Next Article in Journal
An Unconditional Positivity-Preserving Difference Scheme for Models of Cancer Migration and Invasion
Next Article in Special Issue
A Cuboid Registers Topic, Activity and Competency Data to Exude Feedforward and Continuous Assessment of Competencies
Previous Article in Journal
Coefficient Estimates and the Fekete–Szegö Problem for New Classes of m-Fold Symmetric Bi-Univalent Functions
Previous Article in Special Issue
Resolutions of the Jerk and Snap Vectors for a Quasi Curve in Euclidean 3-Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Additive Noise Effects on the Stabilization of Fractional-Space Diffusion Equation Solutions

1
Department of Mathematics, Faculty of Science, University of Ha’il, Ha’il 2440, Saudi Arabia
2
Department of Mathematics, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
3
Department of Mathematics, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
*
Author to whom correspondence should be addressed.
Mathematics 2022, 10(1), 130; https://doi.org/10.3390/math10010130
Submission received: 9 December 2021 / Revised: 30 December 2021 / Accepted: 31 December 2021 / Published: 2 January 2022
(This article belongs to the Special Issue Advanced Methods in Computational Mathematical Physics)

Abstract

:
This paper considers a class of stochastic fractional-space diffusion equations with polynomials. We establish a limiting equation that specifies the critical dynamics in a rigorous way. After this, we use the limiting equation, which is an ordinary differential equation, to approximate the solution of the stochastic fractional-space diffusion equation. This equation has never been studied before using a combination of additive noise and fractional-space, therefore we generalize some previously obtained results as special cases. Furthermore, we use Fisher’s and Ginzburg–Landau equations to illustrate our results. Finally, we look at how additive noise affects the stabilization of the solutions.

