1. Introduction and Preliminaries
Parameter estimation in diffusion processes based on discrete observations is the recent trend of investigation in financial econometrics and mathematical biology since the data available in finance and biology are high-frequency discrete, though the model is continuous. For a treatise on this subject, see Bishwal (2008, 2021) [
1,
2].
Consider the Itô stochastic differential equation
where
is a one-dimensional standard Wiener process,
,
is a compact subset of ℝ,
f is a known real valued function defined on
, the unknown parameter
is to be estimated on the basis of observation of the process
. Let
be the true value of the parameter that is in the interior of
. We assume that the process
is observed at
with
and
for some fixed real number
. We estimate
from the observations
.
The conditional least squares estimator (CLSE) of
is defined as
This estimator was first studied by Dorogovcev (1976) [
3], who obtained its weak consistency under some regularity conditions as
and
. Kasonga (1988) [
4] obtained the strong consistency of the CLSE under some regularity conditions as
assuming that
for some fixed real number
. Prakasa Rao (1983) [
5] obtained asymptotic normality of the CLSE as
and
.
Florens-Zmirou (1989) [
6] studied the minimum contrast estimator, based on a Euler–Maruyama-type first-order approximate discrete time scheme of the SDE (
1), which is given by
The log-likelihood function of
is given by
where
C is a constant independent of
. A contrast for the estimation of
is derived from the above log-likelihood by substituting
with
. The resulting contrast is
and the resulting minimum contrast estimator, called the Euler–Maruyama estimator, is given by
Florens-Zmirou (1989) [
6] showed the
-consistency of the estimator as
and
and asymptotic normality as
and
.
Notice that the contrast
would be the log-likelihood of
if the transition probability was
. This led Kessler (1997) [
7] to consider Gaussian approximation of the transition density. The most natural one is achieved through choosing its mean and variance to be the mean and variance of the transition density. Thus, the transition density is approximated by
, which produces the contrast
Since the transition density is unknown, in general, there is no closed-form expression for
. Using the stochastic Taylor formula obtained in Florens-Zmirou (1989) [
6], he obtained a closed-form expression of
The contrast
is an example of such an approximation when
.
The resulting minimum contrast estimator, which is also the quasi-maximum likelihood estimator (QMLE), is given by
Kessler (1997) [
7] showed the
-consistency of the estimator as
and
and asymptotic normality as
and
for an arbitrary integer
p.
Denote
which is the mean function of the transition probability distribution. Hence, the contrast is given by
If continuous observation of
on the interval
were available, then the likelihood function of
would be
(see Liptser and Shiryayev (1977) [
8]). Since we have discrete data, we have to approximate the likelihood to obtain the MLE. Taking Itô-type approximation of the stochastic integral and rectangle rule approximation of the ordinary integral in (
9), we obtain the approximate likelihood function
The Itô approximate maximum likelihood estimate (IAMLE) based on
is defined as
Weak consistency and asymptotic normality of this estimator were obtained by Yoshida (1992) [
9] as
and
.
Note that the CLSE, the Euler–Maruyama estimator and the IAMLE are the same estimator (see Shoji (1997) [
10]). For the Ornstein–Uhlenbeck process, Bishwal and Bose (2001) [
11] studied the rates of weak convergence of approximate maximum likelihood estimators, which are of conditional least squares type. For the Ornstein–Uhlenbeck process, Bishwal (2010) [
12] studied the uniform rate of weak convergence for the minimum contrast estimator, which has a close connection to the Stratonovich–Milstein scheme. Bishwal (2009) [
13] studied Berry–Esseen inequalities for conditional least squares estimator in discretely observed nonlinear diffusions. Bishwal (2009) [
14] studied the Stratonovich-based approximate M-estimator of discretely sampled nonlinear diffusions. Bishwal (2011) [
15] studied Milstein approximation of the posterior density of diffusions. Bishwal (2010) [
16] studied conditional least squares estimation in nonlinear diffusion processes based on Poisson sampling. Bishwal (2011) [
17] obtained some new estimators of integrated volatility using the stochastic Taylor-type schemes, which could be useful for option pricing in stochastic volatility models; see also Bishwal (2021) [
2].
