr. 1951. The kolmogorov-smirnov test for goodness of fit. Journal of the American Statistical Association 46: 68–78. [Google Scholar] [CrossRef]
  32. Massing, Till. 2024. Parametric estimation of tempered stable laws. ALEA Latin American Journal of Probability and Mathematical Statistics 21: 1567–600. [Google Scholar] [CrossRef]
  33. Mensah, Eric Teye, Alexander Boateng, Nana Kena Frempong, and Daniel Maposa. 2023. Simulating stock prices using geometric brownian motion model under normal and convoluted distributional assumptions. Scientific African 19: e01556. [Google Scholar] [CrossRef]
  34. Nakamoto, Satoshi. 2008. Bitcoin: A Peer-to-Peer Electronic Cash System. Decentralized Business Review. Available online: https://www.ussc.gov/sites/default/files/pdf/training/annual-national-training-seminar/2018/Emerging_Tech_Bitcoin_Crypto.pdf (accessed on 10 May 2024).
  35. Nolan, John P. 2020. Modeling with Stable Distributions. Cham: Springer International Publishing, chp. 2. pp. 25–52. [Google Scholar] [CrossRef]
  36. Nzokem, Aubain H. 2021a. Fitting infinitely divisible distribution: Case of gamma-variance model. arXiv arXiv:2104.07580. [Google Scholar]
  37. Nzokem, Aubain H. 2021b. Gamma variance model: Fractional fourier transform (FRFT). Journal of Physics: Conference Series 2090: 012094. [Google Scholar] [CrossRef]
  38. Nzokem, Aubain H. 2021c. Numerical solution of a gamma—Integral equation using a higher order composite newton-cotes formulas. Journal of Physics: Conference Series 2084: 012019. [Google Scholar] [CrossRef]
  39. Nzokem, Aubain H. 2023a. Enhanced the fast fractional fourier transform (frft) scheme using the closed newton-cotes rules. arXiv arXiv:2311.16379. [Google Scholar]
  40. Nzokem, Aubain H. 2023b. European option pricing under generalized tempered stable process: Empirical analysis. arXiv arXiv:2304.06060. [Google Scholar]
  41. Nzokem, Aubain H. 2023c. 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]
  42. Nzokem, Aubain H. 2024. Self-decomposable laws associated with general tempered stable (gts) distribution and their simulation applications. arXiv arXiv:2405.16614. [Google Scholar]
  43. Nzokem, Aubain H., and Daniel Maposa. 2024. Bitcoin versus s&p 500 index: Return and risk analysis. Mathematical and Computational Applications 29: 44. [Google Scholar] [CrossRef]
  44. Nzokem, Aubain H., and V. T. Montshiwa. 2022. Fitting generalized tempered stable distribution: Fractional fourier transform (frft) approach. arXiv arXiv:2205.00586. [Google Scholar]
  45. Nzokem, Aubain H., and V. T. Montshiwa. 2023. The ornstein–uhlenbeck process and variance gamma process: Parameter estimation and simulations. Thai Journal of Mathematics, 160–68. Available online: https://thaijmath2.in.cmu.ac.th/index.php/thaijmath/article/view/1477 (accessed on 17 June 2024).
  46. Olive, David J. 2014. Statistical Theory and Inference. New York: Springer. [Google Scholar]
  47. Poloskov, Igor E. 2021. Relations between cumulants and central moments and their applications. Journal of Physics: Conference Series 1794: 012004. [Google Scholar] [CrossRef]
  48. Rachev, Svetlozar T., Young Shin Kim, Michele L. Bianchi, and Frank J. Fabozzi. 2011. Stable and tempered stable distributions. In Financial Models with Lévy Processes and Volatility Clustering. Edited by Svetlozar T. Rachev, Young Shin Kim, Michele L. Bianchi and Frank J. Fabozzi. Volume 187 of The Frank J. Fabozzi Series; Hoboken: John Wiley & Sons, Ltd., chp. 3. pp. 57–85. [Google Scholar] [CrossRef]