1. Introduction

Stochastic partial differential equations (SPDEs) are crucial in understanding the dynamics of many fascinating phenomena. In recent years, the significance of taking random influences into consideration in modeling, analyzing, simulating, and predicting complex phenomena has become widely realized in physics, chemistry, biology, materials science, and climate dynamics, as well as in geophysical and other areas; see [1,2].
Furthermore, fractional derivatives have drawn tremendous interest mainly because of their possible implementations in different areas, such as, for example, in physics [3,4,5,6], biology [7], finance [8,9,10], biochemistry and chemistry [11], and hydrology [12,13]. These fractional-order equations are better suited than equations with integer-orders because derivatives of the fractional order are allowed the memory and hereditary properties of various substances to be represented [14].
It seems that examining fractional equations with some random force is more significant. Therefore, we are concerned, here, with the fractional space-diffusion equation perturbed by additive noise on a bounded domain G R :
d φ = [ ε 2 D ( Δ ) r 2 φ + P ( φ ) ] d t + ε 1 d W , t 0 , x G ,
where ε 1 is a small parameter, D is the diffusion coefficient, ( Δ ) r 2 is the fractional Laplacian with r ( 1 , 2 ] , P is a polynomial with the degree m and represents reaction kinetics, and W is a finite dimensional Wiener process.
In normal diffusion with time, the mean square displacement of an equation particle linearly increases, i.e., x 2 ( t ) t . In contrast, anomalous diffusion is a diffusion process not following this linear relation. In some cases, they have a power-law scaling relation, namely x 2 ( t ) t r , that is present in various types of equations. r is defined as the anomalous exponent of diffusion in the case of r = 1 of normal diffusion, whereas r = 2 , r ( 0 , 1 ) , and r ( 1 , 2 ) correspond to a ballistic diffusion, a sub-diffusion, and a Levy super-diffusion, respectively [4]. By the transformation of a Fourier, the anomalously diffusive operator ( Δ ) r 2 is defined [8,15,16,17] as:
L { ( Δ ) r 2 φ } ( η ) = η r L { φ } ( η ) ,
where L { φ } is the Fourier transform of φ .
It is interesting to note that if we input P ( φ ) = φ ( 1 φ ) ( φ a ) , Equation (1) becomes the stochastic fractional space Fitzhugh–Nagumo equation, which is used in the field of biology and population genetics, and also is used to model nerve impulse transmission [18,19]. We derive the stochastic space-fractional heat equation if we set P ( φ ) = φ φ 3 , which is used in physics and describes the heat distribution within a given time interval in a given region [20]. Furthermore, if P ( φ ) = φ ( 1 φ N ) , then (1) gives rise to the stochastic fractional-space Fisher equation, which is used as the spatial and temporal propagation model in an infinite medium of a virile gene [21]. Additionally, it is used in chemical kinetics [22], auto catalytic chemical reactions [23], flame propagation [24], neurophysiology [25], and nuclear reactor theory [26].
Recently, Equation (1) with r = 2 was addressed in the stochastic case by [27,28,29,30]. This equation with r = 2 was studied analytically by [31,32] in the deterministic case, i.e without noise. Recently, this Equation (1) was discussed by [33,34] with multiplicative noise. Furthermore, many analytical and numerical methods have been proposed to find the solution of the fractional-space partial differential Equation (1) without noise, such as in [35,36,37,38,39,40,41]. In this paper, we analytically approximate the solution of Equation (1) by using the perturbation method. This equation has never been addressed before using a combination of additive noise and fractional-space, therefore we generalize some previously obtained results as a special case.
The first aim of this paper is to show that the approximate solution of (1) is given by
φ ( t , x ) = ξ ( t ) + χ ( t , x ) + error ,
where ξ solves
d ξ = [ P ( ξ ) + G ( ξ ) ] d t .
The polynomial G ( ξ ) , is defined later in (22) and has a degree of m 2 . The term χ ( t , x ) in Equation (2) is referred to as a fast Ornstein–Uhlenbeck process (FOUP for short) and it will be defined later in (10). We note that the ordinary differential Equation (3) contains the same polynomial P as in Equation (1), plus an additional polynomial G that occurs as a result of the interaction between the non-linear term and additive noise. The second aim of this paper is to discuss the impact of additive noise on the solutions of Equation (1).
As an applications of how our results can be applied, we provide theoretical examples from physics (the real Ginzburg–Landau equation) and biology (the Fisher’s equation). To clarify our results, let us consider the very simple real-valued Ginzburg–Landau equation with Neumann boundary conditions on [ 0 , π ] as follows
d φ = [ ε 2 ( Δ ) r 2 φ + φ φ 3 ] d t + ε 1 d W .
The approximation Theorem 2 shows us that the solution of the Ginzburg–Landau Equation (4) shall be of the kind (2), where ξ is the solution of
d ξ = [ 1 k = 1 N 3 α k 2 2 k r ξ ξ 3 ] d t ,
where α k for k = 1 , 2 , , N are real numbers and where there is noise intensity. If we input r = 2 , we have the previous result that was obtained by [28].
In this paper, one great innovation of our approach is the explicit estimation of error in terms of arbitrarily high moments of error, as usually only weak convergence is handled against approximation. Moreover, this paper is the first paper, to the best of our knowledge, to analytically find the approximate solution of stochastic fractional-space partial differential equations.
The rest of this article is set out as follows. In the next section, we present some notations, assumptions, and preliminaries that we need in this paper. We also estimate an equation representing the high modes and give bounds on it in Section 3. We will state a general case of the averaging-over OU-process in Section 4. After that, we deduce the limiting equation and prove the main result of Theorem 2 in Section 5. In Section 6, there are two examples to clarify our results, including the Ginzburg–Landau and Fisher’s equations. Finally, the conclusion of this paper is given.