Prime denotes the derivative with respect to
, dot denotes the derivative with respect to
x and ⋁ denotes the max symbol throughout the paper. In order to obtain a better estimator in terms of lowering variance in Monte Carlo simulation, which may have a faster rate of convergence, first, we use the algorithm proposed in Bishwal (2008) [
1]. Note that the Itô integral and the Fisk–Stratonovich (FS, henceforth; Fisk, while introducing the concept of quasimartingale, had the trapezoidal approximation and Stratonovich had the midpoint approximation, converging to the same mean square limit) integral are connected by
where o is the Itô’s circle for the FS integral. We transform the Itô integral (the limit of the rectangular approximation to preserve the martingale property) in (
9) to the FS integral and apply FS-type trapezoidal approximation of the stochastic integral and rectangular rule-type approximation of the Lebesgue integrals and obtain the approximate likelihood
The Fisk–Stratonovich approximate maximum likelihood estimator (FSAMLE) based on
is defined as
Weak consistency as
and
and asymptotic normality as
and
of the FSAMLE were shown in Bishwal (2008) [
1]. Berry–Esseen bounds for the IAMLE and the FSAMLE for the Ornstein–Uhlenbeck processes were obtained in Bishwal and Bose (2001) [
11].
We shall use the following notations: , , C is a generic constant independent of and other variables (it may depend on ). Throughout the paper, denotes the derivative with respect to x and denotes the derivative with respect to of the function . Suppose that denotes the true value of the parameter and . We assume the following conditions:
Assumption 1.
(A1) ,
.
(A2) for all
where for any integer r.
(A3) The diffusion process X is stationary and ergodic with invariant measure ν, i.e., for any g with (A4) for all .
(A5)
(A6) f is continuously differentiable function in x up to order p for all θ.
(A7) and all its derivatives are three times continuously differentiable with respect to θ for all . Moreover, these derivatives up to third order with respect to θ are of polynomial growth in x uniformly in θ.
The Fisher information is given by and for any , or any compact , (A8) The Malliavin covariance of the process is nondegenerate.
The Malliavin covariance matrix of a smooth random variable S is defined as , where is the Malliavin derivative. The Malliavin covariance is nondegenerate if is almost surely positive and, for any , one has This, associated with the functional , is given by where and , respectively, satisfy In the case of independent observations, in order to prove the validity of asymptotic expansion, one usually needs a certain regularity condition for the underlying distribution, such as the Cramér condition; see Bhattacharya and Ranga Rao (1976) [
18]. This type of condition then ensures the regularity of the distribution and hence the smoothness assumption of the functional under the expectation whose martingale expansion is desired can be removed. This type of condition for dependent observations leads to the regularity of the distribution of a functional with nondegenerate Malliavin covariance, which is known in Malliavin calculus; see Ikeda and Watanabe (1989) [
19] and Nualart (1995) [
20]. Malliavin covariance is connected to the Hörmander condition, which is a sufficient condition for a second-order differential operator to be hypoelliptic; see Bally (1991) [
21]. For operators with analytic coefficients, this condition turns out to be also necessary, but this is not true for general smooth coefficients.
More precisely, let
X be a differentiable ℝ-valued Wiener functional defined on a Wiener space. Assume that there exists a functional
such that
Thus, it is a regularity condition of the characteristic function, which is a consequence of the nondegeneracy of the Malliavin covariance in the case of Wiener functionals. The functional
, which is a random variable satisfying
, is a truncation functional extracting from the Wiener space, the portion on which the distribution is regular. If
X is almost regular, one may take
nearly equal to one. Uniform degeneracy of the Malliavin covariance of the functional
can be shown under (A8); see Yoshida (1997) [
22].
Bishwal (2009) [
13] obtained the rate of convergence to normality of the Itô AMLE and the Fisk–Stratonovich AMLE of the order
and
, respectively, under the regularity conditions given above with
for (A4). We obtain the rate of convergence to normality, i.e., Berry–Esseen bound of the order
for the QMLE
for arbitrary integer
p.
