Next Article in Journal
Fractal Features of Fracture Networks and Key Attributes of Their Models
Next Article in Special Issue
Correlation Structure of Time-Changed Generalized Mixed Fractional Brownian Motion
Previous Article in Journal
The Effects of Stochastic Circular Pores on Splitting Tensile Behavior of Concrete Based on the Multifractal Theory
Previous Article in Special Issue
Arbitrage in the Hermite Binomial Market
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Parameter Estimation in Rough Bessel Model

by
Yuliya Mishura
1,2 and
Anton Yurchenko-Tytarenko
3,*
1
Department of Probability, Statistics and Actuarial Mathematics, Taras Shevchenko National University of Kyiv, 03127 Kyiv, Ukraine
2
Division of Mathematics and Physics, Mälardalen University, 722 18 Västerås, Sweden
3
Department of Mathematics, University of Oslo, 0851 Oslo, Norway
*
Author to whom correspondence should be addressed.
Fractal Fract. 2023, 7(7), 508; https://doi.org/10.3390/fractalfract7070508
Submission received: 25 May 2023 / Revised: 21 June 2023 / Accepted: 25 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Fractional Processes and Multidisciplinary Applications)

Abstract

:
In this paper, we construct consistent statistical estimators of the Hurst index, volatility coefficient, and drift parameter for Bessel processes driven by fractional Brownian motion with H < 1 / 2 . As an auxiliary result, we also prove the continuity of the fractional Bessel process. The results are illustrated with simulations.

1. Introduction

The Bessel process, defined as the square root X = Z of the solution to the stochastic differential equation (SDE)
d Z ( t ) = k d t + 2 Z ( t ) d W ( t ) , Z ( 0 ) = x 0 2 > 0 , k > 0 ,
is a well-known probabilistic model used in a wide range of fields including physics (e.g., in non-Abelian gauge field theories or for modeling Fermi acceleration and non-colliding particle systems, as can be seen in [1,2,3,4]) and finance (for modeling interest rates and stochastic volatility, as can be seen in [5,6,7,8]). If k N , X can be interpreted as the Euclidean norm
X ( t ) = B 1 2 ( t ) + + B k 2 ( t )
of a k-dimensional Brownian motion ( B 1 , . . . , B k ) that is connected with W in (1) via the relation
W ( t ) = i = 1 k 0 t B i ( s ) X ( s ) d B i ( s ) ,
and therefore the parameter k is often referred to as the dimension of the Bessel process X. Moreover, as it is shown in Section 3 of [9] (also see [10]), for all real k > 1 , the Bessel process X is the unique non-negative strong solution to the SDE
d X ( t ) = k 1 X ( t ) d t + d W ( t ) , X ( 0 ) = x 0 > 0 .
For more details on the various properties of the Bessel processes, we refer the reader to Chapter XI of [11], the book [12] which considers general equations of the type (3) or [13,14,15] which deal with the case 0 < k < 1 .
However, in many applications, standard Brownian motion may not adequately capture the desired level of complexity observed in real-life phenomena. For example, a number of empirical studies [16,17,18,19,20] point out the presence of memory in financial markets. Other sources [21,22,23,24] indicate that models exhibiting very low Hölder regularity are better suited to reflect the behavior of market volatility. Given that Bessel-type processes are frequently employed in stochastic volatility modeling (see, e.g., Chapter 6 of [25]), there is a natural inclination to enhance them by incorporating the aforementioned memory or roughness. A common way to achieve such an effect is to replace (in some sense) the standard Brownian driver W with a fractional Brownian motion B H = { B H ( t ) , t 0 } , i.e., a centered Gaussian process with the covariance function
E B H ( t ) B H ( s ) = 1 2 t 2 H + s 2 H | t s | 2 H , s , t 0 .
For instance, Refs. [26,27,28] used the property (2) as the starting point of their modification and defined fractional Bessel process ρ H for k N as
ρ H ( t ) = ( B 1 H ( t ) ) 2 + + ( B k H ( t ) ) 2
with B 1 H ,…, B k H being k-independent fractional Brownian motions. In particular, they prove that ρ H admits a representation
d ρ H ( t ) = H ( k 1 ) t 2 H 1 ρ H ( t ) d t + i = 1 k B i H ( t ) ρ H ( t ) d B i H ( t ) ,
where the integrals with respect to fractional Brownian motions are understood in the divergence sense.
Another possible “fractionalization” of the Bessel process may be achieved by replacing W with B H directly in (3). This approach was discussed in detail in the series of papers [29,30,31,32]: according to it, the fractional Bessel process X H = { X H ( t ) , t 0 } is defined as the a.s. point-wise limit
X H ( t ) : = lim ε 0 X ε H ( t )
of stochastic processes given by the SDE
X ε H ( t ) = x 0 + 0 t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s + σ B H ( t ) .
In the present paper, we consider the latter notion of the fractional Bessel process and assume that it is driven by a fractional Brownian motion with the Hurst index H 0 , 1 2 . Our goal is to perform the statistical estimation of three parameters: the Hurst index H mentioned above, the diffusion coefficient σ > 0 regulating the magnitude of the driving noise as well as the drift parameter a > 0 that serves as a direct counterpart of k 1 in (3).
  • In order to estimate H and σ , we use the standard technique based on quadratic variations of fractional Brownian motion. The main challenge arises in the limit L H ( t ) : = lim ε 0 0 t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s : when H < 1 2 , Corollary 3.3 of [32] only guarantees its continuity in t almost everywhere with respect to the Lebesgue measure. It is not enough for the analysis of quadratic variations of L H , so we prove that L H is continuous at every t 0 and utilize this fact to obtain consistent estimators of H and σ .
  • Our estimator of a is, in turn, based on the unconventional technique tailored for our specific model. The problem here is that it is currently not known whether the fractional Bessel process exhibits any ergodic properties typically utilized for drift parameter estimation. In our analysis, we exploit the explicit dynamics of X H instead to compare the behavior of X H ( T ) and 0 T 1 X H ( t ) d t when T .
This paper is structured as follows. In Section 2, we provide the definition of the fractional Bessel process and prove the continuity of the latter. Section 3 is devoted to the estimation of H and σ . In Section 4, we present our estimator of the drift and prove its consistency. Section 5 contains simulations. In Appendix A, we prove a technical result related to the finiteness of the limit (5).