2. Preliminaries

Let H = L 2 ( [ 0 , π ] ) be a separable Hilbert space with norm · and inner product · , · .
Since the operator Δ is self-adjoint, there exists a complete orthonormal system e j j = 0 and a sequence λ j j = 0 such that
Δ e j = λ j e j ,
with
0 = λ 0 λ 1 · · · · λ j · · · ·
Here, we consider Δ with the Neumann boundary condition on [ 0 , π ] . Therefore,
e j ( x ) = 1 i f j = 0 , 2 π cos ( j x ) i f j 0 ,
and
λ j = j 2 .
Define H c and H s as
H c : = kernel { Δ } = s p a n { 1 } and H s = ( H c ) .
Define the projections
P c : = H H c and P s : = I P c ,
where I is the identity operator on H .
Consider the fractional space H r to be the domain of A r 2 for r > 0 , which is defined as
( Δ ) r 2 φ = j = 0 λ j r 2 φ , e j e j ,
H r = D ( ( Δ ) r 2 ) = φ H : j = 0 λ j r φ j 2 < ,
with norm
φ r 2 = ( Δ ) r 2 φ 2 = k = 0 λ j r φ j 2 .
Furthermore, let T r ( t ) = exp ( t ( Δ ) r 2 ) for t 0 be the analytic semigroup generated by the fractional Laplacian ( Δ ) r 2 and satisfy
T r ( t ) φ r e ω t φ r for all φ H r ,
where a constant ω > 0 exists.
For the non-linear P in Equation (1), we assume:
Assumption 1.
Let P : H r H r satisfy for all φ H r such that
P ( φ ) r C ( 1 + φ r m ) ,
where m is the degree of P .
Put shortly, we are using P c ( φ ) = P c P ( φ ) and P s ( φ ) = P s P ( φ ) .
Assumption 2.
Let F ( u ) = P c ( u ) + G ( u ) , where G ( u ) is defined in (22). Assume for u , w , v H c that
F ( u + v + w ) F ( v ) , u C ( u 2 + w 2 + w m ) .
We note that from Assumption 2, if we input v = w = 0 , , we derive
F ( u ) , u C u 2 for u H c
For the noise in Equation (1), see the following.
Assumption 3.
Suppose that the Wiener process W ( t ) for t 0 , is finite dimensional and acts only on H s . Corresponding to [42], one can write it as
W ( t , x ) = j = 1 N α j β j ( t ) e j ( x ) ,
where α j R for all j { 1 , 2 , , N } and ( β j ) j { 1 , 2 , , N } are mutually independent real-valued Brownian motions.
Definition 1.
Define the FOUP χ as
χ ( t ) = j = 1 N χ j ( t ) e j ,
where
χ j ( t ) = α j ε 1 0 t e ε 2 D λ j r 2 ( t s ) d β j ( s ) .
In the following definition, we assume that the solution of Equation (1) is not too large.
Definition 2.
Stopping time: define the stopping time τ * as
τ * : = inf t > 0 : φ ( t ) r > ε κ T 0 ,
for some T 0 > 0 and κ ( 0 , 1 2 m ) .

3. High Modes and Its Bounds

In this section, we deduce an equation representing high modes and bound it. We start by splitting the solution φ of (1) into
φ ( t , x ) = φ c ( t ) + φ s ( t , x ) ,
where φ c H c and φ s H s . By substituting (12) into (1), we have
d ( φ c + φ s ) = [ ε 2 D ( Δ ) r 2 φ s + P ( φ c + φ s ) ] d t + ε 1 d W .
By projecting to H s , we obtain
d φ s = [ ε 2 D ( Δ ) r 2 φ s + P s ( φ c + φ s ) ] d t + ε 1 α d W .
This equation can be expressed in the integral form as
φ s ( t ) = T r ( ε 2 D t ) φ s ( 0 ) + 0 t T r ( ε 2 D ( t τ ) ) ( φ c + φ s ) d τ + χ ( t ) ,
where χ is defined in Definition 1.
In the following lemma, we will show how φ s ( t ) equals to χ ( t ) , plus a small term.
Lemma 1.
Assume that Assumption 1 is satisfied. Then, there is C > 0 such that
E sup t 0 , τ * φ s ( t ) χ ( t ) T r ( ε 2 D t ) φ s ( 0 ) r p C ε p m p κ ,
for p 1 and κ > 0 from the definition of τ * .
Proof. 
Using the triangle inequality for (15), the equation yields
φ s ( t ) χ ( t ) T r ( ε 2 D t ) φ s ( 0 ) r 0 t T r ( ε 2 D ( t s ) ) P s ( φ c + φ s ) d s r C P s ( φ c + φ s ) r 0 t e ε 2 ω D ( t s ) d s C ε 2 ( 1 + φ c + φ s r m ) .
By taking E sup t 0 , τ * on both sides, we find that
E sup t 0 , τ * φ s ( t ) χ ( t ) T r ( ε 2 D t ) φ s ( 0 ) r C ε 2 ( 1 + E sup t 0 , τ * φ c + φ s r m ) C ε 2 ,
where we used (5), representing Assumption 1 and the definition of τ * , respectively. □
Now, let us, without proof, declare the uniform bounds on χ ( t ) . For the proof, see Lemma 4.2 in [28].
Lemma 2.
Let χ ( t ) be defined in Definition 1. Then, there is C > 0 such that
E sup t 0 , T 0 χ ( t ) r p C ε κ 0 ,
for every p 1 and κ 0 > 0 .
The next corollary declares that φ s ( t ) is much smaller than ε κ , as stated in the definition of stopping time τ * .
Corollary 1.
Assume that the assumptions of Lemmas 1 and 2 are satisfied. Let φ s ( 0 ) = O ( 1 ) . Then, for p 1 , κ < 1 m and C > 0 ,
E sup t 0 , τ * φ s ( t ) r p C ε κ 0 for p 1 .
Proof. 
Using Equation (16) and triangle inequality, we find that
E sup 0 , τ * φ s r p c E sup 0 , τ * φ s χ T r ( ε 2 D t ) φ s ( 0 ) r p + c E sup 0 , τ * T r ( ε 2 D t ) φ s ( 0 ) r p + c E sup 0 , τ * χ r p .
Apply Lemmas 1 and 2 to finish the proof. □
Lemma 3.
Let φ s ( 0 ) = O ( 1 ) , then
0 t T r ( ε 2 D s ) φ s ( 0 ) n d s C ε 2 for n 1 .
Proof. 
0 t T r ( ε 2 D s ) φ s ( 0 ) n d s 0 t e ε 2 ω n D τ φ s ( 0 ) n d s = C ε 2 .

