1. Introduction
The mathematical literature has extensively analyzed stochastic differential equations (SDEs) driven by fractional Brownian motion. Most of these efforts have been motivated by problems arising in the financial applications of SDEs, such as option pricing, stochastic volatility, and interest rate modeling. However, there are few results concerning SDEs with boundary conditions. Typically, only SDEs involving reflection at the boundary are considered (see [
1,
2]). Here, we consider a new class of SDEs with stochastic forcing. This class allows us to consider boundaries of a new type.
We consider stochastic differential equations of the following form:
where
is a continuous function,
are measurable functions, and
is a fractional Brownian motion. The stochastic integral in Equation (
1) is a pathwise generalized Lebesgue–Stieltjes integral. Thus, we can use the pathwise approach to consider these fractional stochastic differential equations (FSDEs). We call Equation (
1) the FSDE equation with stochastic forcing term
. Examining such a model can be interpreted as studying the environment’s influence on the behavior of a process. Such equations can be used to consider FSDEs with a permeable wall. The permeable wall model describes a process that can cross the wall, but where the force does not allow the process to move far from the wall. In [
3,
4], the fractional Vasicek process with soft wall was considered as a modeling example. This example explains what a fractional SDE with a permeable wall is. These types of processes can be applied in the natural sciences. In particular, such processes can be used in financial mathematics as models for stochastic volatility. Indeed, it has recently been irrefutably proven that financial markets have a memory that is best interpreted in the framework of stochastic volatility. A stochastic differential equation involving fractional Brownian motion is a natural model with a memory [
5,
6,
7]. On the other hand, volatility should have certain limits and reasonable sizes, and should not deviate infinitely far; otherwise, such a market model would not allow equivalent Martingale measures and would poorly describe real financial processes. Thus, the presence of a permeable wall allows us to construct a model of stochastic volatility with reasonable behavior.
In general, SDEs rarely possess closed analytic-form solutions; therefore, both in general and in our case, it is important to consider certain numerical methods for their solution. The existence and uniqueness of the solution of Equation (
1) was obtained in [
8]. A special case of Equation (
1) with constant and strictly positive diffusion coefficients was considered in [
3]. In the article, we are interested in pathwise numerical approximations of the solution to Equation (
1).
Much of the literature is devoted to numerical methods for SDEs driven by fBm or a combination of Brownian motion and fBm. Strong SDE approximation schemes are usually considered in the literature. Euler, modified Euler, and other higher-order approximation schemes should be mentioned here (see, e.g., [
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24,
25,
26] and references therein).
The rate of convergence for Euler approximations
of solutions of pathwise SDEs driven by fBm with Hurst index
was obtained in [
17]. It was proved that for any natural number
there exists a random variable
such that it is almost certain that
Here, we apply the implicit Euler- and implicit Milstein-type approximations to the solution of Equation (
1) and find the pathwise convergence rate. These results were obtained for the first time under fairly general coefficient restrictions. For simplicity, we consider an implicit Milstein-type approximation for the time-homogeneous Equation (
1).
The paper is organized in the following way. In
Section 2, we present the paper’s main results.
Section 3 contains definitions of considered spaces of functions and a priori estimates for the Lebesgue-Stieltjes integral.
Section 4 defines a deterministic differential equation corresponding to FSDE (
1) and considers its implicit Euler approximation properties. Some results are taken from [
8]. In
Section 5, we obtain a convergence rate for implicit Euler approximation for a deterministic differential equation corresponding to FSDE (
1).
Section 6 presents the implicit Milstein-type approximation and auxiliary results. In
Section 7, the convergence rate of the Milstein-type approximation is obtained. Finally,
Section 8 considers the fractional Pearson diffusion process as an example.
2. Main Result
We assume that the coefficients satisfy the following conditions with some nonrandom constants:
is differentiable in x, and there exist some constants ; moreover, for every there exists such that the following properties hold:
(i) Lipschitz continuity in
x:
(ii) Local Hölder continuity of the derivative in
x:
(iii) Hölder continuity in
t:
There exists a constant , and for every there exists such that the following properties hold:
(i) Local Lipschitz continuity in
x:
(ii) Linear growth condition:
(iii) Hölder continuity in
t:
The function is differentiable and there exist some constants , ’ moreover, for every there exists such that the following properties hold:
(i) for all .
(ii) Local Hölder continuity of the following derivative:
The function , where , has the following properties:
(i) It is strictly monotonic and surjective.