2. Rough Bessel Processes

Let B H = { B H ( t ) , t 0 } be a fractional Brownian motion with Hurst index H 0 , 1 2 .
Remark 1. 
Using the Kolmogorov–Chentsov theorem, it is possible to prove that B H has a modification with paths that are locally Hölder continuous up to the order H, i.e., for any T > 0 and λ ( 0 , H ) , there exists a positive random variable Λ = Λ T , λ such that
| B H ( t 1 ) B H ( t 2 ) | Λ | t 1 t 2 | λ , t 1 , t 2 [ 0 , T ] .
Moreover, by [33], the random variable Λ can be chosen in such a way that for any p > 0
E [ Λ p ] < .
In what follows, we always consider this modification of B H .
For any ε > 0 , x 0 > 0 , a > 0 , b 0 , and σ > 0 , consider a random process X ε H = { X ε H ( t ) , t 0 } given by a stochastic differential Equation (6). Note that the SDE (6) has Lipschitz continuous drift and additive noise and hence, by the standard Picard iteration argument applied pathwise, (6) has a unique solution for each ε > 0 . Moreover, by Lemma 2.1 of [32] (see also Lemma 1.6 in [29]), for any ε 1 < ε 2 ,
X ε 1 H ( t ) X ε 2 H ( t ) , t [ 0 , T ] , ω Ω ,
and hence, for any t 0 , one can define the limit
X H ( t ) : = lim ε 0 X ε H ( t ) .
Definition 1. 
The process X H defined by (9) will be called a fractional or rough Bessel process.
Remark 2. 
This construction of rough Bessel processes was introduced and extensively studied in [32]. In particular, it was shown that
(1) 
The limit (9) is finite, i.e., with probability 1,
X H ( t ) <
for all t 0 ;
(2) 
With probability 1, X H ( t ) 0 for all t 0 and, moreover, X H ( t ) > 0 for almost all t 0 .
It should be noted that one of the possible cases is missing in Step 2 of the proof of Theorem 2.1 in [32] concerning (10). For the reader’s convenience, we provide the completed version of it in Appendix A.
Since the limit (9) is well defined and finite, the limit
lim ε 0 0 t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s = lim ε 0 X ε H ( t ) x 0 σ B H ( t ) = X H ( t ) x 0 σ B H ( t )
also exists and is finite. In what follows, we will use the notation
L H ( t ) : = lim ε 0 0 t 1 X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s ,
i.e., X H satisfies the equation
X H ( t ) = x 0 + a L H ( t ) + σ B H ( t ) , t 0 .
By Fatou’s lemma, for any t 0 ,
0 t 1 X H ( s ) d s lim inf ε 0 0 t 1 X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s = L H ( t ) < .
Denote
R H ( t ) : = L H ( t ) 0 t 1 X H ( s ) d s
and re-write (13) as
X H ( t ) = x 0 + a 0 t 1 X H ( s ) d s + σ B H ( t ) + a R H ( t ) .
At the moment, it is not clear whether R H ( t ) = 0 for all t 0 . However, we have the following result.
Proposition 1. (1) 
With a probability of 1, there exists τ > 0 such that for all t τ
X H ( t ) > 0 .
(2) 
Let τ > 0 be such that X H ( τ ) > 0 . Then, there exists a neighborhood ( τ 1 , τ 2 ) τ such that, for all t 1 , t 2 ( τ 1 , τ 2 ) , R H ( t 1 ) R H ( t 2 ) = 0 , i.e.,
L H ( t 2 ) L H ( t 1 ) = t 1 t 2 1 X H ( s ) d s .
Proof. 
Item (1) directly follows from Theorem 3.3 in [30] so let us prove the claim (2). Let τ > 0 be such that X H ( τ ) > 0 . Since X ε H ( τ ) X H ( τ ) as ε 0 , there exists ε 0 > 0 such that for all ε ε 0 , X ε H ( τ ) > 0 . Moreover, the process X ε 0 H is continuous with respect to t; hence, there exists a neighborhood ( τ 1 , τ 2 ) τ and some positive value x > 0 such that for all t ( τ 1 , τ 2 )
X ε 0 H ( t ) > x .
In particular, (8) implies that, for all ε ε 0
X ε H ( t ) > x , t ( τ 1 , τ 2 ) .
Therefore, for any ε ε 0 and t ( τ 1 , τ 2 ) ,
1 X ε H ( t ) 1 { X ε H ( t ) > 0 } + ε < 1 x
and hence, by the dominated convergence theorem, for any t 1 , t 2 ( τ 1 , τ 2 ) , t 1 < t 2 ,
L H ( t 2 ) L H ( t 1 ) = lim ε 0 t 1 t 2 1 X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s = t 1 t 2 1 X H ( s ) d s .
Corollary 1. 
With a probability of 1, there exists τ > 0 such that, for all t 2 > t 1 τ
R H ( t 1 ) = R H ( t 2 )
and
L H ( t 2 ) L H ( t 1 ) = t 1 t 2 1 X H ( s ) d s .
Note that Corollary 3.3 from [32] only establishes the continuity of L H almost everywhere on R + with respect to the Lebesgue measure. It is not enough for our purposes: in order to estimate H and σ , we want to utilize the behavior of quadratic variations of X H and any possible discontinuities of L H would create substantial obstacles for our analysis. It turns out, however, that L H (and hence X H ) is continuous at all points t 0 . The corresponding theorem is provided below.