4. Averaging over FOUP

Here, we use a comprehensive version of Lemma 5.1 from [28] over the FOUP χ j . This lemma declares that odd powers of χ j are small powers of the order O ( ε 1 κ 0 ) , while the power of χ j averages to a constant.
Lemma 4.
Assume that ϕ is a stochastic process with real values and ϕ ( 0 ) = O ( ε γ ) for some γ 0 . If d ϕ = F d t together with F = O ( ε γ ) , then
0 t ϕ i = 1 N χ i n i d τ = i = 1 N n i ( n i 1 ) α i 2 2 j = 1 N n j D λ j r 2 0 t ϕ j = 1 N χ j n j χ i 2 d τ + O ( ε 1 γ κ 0 ) i f a l l n i a r e e v e n a n d O ( ε 1 γ κ 0 ) i f o n e o f t h e n i i s o d d .
In the next lemma, we utilize Lemma 4 repeatedly and display the outcome that we require afterwards in our application.
Lemma 5.
Assume that ϕ is as in Lemma 4. Then, there is a constant C 2 k for k N such that
0 t ϕ χ 2 k d τ = C 2 k 0 t ϕ d τ + O ( ε 1 γ κ 0 ) ,
where χ is defined in Definition 1.
Proof. 
We address three cases as follows.
First case when k = 1 :
0 t ϕ χ 2 d τ = P c 0 t ϕ j = 1 N χ j e j 2 d τ = j = 1 N ( e j 2 ) 0 t ϕ χ j 2 d τ + 2 j i N ( e j e i ) 0 t ϕ χ j χ i d τ .
From Lemma 4, we have
0 t ϕ χ 2 d τ = j = 1 N e j 2 α j 2 2 D λ j r 2 0 t ϕ d τ + O ( ε 1 γ κ 0 ) .
Thus,
C 2 k = k = 1 C 2 = j = 1 N e j 2 α j 2 2 D λ j r 2 .
Second case when k = 2 :
0 t ϕ χ 4 d τ = 0 t ϕ j = 1 N χ j e j 4 d τ = j = 1 N e j 4 0 t ϕ χ j 4 d τ + 6 j i N e j 2 e i 2 0 t ϕ χ j 2 χ i 2 d τ + 4 k i N e j e i 3 0 t ϕ χ j χ i 3 d τ .
Again, from Lemma 4, we derive
0 t ϕ χ 4 d τ = [ j = 1 N 3 e j 4 α j 4 4 D 2 λ j r + j i N 3 e j 2 e i 2 α j 2 α i 2 2 λ j r 2 λ i r 2 ] 0 t ϕ d τ + O ( ε 1 γ κ 0 ) .
Hence,
C 2 k = k = 2 C 4 = j = 1 N 3 e j 4 α j 4 4 D 2 λ j r + j i N 3 e j 2 e i 2 α j 2 α i 2 2 D 2 λ j r 2 λ i r 2 .
Third case when k > 2 : We can follow the previous cases by expanding
j = 1 N χ j e j 2 k .