(ii) There is a constant
such that
Remark 1 (see Remark 8 in [
3])
. Under Assumption , the function D satisfies Assumption with . For the time-homogeneous version of Equation (
1), we assume that the coefficients
satisfy the following conditions:
There exist constants
such that the following properties hold:
where we write
instead of
to shorten notation.
Remark 2. Linear growth conditions for the functions f, g, and are unnecessary, but simplify the future notation.
For simplicity of presentation, we consider uniform partitions of the interval
. Let
be a sequence of uniform partitions of the interval
and let
,
,
. We define the time-continuous interpolation of the implicit Euler approximation for the Equation (
1) as
and the time-continuous interpolation of the implicit Milstein-type approximation for the time-homogeneous Equation (
1) as
where
and
if
,
We introduce the symbol for simplicity of notation. Let be a sequence of r.v.s, let be an a.s. nonnegative r.v., and let be a vanishing sequence. Then, means that for all n.
Set
and denote by
and
the implicit Euler- and Milstein-type approximations. The norm
is defined in
Section 3.1.
Theorem 1 (See Theorem 1 [
8])
. Suppose that the functions and satisfy Assumptions and with , . If Assumption is satisfied and , then there exists a unique stochastic process satisfying FSDE (1), where is the space of γ-Hölder continuous functions. Theorem 2. Under the hypotheses of Theorem 1 with replaced by , we have Theorem 3. Suppose that the functions and satisfy Assumption with . If Assumption is satisfied and , then there exists a unique stochastic process and The statements of Theorems 2 and 3 follow directly from the results for deterministic differential equations, as we can apply the pathwise approach for FSDE (1).
3. Preliminaries
3.1. Spaces of Functions and Norms
Let us recall some functional spaces that are used in the future.
We use
, where
, to denote the space of real-valued measurable functions
, meaning that we have
The space
is a Banach space with respect to the norm
; for
, the equivalent norm is defined by
For any
, we denote by
the space of
-Hölder continuous functions
equipped with a norm
, where we have
Clearly, we have
for
and
We denote by
,
the space of measurable functions
, meaning that we have
such that
(see [
21]).
In addition, we denote by
,
the space of measurable functions
f on
, such that we have
Fixing
and letting
denote a set of all possible partitions of
, for any
we define the following:
Recall that is called the p-variation of f on . We denote by (resp. ) the class of (resp., continuous) functions on with bounded p-variation, .
We define , which is a seminorm on ; in addition, is 0 if and only if f is constant. For each f, is a non-increasing function of , i.e., if , then . Thus, if . If , then f is bounded.
Let and . Then, is a norm, and equipped with the p-variation norm is a Banach space.
3.2. Riemann–Stieltjes Integral
Assume that
and
, where
. The generalized Lebesgue–Stieltjes integral (see [
21])
exists for all
and for any
where
is the Weyl derivative, and
is a Gamma function. Furthermore, the integral
exists if
.
If
and
with
, then the generalized Lebesgue–Stieltjes integral exists and coincides with the Riemann–Stieltjes integral (see [
22]).
From Young’s Stieltjes integrability theorem [
23] (p. 264), the Riemann–Stieltjes integral
can be defined for functions having bounded
p-variation on
(see [
24]).
Let
and
with
,
,
. If
f and
h have no common discontinuities, then the extended Riemann–Stieltjes integral
exists and the Love–Young inequality
holds for any
, where
,
denotes the Riemann zeta function, i.e.,
.
Proposition 1 (Chain rule [
22] (comment on Theorem 4.3.1))
. Let and be real-valued functions such that with . Then, for any ,where and are the partial derivatives of F with respect to the first and second variables, respectively. 3.3. Estimation of the Generalized Lebesgue–Stieltjes Integrals
From now on, we fix
. For any function
, we define
where
f satisfies Assumptions
.
Proposition 2 (See Proposition 4.4 [
21])
. If , then .If are such that and , thenfor all , where , , from . Given two functions
and
, we denote
where
g satisfies Assumption
with constant
.
Proposition 3 (See Proposition 4.1 [
21])
. Let ; then, the following estimates hold for :andwhere , is the Beta function. Proposition 4 (See Proposition 4.2 [
21])
. If , then .If are such that and , thenfor all , whereand Remark 3. If and , then 4. The Implicit Euler Approximation and Auxiliary Results
Let
. Recall that almost all trajectories of fBm
belong to
. Instead of considering Equation (
1), we consider the deterministic differential equation on
:
where
,
,
.