Theorem 1. 
1. 
The process L H defined by (12) is non-decreasing.
2. 
The processes X H = { X H ( t ) , t 0 } and L H = { L H ( t ) , t 0 } have continuous paths.
Proof. 
First of all, observe that for any 0 t 1 < t 2 T
L H ( t 1 ) = lim ε 0 0 t 1 1 X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s lim ε 0 0 t 2 1 X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s = L H ( t 2 ) ,
i.e., the process L H is indeed non-decreasing. Next, fix an arbitrary deterministic T > 0 , choose an arbitrary λ ( 0 , H ) and let Λ = Λ T , λ be such that, for all s , t [ 0 , T ] ,
| B H ( t ) B H ( s ) | Λ | t s | λ .
Note that the monotonicity of L H implies that its discontinuities can only take the form of positive jumps and, moreover, left and right limits
L H ( t ) : = lim δ 0 L H ( t δ ) , L H ( t + ) : = lim δ 0 L H ( t + δ )
are well defined at any point t ( 0 , T ) . Next, observe that the limit process X H defined by (9) satisfies the equation
X H ( t ) = x 0 + a L H ( t ) + σ B H ( t ) , t [ 0 , T ] ,
and hence points of discontinuity of X H coincide with the ones of L H , and can only occur in the form of positive jumps of the same size as the corresponding jumps of L H and, finally, one can define the limits
X H ( t ) : = lim δ 0 X H ( t δ ) , X H ( t + ) : = lim δ 0 X H ( t + δ ) .
Assume that τ [ 0 , T ] is a point of discontinuity of L H (and hence X H ) for some ω Ω and observe that X H ( τ ) = 0 since, otherwise, τ cannot be a point of discontinuity of X H and L H by (15). Let α > 0 be such that
X H ( τ + ) = L H ( τ + ) = α
and δ > 0 be such that for all t ( τ , τ + δ )
X H ( t ) > 9 α 10 .
Next, choose δ 0 < δ such that
4 a α δ 0 + σ Λ δ 0 λ α 4 ,
where λ ( 0 , H ) and Λ > 0 are from (16). Since X ε H ( τ + δ 0 ) X H ( τ + δ 0 ) as ε 0 , there exists ε 0 > 0 such that for all ε ( 0 , ε 0 ]
X ε H ( τ + δ 0 ) > 4 α 5 , X ε H ( τ ) 0
Furthermore, X ε H ( τ ) X H ( τ ) = 0 as ε 0 hence X ε H ( τ ) 0 for all ε > 0 , and therefore, for any ε ( 0 , ε 0 ] one can define
τ ε : = sup s ( τ , τ + δ 0 ) : X ε H ( s ) = α 4
and
τ ε + : = sup s ( τ ε , τ + δ 0 ) : X ε H ( s ) = 3 α 4 .
By continuity, it is evident that X ε H ( τ ε ) = α 4 , X ε H ( τ ε + ) = 3 α 4 and X ε H ( t ) α 4 for any t [ τ ε , τ ε + ] . Therefore for any ε ( 0 , ε 0 ]
α 2 = X ε H ( τ ε + ) X ε H ( τ ε ) = τ ε τ ε + a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s + σ B H ( τ ε + ) B H ( τ ε ) 4 a α ( τ ε + τ ε ) + σ Λ ( τ ε + τ ε ) λ 4 a α δ 0 + σ Λ δ 0 λ < α 4 ,
which gives a contradiction. Therefore, α = 0 and X H ( τ ) = X H ( τ + ) , i.e., X H (and hence L H ) is right-continuous.
It remains to notice that X H ( t ) = X H ( t ) for any t ( 0 , T ] . Indeed, as mentioned above, X H is continuous at t if X H ( t ) > 0 . If X H ( t ) = 0 , since X H ( t ) 0 can only have positive jumps,
X H ( t ) X H ( t ) = X H ( t )
and, since X H can potentially have only positive jumps, X H ( t ) = 0 . □
Note that the pre-limit processes X ε H given by (6) are monotonically non-decreasing with respect to ε . Moreover, with probability 1, their point-wise limit is a continuous function by Theorem 1. Therefore, by Dini’s theorem, we immediately obtain uniform convergence which is summarized in the following corollary.
Corollary 2. 
For any T > 0 ,
sup t [ 0 , T ] | X H ( t ) X ε H ( t ) | 0 a . s . , ε 0 .

3. Estimation of Hurst Index and Volatility Coefficient

Let us now move on to the parameter estimation for the rough Bessel processes. Our first goal is to estimate the Hurst index H 0 , 1 2 and the volatility parameter σ > 0 . Note that, by Theorem 1, the process L H is continuous and non-decreasing and hence has zero quadratic variation. Therefore, we can utilize the standard estimation technique based on power variations. For more details on this method, we refer the reader to [34].
Throughout this section, we assume that the rough Bessel process X H = { X H ( t ) , t 0 } satisfies Equation (14) with (unknown) parameters a , σ > 0 , H < 1 / 2 and is observed on a discrete uniform partition 0 = t 0 < t 1 < . . . < t n = T of a fixed compact [ 0 , T ] , t k : = k T n .