5. Limiting Equation and Main Theorem

Here, the limiting equation is derived for Equation (1). Additionally, the main theorem of this paper is stated and proved.
Lemma 6.
Assume that Assumptions 1, 2, and 3 are satisfied. If φ s ( 0 ) = O ( 1 ) , then
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c ) d τ + 0 t G ( φ c ) d τ + R ( t ) ,
where
G ( φ c ) = k 1 C 2 k ( 2 k ) ! [ P c ( φ c ) ] ( 2 k ) ,
and
R = O ( ε 1 ( 2 m 1 ) κ ) .
Proof. 
By recalling (13) and projecting to H c , we have
d φ c = P c ( φ c + φ s ) d t .
W rewrite the above equation in the integral form as
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c + φ s ) d τ .
Recall Lemma 1 which states
φ s ( t ) = χ ( t ) + ψ ( t ) + z ( t ) ,
with
ψ ( t ) = T r ( ε 2 D s ) φ s ( 0 ) and z ( t ) = O ( ε 1 m κ ) .
By substituting (26) into (25), we have
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c + χ + ψ + z ) ( τ ) d τ .
Now, by applying Taylor’s expansion to P c , we derive
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c + χ ) ( τ ) d τ + z ˜ ( t ) ,
where
z ˜ ( t ) = k = 1 m P c 0 t [ P c ( φ c + χ ) ] ( k ) k ! ( ψ + z ) k d τ .
We next apply Taylor’s expansion again to polynomial P c
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c ) ( τ ) d τ + k = 1 m 0 t [ P c ( φ c ) ] ( k ) k ! χ k d τ + z ˜ ( t ) ,
where m is the degree of P . Using (20), we derive
φ c ( t ) = φ c ( 0 ) + 0 t P c ( φ c ) d τ + 0 t G ( φ c ) d τ + R ( t ) ,
where
R ( t ) = z ˜ ( t ) + O ( ε 1 κ 0 ) .
To bound the error R , we take the E sup t 0 , τ * · r p on both sides of (28).
E sup 0 , τ * z ˜ r p C k = 1 m E sup 0 , τ * 0 t [ P c ( φ c + χ ) ] ( k ) k ! ( ψ + z ) k d τ r p .
Using Lemmas 1 and 2, Assumption 1, and the theorem of Burkholder–Davis–Gundy (cf. Theorem 1.2.4 in [43]), the equation yields
E sup 0 , τ * z ˜ r p C ε 1 ( 2 m 1 ) κ .
Substituting (30) into (29) yields (23). □
Lemma 7.
Assume that Assumption 1 is satisfied. Define ξ ( t ) in H c as a solution of (3) with E ξ ( 0 ) p C . Then, for all T 0 > 0 there is C > 0 such that
E sup [ 0 , T 0 ] ξ ( t ) p C .
Proof. 
By taking the scalar product · , ξ on both sides of (3), we derive
1 2 d t ξ 2 = P c ( ξ ) + G ( ξ ) , ξ .
Using Equation (8) yields
1 2 d t ξ 2 C ξ 2 .
By integrating from 0 to t, we have
ξ ( t ) 2 ξ ( 0 ) 2 + 2 C 0 t ξ ( τ ) 2 d τ .
By applying Gronwall’s lemma, we attain for t [ 0 , T 0 ]
ξ ( t ) ξ ( 0 ) e C T 0 .
We take the expectation on both sides after the supremum for Equation (32) to obtain (31). □
In fact, we cannot control the error terms that are defined in terms of φ s or φ c . Therefore, we are limited to a sufficiently large subset of Ω where all our estimates of errors are true.
Definition 3.
Define the set Ω * Ω so that all of these estimations are included:
sup [ 0 , τ * ] φ s r < ε κ 0 1 2 κ ,