  49. Rota, Gian-Carlo, and Jianhong Shen. 2000. On the combinatorics of cumulants. Journal of Combinatorial Theory, Series A 91: 283–304. [Google Scholar] [CrossRef]
  50. Sato, Ken-Iti. 1999. Lévy Processes and Infinitely Divisible Distributions. Cambridge: Cambridge University Press. [Google Scholar]
  51. Schoutens, Wim. 2003. Lévy Processes in Finance: Pricing Financial Derivatives. West Sussex: John Wiley & Sons. [Google Scholar]
  52. Smith, Peter J. 1995. A recursive formulation of the old problem of obtaining moments from cumulants and vice versa. The American Statistician 49: 217–18. [Google Scholar] [CrossRef]
  53. Stephens, Michael A. 1974. Edf statistics for goodness of fit and some comparisons. Journal of the American Statistical Association 69: 730–37. [Google Scholar] [CrossRef]
  54. Tsallis, Constantino. 1997. Lévy distributions. Physics World 10: 42. [Google Scholar] [CrossRef]
  55. Van den Bos, Adriaan. 2007. Precision and Accuracy. Hoboken: John Wiley & Sons, Ltd., chp. 4. pp. 45–97. [Google Scholar] [CrossRef]
  56. Vuong, Quang H. 1989. Likelihood ratio tests for model selection and non-nested hypotheses. Econometrica: Journal of the Econometric Society 57: 307–33. [Google Scholar] [CrossRef]
Figure 1. Daily price.
Figure 1. Daily price.
Jrfm 17 00531 g001
Figure 2. Daily return.
Figure 2. Daily return.
Jrfm 17 00531 g002
Figure 3. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (Bitcoin returns).
Figure 3. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (Bitcoin returns).
Jrfm 17 00531 g003
Figure 4. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (Ethereum returns).
Figure 4. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (Ethereum returns).
Jrfm 17 00531 g004
Figure 5. Daily price.
Figure 5. Daily price.
Jrfm 17 00531 g005
Figure 6. Daily return.
Figure 6. Daily return.
Jrfm 17 00531 g006
Figure 7. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (S&P 500 index).
Figure 7. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (S&P 500 index).
Jrfm 17 00531 g007
Figure 8. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (SPY EFT).
Figure 8. d f ( x , V ) d V j f ( x , V ) : Effect of parameters on the GTS probability density (SPY EFT).
Jrfm 17 00531 g008
Figure 9. Asymptotic statistic ( m D m ) probability density function (PDF).
Figure 9. Asymptotic statistic ( m D m ) probability density function (PDF).
Jrfm 17 00531 g009
Figure 10. Asymptotic Anderson–Darling statistic ( A m 2 ) probability density function (PDF).
Figure 10. Asymptotic Anderson–Darling statistic ( A m 2 ) probability density function (PDF).
Jrfm 17 00531 g010
Table 1. Maximum-likelihood GTS parameter estimation for Bitcoin.
Table 1. Maximum-likelihood GTS parameter estimation for Bitcoin.
ModelParameterEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.121571(0.375)−0.327.5 × 10 01 −0.8560.613
β + 0.315548(0.136)2.332.0 × 10 02 0.0500.581
β 0.406563(0.117)3.484.9 × 10 04 0.1780.635
α + 0.747714(0.047)15.766.2 × 10 56 0.6550.841
α 0.544565(0.037)14.564.8 × 10 48 0.4710.618
λ + 0.246530(0.036)6.914.9 × 10 12 0.1770.316
λ 0.174772(0.026)6.692.2 × 10 11 0.1240.226
Log(ML)−10,606
AIC21,227
BIK21,271
GBM μ 0.151997(0.060)2.511.2 × 10 02 0.0330.271
σ 3.865132(0.330)11.697.2 × 10 32 3.2174.513
Log(ML)−11,313
AIC22,630
BIK22,638