3.1. Estimation of H

Let us start with some useful notation for quadratic variations.
Notation 1. 
For a stochastic process ξ = { ξ ( t ) , t [ 0 , T ] } observed on a discrete uniform partition 0 = t 0 < t 1 < . . . < t n = T , t k : = k T n , denote
V 1 , 2 n ( ξ ) : = k = 0 n 1 ( ξ ( t k + 1 ) ξ ( t k ) ) 2
and
V 2 , 2 n ( ξ ) : = k = 0 n 2 ( ξ ( t k + 2 ) 2 ξ ( t k + 1 ) + ξ ( t k ) ) 2 .
In order to estimate the unknown Hurst parameter, we utilize the following result from Proposition 4.2 in [35] and Lemma 2.10 in [34].
Theorem 2. 
Let H ( 0 , 1 ) , σ > 0 and V 1 , 2 n ( σ B H ) , V 2 , 2 n ( σ B H ) be as defined in Notation 1. Then, with the probability of 1,
n T 1 + 2 H V 1 , 2 n ( σ B H ) σ 2 T
and
n T 1 + 2 H V 2 , 2 n ( σ B H ) ( 4 2 2 H ) σ 2 T
as n .
As established in Theorem 1, the process a L H in the right-hand side of (14) is a continuous process of bounded variation and hence, with a probability of 1,
V 1 , 2 n ( a L H ) 0 , V 2 , 2 n ( a L H ) 0 , n .
Therefore, we immediately obtain the following corollary.
Corollary 3. 
Let X H be a rough Bessel process given by (13) observed on a discrete uniform partition 0 = t 0 < t 1 < . . . < t n = T , t k = k T n . Then, with probability 1,
n T 1 + 2 H V 1 , 2 n ( X H ) σ 2 T
and
n T 1 + 2 H V 2 , 2 n ( X H ) ( 4 2 2 H ) σ 2 T
as n .
Next, define
H ^ : = log 4 V 2 , 2 n ( X H ) V 1 , 2 n ( X H ) 2 log 2 .
Theorem 3. 
The estimator H ^ given by (19) is a (strongly) consistent estimator of the Hurst index H, i.e., with a probability of 1,
H ^ H , n .
Proof. 
The statement immediately follows from Corollary 3. □
Remark 3. 
By Corollary 3, with probability 1,
4 V 2 , 2 n ( X H ) V 1 , 2 n ( X H ) 2 2 H > 1 , n ,
and hence the logarithm in (19) is well defined for large enough values of n. Moreover,
V 2 , 2 n ( X H ) = k = 0 n 2 ( X H ( t k + 2 ) 2 X H ( t k + 1 ) + X H ( t k ) ) 2 = k = 0 n 2 ( X H ( t k + 2 ) X H ( t k + 1 ) ) ( X H ( t k + 1 ) X H ( t k ) ) 2 2 k = 0 n 2 ( X H ( t k + 2 ) X H ( t k + 1 ) ) 2 + 2 k = 0 n 2 ( X H ( t k + 1 ) X H ( t k ) ) 2 4 V 1 , 2 n ( X H ) ,
i.e.,
4 V 2 , 2 n ( X H ) V 1 , 2 n ( X H ) 0 ,
and the equality in (20) only occurs if x 0 = X H ( t 0 ) = X H ( t 1 ) = . . . = X H ( t n ) . It remains an open question whether such an event can occur with positive probability; however, in practice, it is never the case.

3.2. Estimation of σ

Next, assume that the Hurst index H is known and the goal is to estimate the volatility coefficient σ . Using Corollary 3, it can be performed in a straightforward manner. Namely, we have the following result.
Theorem 4. 
Let the Hurst index H be known. Then, the estimator
σ ^ = 1 T H n 1 + 2 H V 1 , 2 n ( X H )
is a (strongly) consistent estimator of the volatility coefficient σ, i.e., with a probability of 1,
σ ^ σ , n .
Proof. 
The statement immediately follows from Corollary 3. □