sup [ 0 , τ * ] φ s χ T r ( ε 2 D s ) φ s ( 0 ) r < ε 1 m κ κ ,
sup [ 0 , τ * ] ξ < ε 1 2 κ ,
and
sup [ 0 , τ * ] R r < ε 1 2 m κ ,
are valid on Ω * .
As shown below, the set Ω * has a probability of nearly to one.
Proposition 1.
Assume that Assumptions 1 and 2 are satisfied. Then, Ω * has probability
P ( Ω * ) 1 C ε p .
Proof. 
We notice that
P ( Ω * ) 1 P ( sup [ 0 , τ * ] φ s r ε κ 0 1 2 κ ) P ( sup [ 0 , τ * ] φ s χ T r ( ε 2 D s ) φ s ( 0 ) r ε 1 κ ( m + 1 ) )
P ( sup [ 0 , τ * ] ξ > ε 1 2 κ ) P ( sup [ 0 , τ * ] R r ε 1 2 m κ ) .
By first using the Chebychev inequality and afterwards using Lemmas 1, 6, and 7, and Corollary 1, we derive
P ( Ω * ) 1 C [ ε q κ + ε q κ + ε q κ ] 1 C ε q κ 1 C ε p .
Theorem 1.
Let Assumption 2 hold. Assume that ξ and φ c are solutions of (3) and (21), respectively, with ξ ( 0 ) = φ c ( 0 ) = O ( 1 ) . Then,
sup t 0 , τ * φ c ( t ) ξ ( t ) r C ε 1 2 m κ ,
and
sup t 0 , τ * φ c ( t ) r C ε 1 2 κ .
Proof. 
By subtracting (3) from (21), we derive
φ c ( t ) ξ ( t ) = 0 t P c ( φ c ) P c ( ξ ) d τ + 0 t G ( φ c ) G ( ξ ) d τ + R ( t ) .
Let F ( u ) = P c ( u ) + G ( u ) to have
φ c ( t ) ξ ( t ) = 0 t [ F ( φ c ) F ( ξ ) ] d τ + R ( t ) .
Now, define Θ = φ c ξ R to obtain
Θ ( t ) = 0 t [ F ( Θ + ξ + R ) F ( ξ ) ] d τ .
Thus,
d t Θ ( t ) = F ( Θ + ξ + R ) F ( ξ ) .
By taking the scalar product · , Θ on both sides and using Assumption 2, we have
1 2 d t Θ ( t ) 2 = F ( Θ + ξ + R ) F ( ξ ) , Θ Θ ( t ) 2 + R ( t ) 2 + R ( t ) m Θ ( t ) 2 + C ε 2 4 m κ on Ω * .
By using Gronwall’s lemma, we obtain
sup 0 , τ * Θ C ε 1 2 m κ on Ω * .
We finish the first part by using
sup 0 , τ * φ c ξ sup 0 , τ * Θ + sup 0 , τ * R C ε 1 2 m κ .
For the second part, consider
sup 0 , τ * φ c sup 0 , τ * φ c ξ + sup 0 , τ * ξ C ε 1 2 m κ + C ε 1 2 κ C ε 1 2 κ ,
for κ < 1 2 m , where we used the first part and (35). □
Theorem 2. 
(Approximation: suppose that Assumptions 1–3 are satisfied. Let φ be a solution of (1) with splitting φ = φ c + φ s as defined in (12). Additionally, let ξ be a solution of (3) with ξ ( 0 ) = φ c ( 0 ) . Then, for all κ ( 0 , 1 2 m ) and T 0 > 0 , there is C > 0 such that
P sup t 0 , T 0 φ ( t ) ξ ( t ) χ ( t ) T r ( ε 2 D s ) φ s ( 0 ) r > ε 1 2 m κ C ε p ,
for all p > 0 .
Proof. 
We notice that for τ * :
Ω * { sup [ 0 , τ * ] φ c r < ε κ , sup [ 0 , τ * ] φ s r < ε κ } { τ * = T 0 } Ω .
By using Equation (12) and the triangle inequality, we obtain
sup t [ 0 , T 0 ] φ ( t ) ξ ( t ) χ ( t ) T r ( ε 2 D s ) φ s ( 0 ) r
= sup t [ 0 , τ * ] φ ( t ) ξ ( t ) χ ( t ) T r ( ε 2 D s ) φ s ( 0 ) r sup [ 0 , τ * ] φ c ξ r + sup [ 0 , τ * ] φ s χ T r ( ε 2 D s ) φ s ( 0 ) r C ε 1 2 m κ on Ω * .
where we used (34) and (38). Hence,
P sup t [ 0 , T 0 ] φ ( t ) ξ ( t ) χ ( t ) T r ( ε 2 D s ) φ s ( 0 ) r > C ε 1 2 m κ 1 P ( Ω * ) .
By using (37), the equation yields (40). □