Theorem 4. Suppose that the functions and satisfy Assumptions and with , . If Assumption is satisfied, then Equation (17) has a unique solution , where , is defined in (5). Proof. The theorem statement follows directly from Theorem 1. Set . It is sufficient to note that if . □
We define the implicit Euler approximations for Equation (
17) as
and their continuous interpolations as
where
and
if
,
and where
. For abbreviation, let
stand for
.
We can rewrite the implicit Euler approximations (
18) and (19) in a more compact way:
with
and
where
The implicit Euler approximation scheme (
20) is correctly defined. From the recursive expression (
20), we calculate
. The properties of the function
provide a single value of
. Because
is a continuous function,
is a continuous function. Indeed, because
and
are continuous functions,
is a continuous function.
The following properties hold for the implicit Euler approximation:
Proposition 5 (See Propositions 4 and 5 [
8])
. Under the assumptions of Theorem 4, we obtain 5. Rate of Convergence of the Implicit Euler Approximation
Lemma 1 (See Lemma 7.1 in [
21]; see also Lemma 3 [
8])
. Let Φ be function satisfying Assumption . Then, for all and , , , : Theorem 5. Under the hypotheses of Theorem 1 with replaced by , we havewhere and where x is the solution of Equation (17). Proof. We denote
. Because
x is an element of the space
and because
, there exists
N such that
and
for all
n. Furthermore,
and
,
(see Propositions 2 and 4), and
. From Lemma 1 in [
8] we have
,
for any fixed
. □
Recall that elements of the space
have the finite norm
with
. Thus,
Now, we can evaluate the terms on the right-hand side of (
23). If
, then the estimate
follows from Lemma 1 and the arguments used to prove the uniqueness of the solution in [
8]. The estimates of the second and fourth terms follow from Propositions 2 and 4 if
. Because
, we have
and
and our restrictions for
and
are satisfied. Thus,
where
For any
, we can choose a sufficiently large
such that
Thus,
To complete the estimate of
, it remains to estimate the right-hand side of the above inequality. From (
9) it follows that instead of the norm
it is sufficient to estimate the norm
.
We first estimate the norm
. Combining Assumptions
–
, we obtain
and
where
Thus,
Because
and
, it is the case that
and
Moreover, from (
9) and the inequality
, which is valid for all
, we obtain
Now, we can go back to the estimate of the norm
. Because
, we have
(see Proposition 4.2 in [
21]). From Lemma 2 in [
8], it follows that
for any fixed
if
for any fixed
. Note that
.
From the above and Proposition 3, it follows that
Applying (
27), we have
and
To estimate the second term in (
29), we note that
and (see Proposition 5 in [
8])
Consequently,
It is easy to check that
as
for
and
Thus,
Therefore,
and
as
.
6. The Implicit Milstein Type Approximation and Auxiliary Results
For a given partition
, we define the implicit Milstein-type approximations for the time-homogenous equation
as
where
,
,
,
.
Theorem 6. Let and let the functions and satisfy Assumptions and , where is defined in (7). Then, Equation (33) has a unique solution , where . Proof. From Assumption , we have . Because , it follows that . Consequently, . Therefore, the conditions of Theorem 4 are satisfied and the theorem’s statement holds. □
Applying the chain rule, we can rewrite (
34) as
The continuous-time interpolation of the Milstein scheme is defined by
where
Because
, we have
and
where
. From now on, we assume that
.
The method of proving the convergence of the implicit Milstein approximation to the solution of Equation (
1) repeats the idea of the proof for the implicit Euler approximation.
Lemma 2. Let Assumption be satisfied. For any fixed , the functions , , , and belong to .
Proof. From Lemma 1 in [
8], we have
,
. Indeed, the proof does not change when the Euler approximation is replaced with the Milstein one. Now, we can consider
.
We first note that the function
has a bounded variation on
for any fixed
n; thus, it is bounded and has
p-bounded variation. Because
, it is the case that
To simplify the notation, we write
instead of
.
Now, it remains to prove that
for fixed
. Assume that
for some
and
. Then,
as the chain rule implies an inequality
If
, then from (
38), the Love–Young inequality, and (
36) it follows that
The boundedness of the last term in the above inequality follows from (
37). Consequently,
.
From Assumption
, it follows that
Thus,
for any fixed
. □
The next lemma allows us to apply the estimate (
11) to the integral
.
Lemma 3. Let Assumptions and be satisfied. If for any fixed , then for any fixed .
Proof. It follows from Lemma 2 that
for any fixed
and that there exists a constant
depending on
n such that
First, we note that
for any fixed
(see Lemma 2 in [
8]).