4. Drift Parameter Estimation

Let us move to the estimation of the drift parameter a in (14). In contrast to the high-frequency setting considered in Section 3, we will now assume that we observe a continuous path of X H on [ 0 , T ] with T . We study the following estimator of the parameter a:
a ^ ( T ) : = X H ( T ) 0 T 1 X H ( t ) d t .
Remark 4. 
Note that we do not assume any prior knowledge of the parameters H or σ.
Let us start by presenting a well-known result related to the growth of fractional Brownian motion (for more details, see [36], Section B 3.5 of [34] or Corollary 2.2 in [30]).
Theorem 5. 
For any δ > 0 , there exist random time τ δ and positive random variable ξ such that, with a probability of 1,
max t [ 0 , T ] | B H ( t ) | ξ T H + δ
for all T > τ δ .
Lemma 1. 
Let X H be a rough Bessel process satisfying (14). Then, with a probability of 1,
lim inf T 1 T 0 T 1 X H ( t ) d t > 0 .
Proof. 
Take an arbitrary δ 0 , 1 2 H and fix an ω Ω such that the corresponding path of B H is continuous, (23) holds and R H becomes constant after some random time point τ as described in Corollary 1.
Assume that (24) does not hold, i.e., there exists an increasing sequence { T n , n 1 } such that T n as n , T 1 exceeds τ δ from Theorem 5, and
1 T n 0 T n 1 X H ( t ) d t 0 , n .
Integrating both sides of (14) from 0 to T n , we obtain that, for all n 1
0 T n X H ( t ) d t = x 0 T n + a 0 T n 0 t 1 X H ( s ) d s d t + σ 0 T n B H ( t ) d t + a 0 T n R H ( t ) d t .
Let us prove that the equality (25) cannot hold by comparing the asymptotics of its left- and right-hand sides.
On the one hand, using the Cauchy–Schwartz inequality, it is easy to see that, for all T > 0 ,
0 T X H ( t ) d t 0 T 1 X H ( t ) d t T 2 ,
so
1 T n 3 / 2 0 T n X H ( s ) d s , n .
On the other hand,
x 0 T n T n 3 / 2 0 , n ,
1 T n 3 / 2 0 T n 0 t 1 X H ( s ) d s d t = 1 T n 3 / 2 0 T n 1 X H ( t ) ( T n t ) d t T n T n 3 / 2 0 T n 1 X H ( t ) d t 0 , n ,
and, by (23),
1 T n 3 / 2 0 T n B H ( t ) d t T 1 + H + δ T n 3 / 2 ξ 1 + H + δ 0 , n .
Finally, since R H is constant after some time point τ ,
1 T n 3 / 2 0 T n R H ( t ) d t 0 , n .
In other words, if such a sequence { T n , n 1 } exists, the left-hand side of (25) divided by T n 3 / 2 converges to whereas the right-hand side divided by T n 3 / 2 converges to zero. We obtain a contradiction that proves (24). □
We are now ready to move to the main result of this section.
Theorem 6. 
With a probability of 1,
a ^ ( T ) a , T .
Proof. 
First of all, note that, by Lemma 1,
x 0 0 T 1 X H ( s ) d s 0 a . s .
when T . Next, since there exists τ > 0 such that R H ( t ) = 0 for all t τ , Lemma 1 also implies that
a R H ( t ) 0 T 1 X H ( s ) d s 0 a . s .
when T . Finally, fixing δ 0 , 1 2 H and using (23), we can deduce that
| B H ( T ) | 0 T 1 X H ( s ) d s 1 T 1 2 H δ ξ 1 T 0 T 1 X H ( s ) d s 0 a . s .
as T . Therefore, with probability 1,
a ^ 1 ( T ) = x 0 + a 0 t 1 X H ( s ) d s + σ B H ( t ) + a R H ( t ) 0 T 1 X H ( s ) d s = a + x 0 0 T 1 X H ( s ) d s + σ B H ( t ) 0 T 1 X H ( s ) d s + a R H ( t ) 0 T 1 X H ( s ) d s a , T .

5. Simulations

In this section, we illustrate our results with simulations. In all cases, we use the standard Euler scheme to simulate the pre-limit processes X ε H given by (6) with ε = 0.0001 . Each estimator was tested on 1000 samples. All simulations were performed using R programming language; in order to simulate the trajectories of fractional Brownian motion, the package somebm was used.

5.1. Simulation of H ^ and σ ^

We start by testing the estimators H ^ and σ ^ given by (19) and (21), respectively. The real values of the parameters are chosen to be H = 0.3 and σ = 1 whereas the time horizon is T = 1 .
At first, we analyze H ^ for different sizes n of the partition. For each n, 1000 trajectories of X H were generated and the value of H ^ was computed for each of the generated paths. Table 1 contains the mean, variance, and coefficient of variation (CV, i.e., the ratio of the standard deviation to the mean) for H ^ and Figure 1 depicts the corresponding box-and-whisker plots.
Next, we perform the same procedure for the estimator σ ^ given by (21) under the assumption that the real value of H is known. The results are given in Table 2 and Figure 2.
Finally, we consider a more realistic situation when neither H nor σ are known. In this case, we first estimate H using the estimator H ^ given by (19) and then plug in the result in the estimator σ ^ defined by (21). The results of this estimation of σ are given in Table 3 and Figure 3.
Overall, in all cases, the simulations show the convergence of the estimators to the true values. This also applies to the least favorable situation of estimation of σ when H is unknown.

5.2. Simulation of a ^ ( T )

Next, we test the performance of the estimator a ^ ( T ) for different time horizons T using the same numerical techniques as in Section 5.1. The real values of the parameters are once again chosen to be H = 0.3 , σ = 1 and the integral 0 T 1 X H ( t ) d t is approximated by the integral sum
1 n i = 0 [ T n ] 1 1 max X H i n , 0.001
with n = 10 , 000 . For each T, 1000 simulations of a ^ ( T ) were performed. The results are given in Table 4 and Figure 4.
Simulations confirm the convergence of a ^ ( T ) to a. However, they also indicate that relatively high values of T are required to guarantee the reasonable variance of the estimator.

Author Contributions

Investigation, Y.M. and A.Y.-T.; writing—original draft preparation, Y.M. and A.Y.-T. All authors have read and agreed to the published version of the manuscript.

Funding

The present research was carried out within the frame and support of the ToppForsk project nr. 274410 of the Research Council of Norway with title STORM: Stochastics for Time-Space Risk Models. The first author is supported by The Swedish Foundation for Strategic Research, grant Nr. UKR22-0017 and by Japan Science and Technology Agency CREST, project reference number JPMJCR2115.