6. Application

Throughout chemistry, physics, biology, and other fields of reaction-diffusion equations with non-linearities of polynomials, there are many models in which the main theory of approximation is applied; for example, consider Fisher’s and Fitzhugh–Nagumo equations in biology and the real-valued Ginzburg–Landau equation in physics. Here, we are looking at two models, namely one from physics and the other from biology, as follows.

6.1. Physical Example

The first example is the Ginzburg–Landau equation [20]. The Ginzburg–Landau equation is used for modeling a wide variety of physical systems. Additionally, it was first formulated in the sense of pattern formation as a long-wave amplitude equation in the case of convection in binary mixtures close to the onset of instability. The fractional space Ginzburg–Landau equation with additive noise is
d φ = [ ε 2 ( Δ ) r 2 φ + φ φ 3 ] d t + ε 1 d W for t 0 ,
where the variable φ ( t , x ) is a real-valued function of t and x .
To check Assumption 1, we note that P ( φ ) = φ φ 3 and then for r > 1 2
P ( φ ) r = φ φ 3 r φ r + φ 3 r 2 3 + 1 3 φ r 3 + φ r 3 4 3 ( 1 + φ r 3 ) ,
where we used the Young inequality.
Moreover, we use (22) with k = 1 to obtain
G ( ξ ) = j = 1 N 3 α j 2 2 j r ξ ,
Hence, the limiting equation is
d ξ = [ 1 j = 1 N 3 α j 2 2 j r ξ ξ 3 ] d t .
Now, the solution of (41) by our main theorem is approximated by
φ ( t , x ) ξ ( t ) + χ ( t , x ) ,
where ξ is a solution of (42) and χ is defined in (1). If we suppose that the noise acts only in one mode, i.e., W ( t ) = α j β j cos ( j x ) , then Equation (42) takes the form
d ξ = [ 1 3 α j 2 2 j r ξ ξ 3 ] d t .
If we choose α j such that α j 2 < 2 j r 3 for r ( 1 , 2 ] , then the term ( 1 3 α j 2 2 j r ) is negative. We may say, in this case, that the dynamics of the dominant modes were stabilized by the degenerated additive noise.

6.2. Biological Example

The second example is Fisher’s equation [21]. Fisher’s equation becomes one of the most important types of non-linear equations due to its existence in many chemical and biological processes. Fisher’s equation with fractional space and, by being forced by additive noise, takes the form
d φ = [ ε 2 D ( Δ ) r 2 φ + A φ ( 1 φ K ) ] d t + ε 1 d W ,
where A and K are positive constants. Here, φ ( t , x ) describes the evolution of the state over the spatial–temporal domain defined by the coordinates t and x, respectively.
Our main theory shows that the approximate solution of Fisher’s Equation (44) is
φ ( t , x ) = ξ ( t ) + χ ( t , x ) + e r r o r ,
where ξ is the solution of
d ξ = [ A ξ ( 1 ξ K ) A 2 D K j = 1 N α j 2 j r ] d t .

7. Conclusions

In this article, we obtained the approximation solutions of stochastic fractional-space diffusion equations via the solutions of ordinary differential equations, which are called limiting equations. This equation has never been studied before using a combination of additive noise and fractional-space. We applied our results to many example such as Fisher’s equation and the Ginzburg–Landau models. Additionally, we discussed the influence of degenerate additive noise on the stabilization of the approximate solutions. These solutions are of considerable importance in understanding many important complex physical phenomena as fractional diffusion equations arise in the modeling of turbulent flow, contaminant transport in groundwater flow, and chaotic dynamics of classical conservative systems.