Table 2. Maximum-likelihood GTS parameter estimation for Ethereum.
Table 2. Maximum-likelihood GTS parameter estimation for Ethereum.
ModelParamEstimateStd Errz P r ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.4854(1.008)−0.486.3 × 10 01 −2.4611.491
β + 0.3904(0.164)2.381.7 × 10 02 0.0690.712
β 0.4045(0.210)1.935.4 × 10 02 −0.0070.816
α + 0.9582(0.106)9.011.1 × 10 19 0.7501.167
α 0.8005(0.110)7.254.2 × 10 13 0.5841.017
λ + 0.1667(0.029)5.721.1 × 10 08 0.1100.224
λ 0.1708(0.036)4.712.5 × 10 06 0.1100.242
Log(ML)−9552
AIC19,119
BIK19,162
GBM μ 0.267284(0.091)2.933.4 × 10 03 0.0880.446
σ 5.205539(0.672)7.741.0 × 10 14 3.8876.524
Log(ML)−9960
AIC19,925
BIK19,933
Table 3. Evaluation of the method of moments.
Table 3. Evaluation of the method of moments.
Bitcoin BTCEthereum
Empirical(1) Theoretical(2) ( 1 ) ( 2 ) 2 Empirical(1) Theoretical(2) ( 1 ) ( 2 ) 2
Sample size4083 3246
m ^ 1 0.1520.1520.0%0.2670.2670.0%
m ^ 2 14.96015.0200.4%27.16127.3880.8%
m ^ 3 −11.320−15.64027.6%55.36357.8674.3%
m ^ 4 203322569.8%5267630716.5%
m ^ 5 −5823−15,48062.3%22,3683251831.2%
m ^ 6 670,6951,123,21540.2%2,114,7884,361,56251.5%
m ^ 7 −1,997,196−19,777,98889.9%12,411,80939,253,00168.3%
Standard deviation 13.8653.8730.2%5.2065.2260.4%
Skewness 2−0.314−0.38718.8%0.2380.2525.2%
Kurtosis 39.15410.0829.2%7.1128.38515.2%
Max value28.052 29.013
Min value−26.620 −29.174
1  σ = κ 2 ; 2 Skewness is estimated as κ 3 κ 2 3 / 2 ; 3 Kurtosis is estimated as 3 + κ 4 κ 2 2 ; κ 1 , κ 2 and κ 2 are defined in (13).
Table 4. Maximum-likelihood GTS parameter estimation for S&P 500 index.
Table 4. Maximum-likelihood GTS parameter estimation for S&P 500 index.
ModelParamEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.249408(0.208)−1.202.3 × 10 01 −0.6580.159
β + 0.328624(0.308)1.072.9 × 10 01 −0.2750.932
β 0.088640(0.176)0.506.1 × 10 01 −0.2560.433
α + 0.792426(0.350)2.262.4 × 10 02 0.1061.479
α 0.542250(0.107)5.093.6 × 10 07 0.3330.751
λ + 1.279743(0.348)3.682.4 × 10 04 0.5971.962
λ 0.937133(0.144)6.508.0 × 10 11 0.6551.220
Log(ML)−4920
AIC9851
BIK9898
GBM μ 0.044875(0.018)2.511.2 × 10 02 0.0100.080
σ 1.081676(0.027)39.530.0001.0281.135
Log(ML)−5330
AIC10,665
BIK10,677
Table 5. Maximum-likelihood GTS parameter estimation for SPY EFT data.
Table 5. Maximum-likelihood GTS parameter estimation for SPY EFT data.
ModelParamEstimateStd Errz Pr ( Z > | z | ) [95% Conf.Interval]
GTS μ −0.260643(0.135)−1.945.3 × 10 02 −0.5240.003
β + 0.340880(0.189)1.807.1 × 10 02 −0.0300.711
β 0.022212(0.212)0.109.2 × 10 01 −0.3930.437
α + 0.787757(0.225)3.504.6 × 10 04 0.3471.229
α 0.597110(0.141)4.222.4 × 10 05 0.3200.874
λ + 1.288555(0.226)5.701.2 × 10 08 0.8461.731
λ 1.014353(0.177)5.749.4 × 10 09 0.6681.361
Log(ML)−4893
AIC9800
BIK9843
GBM μ 0.054344(0.017)3.131.8 × 10 03 0.0200.088
σ 1.050217(0.026)40.710.0001.0001.101
Log(ML)−54,275
AIC10,554
BIK10,566
Table 6. Likelihood ratio test statistic and p-value.
Table 6. Likelihood ratio test statistic and p-value.
GTSGTS Variants χ 2 -Valuedfp-Value
Log(ML)−10,606.73−10,606.810.152510.6962
BitcoinAIC21,227.4721,225.62
BIK21,271.6721,263.51
Log(ML)−9552.86−9553.902.081020.3533
EthereumAIC19,119.7219,117.81
BIK19,162.3219,148.23
Log(ML)−4920.52−4924.628.182820.0167
S&P 500AIC9851.069859.24
BIK9898.499890.26
Log(ML)−4893.21−4898.6710.923420.0042
SPY ETFAIC9800.429807.34
BIK9843.849838.36
Table 7. Evaluation of the methods of moments.
Table 7. Evaluation of the methods of moments.