Data Availability Statement

Data sharing not applicable to this paper.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Finiteness of Rough Bessel Processes

As mentioned in Remark 2, the goal of this appendix is to prove the following result.
Theorem A1. 
Let B H be a fractional Brownian motion with H < 1 / 2 and X H be defined by (9). Then, with probability 1, for any t 0
X H ( t ) < .
Before moving to the proof of Theorem A1, let us first give some auxiliary results.
Proposition A1. 
Let X 1 H = { X 1 H ( t ) , t 0 } be a stochastic process defined by
X 1 H ( t ) = x 0 + 0 t a X 1 H ( s ) 1 { X 1 H ( s ) > 0 } + 1 d s + σ B H ( t ) .
Fix T > 0 , λ ( 0 , H ) and let Λ = Λ T , λ be a random variable such that (7) holds. Then, with probability 1,
sup t [ 0 , T ] | X 1 H ( t ) | x 0 + a T + σ T λ Λ .
Proof. 
Fix ω Ω such that the corresponding path t B H ( ω , t ) is Hölder continuous. In particular,
max t [ 0 , T ] | B H ( ω , t ) | Λ T λ .
Taking into account that a x 1 { x > 0 } + 1 a for any x R , we can write
| X 1 H ( t ) | x 0 + 0 t a d s + σ T λ Λ x 0 + a T + σ T λ Λ .
Proposition A2. 
Let X ε H be defined by (6). Fix T > 0 , λ ( 0 , H ) and let Λ = Λ T , λ be a random variable such that (7) holds. Then, for any ε ( 0 , 1 ] and t [ 0 , T ] ,
| X ε H ( t ) | < x 0 + a T 2 x 0 1 + σ T λ Λ a . s .
Proof. 
Fix ε ( 0 , 1 ] and take ω Ω such that the corresponding path t B H ( ω , t ) is Hölder continuous. In particular,
max t [ 0 , T ] | B H ( ω , t ) | Λ T λ ,
Denote also
τ 1 , ε : = sup t [ 0 , T ] | s [ 0 , t ] : X ε H ( s ) x 0 2 .
Our goal is to prove (A2) separately for t [ 0 , τ 1 , ε ] and t [ τ 1 , ε , T ] .
Step 1:  t [ 0 , τ 1 , ε ] . In this case, X ε H ( s ) x 0 2 for all s [ 0 , t ] and hence, we can write
| X ε H ( t ) | = x 0 + 0 t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s + σ B H ( t ) x 0 + 0 t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s + σ T λ Λ x 0 + a 0 t 2 x 0 d s + σ T λ Λ x 0 + 2 a T x 0 + σ T λ Λ x 0 + a T 2 x 0 1 + σ T λ Λ
as required.
Step 2:  t [ τ 1 , ε , T ] . Consider
τ 2 , ε ( t ) : = sup s [ τ 1 , ε , t ] | | X ε H ( s ) | < x 0 2 .
If τ 2 , ε ( t ) = t , (A2) holds automatically, so let us assume that τ 2 , ε ( t ) < t . In this case, by continuity, X ε H ( τ 2 , ε ( t ) ) = x 0 2 and, moreover, | X ε H ( s ) | x 0 2 for all s [ τ 2 , ε ( t ) , t ] and we have two possibilities.
(1)
If X ε H ( s ) x 0 2 for all s [ τ 2 , ε ( t ) , t ] , then
X 1 H ( s ) X ε H ( s ) < 0
and hence, by Proposition A1,
| X ε H ( s ) | | X 1 H ( s ) | x 0 + a T + σ T λ Λ .
(2)
If X ε H ( s ) x 0 2 for all s [ τ 2 , ε ( t ) , t ] , then, just like in Case 1 above,
| X ε H ( t ) | X ε H ( τ 2 , ε ( t ) ) + τ 2 , ε ( t ) t a X ε H ( s ) 1 { X ε H ( s ) > 0 } + ε d s + σ B H ( t ) B H ( τ 2 , ε ( t ) ) x 0 2 + a τ 2 , ε ( t ) t 2 x 0 d s + σ T λ Λ x 0 2 + 2 a T x 0 + σ T λ Λ .
Hence, summarizing (A3) and (A4), for any t [ τ 1 , ε , T ]
X H ( t ) x 0 + a T 2 x 0 1 + σ T λ Λ ,
which ends the proof. □
Proof of Theorem A1. 
By definition, for any t 0
X ε ( t ) X H ( t ) a . s . , ε 0 ,
so (A1) follows immediately from Proposition A2. □