Now, we prove that
for any fixed
. It is clear that
and
Thus,
and
as
and
The above inequality was proved in Lemma 2 in [
8].
Consequently, it follows that for any fixed . □
Boundedness of the Norm
Proposition 6. Let and the functions and satisfy Assumptions and . Then, there exists a constant C such that we have the following: Proof. It is easy to check (see Proposition 4 in [
8]) that
where the constant
c is taken from Assumption
. From Lemma 2, we know that the norms mentioned above are finite for any fixed
.
To obtain the statement of the proposition, we repeat the proof of Proposition 4 in [
8]. First, we note that from Assumption
it follows that
where
.
Based on the above and the proof of Proposition 4 in [
8], we can use the results obtained in Proposition 4. Thus, we obtain
with certain constants
,
.
Now, we estimate
. From Lemma 3, it follows that we can apply Proposition 3. Thus, we obtain the inequality
From inequalities (
40) and (
41) and the inequality
we have
Thus,
where
.
Now, let us move on to estimating the norm of . We divide the first integral into two parts and estimate each separately.
Let
. Applying change the order of integration, noting that
(as in [
18] (p. 349)), we obtain
By changing the order of integration in the first part of the integral (as in [
18], p. 3497), we obtain
Note that for the second term in (
49) we have
where
For the third term, it is evident that
The estimation of the fourth term was proved in [
8], and we repeat it below:
where
with
as defined in (
44).
Consequently, for certain constants
,
, we obtain
Obviously, from (
43), (
45), (46), and (
50), we have
Note that for
, we have
Thus,
and from Lemma 7.6 in [
21] it follows that
where
and
are positive constants depending only on
:
□
Now, we can strengthen the result of Lemma 2.
Proposition 7. Under the assumptions of Proposition 6, we obtain .
Proof. Recall that from Lemma 2 we have
,
,
,
for any fixed
. Thus, for any fixed
, we have the following:
The proof repeats the arguments of the proof of Proposition 5 in [
8]. The terms
,
, are bounded for all
n. This follows from Proposition 6 and the proof of Proposition 5 in [
8].
The boundedness of the norm
can be proved in much the same way as was done for the norm
. From (
11), it follows that
Applying (
40), we obtain
From (
41), (48), and (
31), it is obvious that
Thus, the norm
is bounded for all
n and the proof is complete. □
7. Rate of Convergence of the Implicit Milstein-Type Approximation
Theorem 7. Under conditions of Theorem 6,where and x is the solution of Equation (33). The proof is similar in spirit to the proof of the rate of convergence of the implicit Euler approximation. For abbreviation, let stand for .
Because
x and
are elements of the space
, there exists
N such that
and
for all
n. It is obvious that
We divide the proof of the convergence rate of the norm
into two steps. We estimate the first, second, and fourth terms in the first step. The estimation of the first term is provided in (
24). Estimates for the second and fourth terms follow from Propositions 2 and 4.
Similarly, as in
Section 4, for sufficiently large
and
b satisfying inequality (
25), we obtain
In the second step, we estimate the right-hand side of the above inequality. Estimation of the first term follows immediately from (
28). It follows from (
9) that to estimate the norm of the second term it is sufficient to estimate the norm of
.
From Proposition 6, Proposition 4.2 in [
21], and Lemma 3, we obtain
for any fixed
. Proposition 3 shows that
Assume that
. First, observe that
From Assumption
and (
44), we obtain
Further,
and
From (
53)–(
55) and the fact that
, we can conclude that there exists a constant
C independent of
n such that
Set
. For the first term in (
52), we obtain
We can rewrite the second integral (
52) as the sum of two integrals:
and evaluate each of them separately.
Applying (
56) and (
31), we obtain
Note that
and that for
we have
We can divide the integral
into two parts and estimate each separately. From (
56), it is evident that
From Assumption
and (
44), we obtain
Thus,
Consequently,
and
The statement of the theorem follows from
8. Example: Fractional Pearson Diffusion with a Stochastic Force
Consider the Pearson diffusion process with a stochastic force
where
and the function
satisfies Assumption
. Assume that the coefficients
,
are such that
and
. Then,
.
For the existence of a unique solution to problem (
57), it is necessary to check the conditions of Theorem 1. Note that
where
is a critical point of the function
.
Straightforward computation shows that
and
Thus, the Pearson diffusion process with a stochastic force has a unique solution under the above conditions.
Note that
Thus, Assumption
is satisfied, and for an implicit Milstein-type approximation we have the following:
where the rate of convergence is
. The convergence rate for Euler approximation is
.