We need the following lemma from Michel and Pfanzagl (1971) [
23] to prove our main results.
Lemma 1. Let and η be any three random variables on a probability space with . Then, for any , we have 2. Main Results
We start with some preliminary lemmas. Let
L denote the generator of the diffusion process,
The k-th iterate of L is denoted as . Its domain is . We set .
Stochastic Taylor formula (Kloeden and Platen (1992) [
24]): For a
times continuously differentiable function
, we have for
and
Lemma 2. With , the stochastic Taylor expansion of is given bywhere R denotes a function for which there exists a constant C such that Proof. Applying the stochastic Taylor formula of Florens-Zmirou (1989, Lemma 1) [
6], one obtains the result. See also Kloeden and Platen (1992) [
24].
Consider the following special cases:
Euler Scheme: For ,
Milstein Scheme: For ,
Simpson Scheme: For ,
Boole Scheme: For ,
□
Remark 1. For , This produces the CLSE. This estimator has been very well studied in the literature (see Shoji (1997) [10]). Remark 2. Note that the Milstein scheme is equivalent to Stratonovich approximation of the stochastic integral after converting the Itô integral to the Stratonovich integral.
Proof. First, we show that, for
,
We emphasize that the Itô formula is a stochastic Taylor formula of order 2. By the Itô formula, we have
where
We employ Taylor expansion in the local neighborhood of
. Let
. Then, we have
where
and
. Further
By Lemma 2, we have
Further
Hence
Observe that, with
we have
On the other hand, with
we have
However,
using (A4) and (A3). On substitution, the last term is dominated by
By the same method, we have
Thus, the proof for
is complete. Next, we consider the general case
. Denote
Observe that, by Lemma 2, we have
Thus, by combining the bounds for
,
and
, we have
□
The following lemma is from Bishwal (2008) [
1].
Lemma 4. Then, under the conditions (A1)–(A8), The following lemma follows from Theorem 7 in Yoshida (1997) [
22].
Lemma 5. Then, under the conditions (A1)–(A8), Our main result is the following theorem.
Theorem 1. Under the conditions (A1)-(A8), for any , we have Proof. We start with
and
Let
By Taylor expansion, we have
where
. Since
, hence we have
However,
from Lemma 4 (see also Pardoux and Veretennikov (2001) [
25] and Yoshida (2011) [
26]). It can be shown that
(see Altmeyer and Chorowski (2018) [
27]). Hence
Further, by Lemma 1 (b), we have
since, by Lemmas 1 (a) and 5, we have
Choosing , we have the result.
On the other hand, by Taylor expansion, we have
where
Since
, hence we have
Let
in
as
and
. Similar to Lemma 4, it can be shown that
(see also Pardoux and Veretennikov (2001) [
25] and Yoshida (2011) [
26]). It can be shown that
(see Altmeyer and Chorowski (2018) [
27]). Hence
Thus, by Lemma 1 (b), we have
since, by Lemmas 1 (a) and 5, we have
Choosing , we have the result.
Now, we study the general case for arbitrary
p. By Taylor expansion, we have
where
. Since
, hence we have
Let
in
as
and
. Similar to Lemma 4, it can be shown that
(see also Pardoux and Veretennikov (2001) [
25] and Yoshida (2011) [
26]). It can be shown that
(see Altmeyer and Chorowski (2018) [
27]). Hence
Further, by Lemma 1 (b), we have
since, by Lemmas 1 (a) and 5, we have
Choosing , we have the result. □
Remark 3. With , for the Euler scheme, which produces the conditional least squares estimator, one obtains the rate With , for the Milstein scheme, one obtains the rate With , for the Simpson scheme, one obtains the rate With , for the Boole scheme, one obtains the rate Thus, the higher the p, the sharper the bound. Thus, the Itô/Euler scheme gives the first-order QMLE, the Milstein/Stratonovich scheme produces the second-order QMLE, the Simpson scheme produces the fourth-order QMLE and the Boole scheme produces the sixth-order QMLE. See Bishwal (2011) [28] for a connection of this area to the stochastic moment problem and hedging of generalized Black–Scholes options.