References

  1. Bray, A.J. Random walks in logarithmic and power-law potentials, nonuniversal persistence, and vortex dynamics in the two-dimensional XY model. Phys. Rev. E 2000, 62, 103–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Guarnieri, F.; Moon, W.; Wettlaufer, J.S. Solution of the Fokker-Planck equation with a logarithmic potential and mixed eigenvalue spectrum. J. Math. Phys. 2017, 58, 093301. [Google Scholar] [CrossRef] [Green Version]
  3. Horibe, M.; Hosoya, A.; Sakamoto, J. Stochastic quantization and the Gribov problem in non-abelian gauge theories. Prog. Theor. Phys. 1983, 70, 1636–1647. [Google Scholar] [CrossRef] [Green Version]
  4. Valenti, D.; Chichigina, O.A.; Dubkov, A.A.; Spagnolo, B. Stochastic acceleration in generalized squared Bessel processes. J. Stat. Mech. Theory Exp. 2015, 2015, P02012. [Google Scholar] [CrossRef]
  5. Cox, J.C.; Ingersoll, J.E.; Ross, S.A. A re-examination of traditional hypotheses about the term structure of interest rates. J. Financ. 1981, 36, 769–799. [Google Scholar] [CrossRef]
  6. Cox, J.C.; Ingersoll, J.E.; Ross, S.A. An intertemporal general equilibrium model of asset prices. Econom. J. Econom. Soc. 1985, 53, 363. [Google Scholar] [CrossRef] [Green Version]
  7. Cox, J.C.; Ingersoll, J.E.; Ross, S.A. A theory of the term structure of interest rates. Econom. J. Econom. Soc. 1985, 53, 385. [Google Scholar] [CrossRef]
  8. Heston, S.L. A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev. Financ. Stud. 1993, 6, 327–343. [Google Scholar] [CrossRef] [Green Version]
  9. Cherny, A.S. On the strong and weak solutions of stochastic differential equations governing Bessel processes. Stochastics Stochastics Rep. 2000, 70, 213–219. [Google Scholar] [CrossRef]
  10. Mishura, Y.; Yurchenko-Tytarenko, A. Standard and fractional reflected Ornstein–Uhlenbeck processes as the limits of square roots of Cox–Ingersoll–Ross processes. Stochastics. Int. J. Probab. Stoch. Process. 2022, 95, 99–117. [Google Scholar] [CrossRef]
  11. Revuz, D.; Yor, M. Continuous Martingales and Brownian Motion; Springer: Berlin/Heidelberg, Germany, 1999. [Google Scholar]
  12. Cherny, A.S.; Engelbert, H.J. Singular Stochastic Differential Equations; Springer: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
  13. Bertoin, J. Complements on the Hilbert transform and the fractional derivative of Brownian local times. Kyoto J. Math. 1990, 30, 651–670. [Google Scholar] [CrossRef]
  14. Bertoin, J. Excursions of a BES0(d) and its drift term (0 < d < 1). Probab. Theory Relat. Fields 1990, 84, 231–250. [Google Scholar] [CrossRef]
  15. Mishura, Y.; Pilipenko, A.; Yurchenko-Tytarenko, A. Low-dimensional Cox-Ingersoll-Ross process. arXiv 2023, arXiv:2303.12911. [Google Scholar] [CrossRef]
  16. Bollerslev, T.; Ole Mikkelsen, H. Modeling and pricing long memory in stock market volatility. J. Econom. 1996, 73, 151–184. [Google Scholar] [CrossRef]
  17. Breidt, F.J.; Crato, N.; de Lima, P. The detection and estimation of long memory in stochastic volatility. J. Econom. 1998, 83, 325–348. [Google Scholar] [CrossRef] [Green Version]
  18. Cont, R. Long range dependence in financial markets. In Fractals in Engineering; Springer: London, UK, 2005; pp. 159–179. [Google Scholar]
  19. Ding, Z.; Granger, C.W.J. Modeling volatility persistence of speculative returns: A new approach. J. Econom. 1996, 73, 185–215. [Google Scholar] [CrossRef]
  20. Ding, Z.; Granger, C.W.J.; Engle, R.F. A long memory property of stock market returns and a new model. J. Empir. Financ. 1993, 1, 83–106. [Google Scholar] [CrossRef]
  21. Alòs, E.; León, J.A.; Vives, J. On the short-time behavior of the implied volatility for jump-diffusion models with stochastic volatility. Financ. Stochastics 2007, 11, 571–589. [Google Scholar] [CrossRef] [Green Version]
  22. Fukasawa, M. Volatility has to be rough. Quant. Financ. 2021, 21, 1–8. [Google Scholar] [CrossRef]
  23. Fukasawa, M.; Takabatake, T.; Westphal, R. Is Volatility Rough? arXiv 2019, arXiv:1905.04852. [Google Scholar] [CrossRef]
  24. Gatheral, J.; Jaisson, T.; Rosenbaum, M. Volatility is rough. Quant. Financ. 2018, 18, 933–949. [Google Scholar] [CrossRef]
  25. Jeanblanc, M.; Yor, M.; Chesney, M. Mathematical Methods for Financial Markets; Springer: London, UK, 2009. [Google Scholar]
  26. Essaky, E.H.; Nualart, D. On the 1/H-variation of the divergence integral with respect to fractional Brownian motion with Hurst parameter H<1/2. Stoch. Process. Their Appl. 2015, 125, 4117–4141. [Google Scholar] [CrossRef]
  27. Guerra, J.M.E.; Nualart, D. The 1/H-variation of the divergence integral with respect to the fractional Brownian motion for H>1/2 and fractional Bessel processes. Stoch. Process. Their Appl. 2005, 115, 91–115. [Google Scholar] [CrossRef] [Green Version]
  28. Hu, Y.; Nualart, D. Some processes associated with fractional Bessel processes. J. Theor. Probab. 2005, 18, 377–397. [Google Scholar] [CrossRef] [Green Version]
  29. Di Nunno, G.; Mishura, Y.; Yurchenko-Tytarenko, A. Sandwiched SDEs with unbounded drift driven by Hölder noises. Adv. Appl. Probab. Appear. 2023, 55, 1–38. [Google Scholar] [CrossRef]
  30. Mishura, Y.; Ralchenko, K. Fractional diffusion Bessel processes with Hurst index H∈(0,12). arXiv 2023, arXiv:2304.06328. [Google Scholar] [CrossRef]