S&P 500 IndexSPY ETF
Empirical(1) Theoretical(2) ( 1 ) ( 2 ) 2 Empirical(1) Theoretical(2) ( 1 ) ( 2 ) 2
Sample size3656 3655
m ^ 1 0.0450.045−0.5%0.0540.0540.0%
m ^ 2 1.0691.083−1.3%1.0531.0440.8%
m ^ 3 −0.447−0.34131.2%−0.214−0.351−39.0%
m ^ 4 8.3719.764−14.3%8.1977.6916.6%
m ^ 5 −16.386−11.12847.3%−3.969−12.717−68.8%
m ^ 6 193.563247.811−21.9%157.645162.048−2.7%
m ^ 7 −840.097−547.88253.3%−85.003−602.447−85.9%
Standard deviation 11.0821.0334.7%1.0501.0212.9%
Skewness 2−0.432−0.535−19.2%−0.358−0.490−26.9%
Kurtosis 38.4137.43513.1%7.4957.1774.4%
Max value6.797 6.501
Min value−7.901 −6.734
1  σ = κ 2 ; 2 Skewness is estimated as κ 3 κ 2 3 / 2 ; 3 Kurtosis is estimated as 3 + κ 4 κ 2 2 ; κ 1 , κ 2 and κ 2 are defined in (13).
Table 8. Kolmogorov–Smirnov statistics and p-values.
Table 8. Kolmogorov–Smirnov statistics and p-values.
GTSGBNGTS VariantsSample Size
Index D ^ m m D ^ m p-Value D ^ m m D ^ m p-Value D ^ m m D ^ m p-Valuem
Bicoin BTC0.0130.8300.4940.1066.8030.0000.014 10.8630.4454083
Ethereum0.0120.7210.6740.0925.2490.0000.013 20.7490.6273246
S&P 5000.0120.7500.6270.0915.5500.0000.014 30.8970.3953656
SPY ETF0.0140.8690.4360.0895.4380.0000.016 31.0100.2583655
1 Kobol distribution ( β = β = β + ); 2 Carr–Geman–Madan–Yor (CGMY) distributions ( β = β = β + ; α = α = α + ); 3 bilateral Gamma distribution ( β = β + = 0 ).
Table 9. Anderson–Darling statistics and p-values.
Table 9. Anderson–Darling statistics and p-values.
GTSGBNGTS VariantsSample Size
Index A ^ m 2 p-Value A ^ m 2 p-Value A ^ m 2 p-Valuem
Bicoin BTC0.10980.999999.7060.00000.1105 10.99994083
Ethereum0.10180.999959.1570.00010.2123 20.98663246
S&P 5000.30070.937654.3040.00010.5010 30.74583656
SPY ETF0.30170.936851.5160.00010.6684 30.58573655
1 Kobol distribution ( β = β = β + ); 2 Carr–Geman–Madan–Yor (CGMY) distributions ( β = β = β + ; α = α = α + ); 3 bilateral Gamma distribution ( β = β + = 0 ).
Table 10. Pearson statistics and p-values.
Table 10. Pearson statistics and p-values.
GTSGBNGTS VariantsClass Number
Index χ ^ 2 ( K 8 ) p-Value χ ^ 2 ( K 3 ) p-Value χ ^ 2 ( K p 1 ) p p-ValueK
Bicoin BTC12.2340.50813750.00012.549 160.56221
Ethereum6.9100.8638050.0008.618 250.85420
S&P 5009.8860.7035740.00012.844 350.61421
SPY ETF13.9550.3776050.00018.228 350.25121
1 Kobol distribution ( β = β = β + ); 2 Carr–Geman–Madan–Yor (CGMY) distributions ( β = β = β + ; α = α = α + ); 3 bilateral Gamma distribution ( β = β + = 0 ).
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

Nzokem, A.; Maposa, D. Fitting the Seven-Parameter Generalized Tempered Stable Distribution to Financial Data. J. Risk Financial Manag. 2024, 17, 531. https://doi.org/10.3390/jrfm17120531

AMA Style

Nzokem A, Maposa D. Fitting the Seven-Parameter Generalized Tempered Stable Distribution to Financial Data. Journal of Risk and Financial Management. 2024; 17(12):531. https://doi.org/10.3390/jrfm17120531

Chicago/Turabian Style

Nzokem, Aubain, and Daniel Maposa. 2024. "Fitting the Seven-Parameter Generalized Tempered Stable Distribution to Financial Data" Journal of Risk and Financial Management 17, no. 12: 531. https://doi.org/10.3390/jrfm17120531

APA Style

Nzokem, A., & Maposa, D. (2024). Fitting the Seven-Parameter Generalized Tempered Stable Distribution to Financial Data. Journal of Risk and Financial Management, 17(12), 531. https://doi.org/10.3390/jrfm17120531

Article Metrics

Back to TopTop