  31. Mishura, Y.; Yurchenko-Tytarenko, A. Fractional Cox–Ingersoll–Ross process with non-zero “mean”. Mod. Stochastics Theory Appl. 2018, 5, 99–111. [Google Scholar] [CrossRef]
  32. Mishura, Y.; Yurchenko-Tytarenko, A. Fractional Cox–Ingersoll–Ross process with small Hurst indices. Mod. Stochastics Theory Appl. 2018, 6, 13–39. [Google Scholar] [CrossRef]
  33. Azmoodeh, E.; Sottinen, T.; Viitasaari, L.; Yazigi, A. Necessary and sufficient conditions for Hölder continuity of Gaussian processes. Statist. Probab. Lett. 2014, 94, 230–235. [Google Scholar] [CrossRef] [Green Version]
  34. Kubilius, K.; Mishura, Y.; Ralchenko, K. Parameter Estimation in Fractional Diffusion Models; Springer International Publishing: Cham, Switzerland, 2017. [Google Scholar]
  35. Viitasaari, L. Necessary and sufficient conditions for limit theorems for quadratic variations of Gaussian sequences. Probab. Surv. 2019, 16, 62–98. [Google Scholar] [CrossRef]
  36. Kozachenko, Y.; Melnikov, A.; Mishura, Y. On drift parameter estimation in models with fractional Brownian motion. Statistics 2015, 49, 35–62. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Performance of the estimator H ^ given by (19) for different sizes of the partition. The real value H = 0.3 is depicted with a gray-dashed line.
Figure 1. Performance of the estimator H ^ given by (19) for different sizes of the partition. The real value H = 0.3 is depicted with a gray-dashed line.
Fractalfract 07 00508 g001
Figure 2. Performance of the estimator σ ^ given by (21) for different sizes of the partition under the assumption that H is known. The real value σ = 1 is depicted with a gray-dashed line.
Figure 2. Performance of the estimator σ ^ given by (21) for different sizes of the partition under the assumption that H is known. The real value σ = 1 is depicted with a gray-dashed line.
Fractalfract 07 00508 g002
Figure 3. Performance of the estimator σ ^ given by (21) for different sizes of the partition; H is unknown and estimated using (19). The real value σ = 1 is depicted with a gray-dashed line.
Figure 3. Performance of the estimator σ ^ given by (21) for different sizes of the partition; H is unknown and estimated using (19). The real value σ = 1 is depicted with a gray-dashed line.
Fractalfract 07 00508 g003
Figure 4. Performance of the estimator a ^ ( T ) for different time horizons T. The real value a = 2 is depicted with a gray-dashed line.
Figure 4. Performance of the estimator a ^ ( T ) for different time horizons T. The real value a = 2 is depicted with a gray-dashed line.
Fractalfract 07 00508 g004
Table 1. Performance of the estimator H ^ given by (19) for different sizes of the partition. The real value is H = 0.3 .
Table 1. Performance of the estimator H ^ given by (19) for different sizes of the partition. The real value is H = 0.3 .
n = 100 n = 1000 n = 10,000n = 100,000
Mean0.30746120.30187240.2996840.2999879
Variance0.0076087590.00069797470.000074383510.000007596678
CV0.28370470.087517810.028778940.009187727
Table 2. Performance of the estimator σ ^ given by (21) for different sizes of the partition under assumption that H is known. The real value is σ = 1 .
Table 2. Performance of the estimator σ ^ given by (21) for different sizes of the partition under assumption that H is known. The real value is σ = 1 .
n = 100 n = 1000 n = 10,000n = 100,000
Mean0.99646490.99997650.99966660.9999794
Variance0.0058959530.00060225940.000055297350.000005777412
CV0.077057520.024541550.0074386990.002403674
Table 3. Performance of the estimator σ ^ given by (21) for different sizes of the partition; H is unknown and estimated using (19). The real value is σ = 1 .
Table 3. Performance of the estimator σ ^ given by (21) for different sizes of the partition; H is unknown and estimated using (19). The real value is σ = 1 .
n = 100 n = 1000 n = 10,000n = 100,000
Mean1.0989721.0215490.99912511.001703
Variance0.17836770.035010520.0061817740.000860674
CV0.38430090.1831640.07869310.02928738
Table 4. Performance of the estimator a ^ ( T ) for different time horizons T. The real value is a = 2 .
Table 4. Performance of the estimator a ^ ( T ) for different time horizons T. The real value is a = 2 .
T = 1 T = 10 T = 100 T = 1000
Mean3.8726422.4096922.1093892.029289
Variance4.090640.59142940.17497050.06310528
CV0.52226180.31914630.19830140.1237909
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

Mishura, Y.; Yurchenko-Tytarenko, A. Parameter Estimation in Rough Bessel Model. Fractal Fract. 2023, 7, 508. https://doi.org/10.3390/fractalfract7070508

AMA Style

Mishura Y, Yurchenko-Tytarenko A. Parameter Estimation in Rough Bessel Model. Fractal and Fractional. 2023; 7(7):508. https://doi.org/10.3390/fractalfract7070508

Chicago/Turabian Style

Mishura, Yuliya, and Anton Yurchenko-Tytarenko. 2023. "Parameter Estimation in Rough Bessel Model" Fractal and Fractional 7, no. 7: 508. https://doi.org/10.3390/fractalfract7070508

APA Style

Mishura, Y., & Yurchenko-Tytarenko, A. (2023). Parameter Estimation in Rough Bessel Model. Fractal and Fractional, 7(7), 508. https://doi.org/10.3390/fractalfract7070508

Article Metrics

Back to TopTop