Author Contributions

Conceptualization, W.W.M.; data curation, N.I.; formal analysis, W.W.M.; funding acquisition, T.B.; investigation, N.I.; methodology, W.W.M.; project administration, T.B.; resources, T.B.; software, W.W.M.; validation, N.I.; visualization, N.I. and T.B.; writing—original draft, W.W.M.; writing—review and editing, W.W.M., N.I. and T.B. 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

Not applicable.

Acknowledgments

The authors would like to thank the anonymous referees for valuable comments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, X.; Vanden Eijnden, E. Some recent progress in multiscale modeling. Multiscale Model. Simul. Lect. Notes Comp. Sci. Eng. 2004, 39, 3–21. [Google Scholar]
  2. Imkeller, P.; Monahan, A.H. Conceptual stochastic climate models. Stoch. Dynam. 2002, 2, 311–326. [Google Scholar] [CrossRef]
  3. Barkai, E.; Metzler, R.; Klafter, J. From continuous time random walks to the fractional Fokker–Planck equation. Phys. Rev. E 2000, 61, 132–138. [Google Scholar] [CrossRef]
  4. Metzler, R.; Klafter, J. The random walk’s guide to anomalous diffusion: A fractional dynamics approach. Phys. Rep. 2000, 339, 1–77. [Google Scholar] [CrossRef]
  5. Saichev, A.I.; Zaslavsky, G.M. Fractional kinetic equations: Solutions and applications. Chaos 1997, 7, 753–764. [Google Scholar] [CrossRef] [Green Version]
  6. Zaslavsky, G.M. Chaos, fractional kinetics and anomalous transport. Phys. Rep. 2002, 6, 461–580. [Google Scholar] [CrossRef]
  7. Yuste, S.B.; Lindenberg, K. Subdiffusion-limited A+A reactions. Phys. Rev. Lett. 2001, 87, 118301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Gorenflo, R.; Mainardi, F. Random walk models for space–fractional diffusion processes. Fract. Calc. Appl. Anal. 1998, 1, 1–191. [Google Scholar]
  9. Raberto, M.; Scalas, E.; Mainardi, F. Waiting-times and returns in high-frequency financial data: An empirical study. Phys. A Stat. Mech. Appl. 2002, 314, 749–755. [Google Scholar] [CrossRef] [Green Version]
  10. Wyss, W. The fractional Black–Scholes equation. Fract. Calc. Appl. Anal. 2000, 3, 51–61. [Google Scholar]
  11. Yuste, S.B.; Acedo, L.; Lindenberg, K. Reaction front in an A+BC reaction–subdiffusion process. Phys. Rev. E 2004, 69, 036126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Benson, D.A.; Wheatcraft, S.W.; Meerschaert, M.M. The fractional-order governing equation of Lévy motion. Water Resour. Res. 2000, 36, 1413–1423. [Google Scholar] [CrossRef]
  13. Liu, F.; Anh, V.; Turner, I. Numerical solution of the space fractional Fokker–Planck equation. J. Comput. Appl. Math. 2004, 166, 209–219. [Google Scholar] [CrossRef] [Green Version]
  14. Podlubny, I. Fractional Differential Equations; Academic Press: New York, NY, USA, 1999. [Google Scholar]
  15. Samko, S.G.; Kilbas, A.A.; Marichev, O.I. Fractional Integrals and Derivatives: Theory and Applications; Gordon and Breach Science Publishers: Philadelphia, PA, USA, 1993. [Google Scholar]
  16. Iqbal, N.; Wu, R.; Mohammed, W.W. Pattern formation induced by fractional cross-diffusion in a 3-species food chain model with harvesting. Math. Comput. Simul. 2021, 188, 102–119. [Google Scholar] [CrossRef]
  17. Iqbal, N.; Wu, R. Pattern formation by fractional cross-diffusion in a predator-prey model with Beddington-DeAngelis type functional response. Int. J. Mod. Phys. 2019, 33, 1950296. [Google Scholar] [CrossRef]
  18. Fitzhugh, R. Impulse and physiological states in models of nerve membrane. Biophys. J. 1961, 1, 445–466. [Google Scholar] [CrossRef] [Green Version]
  19. Nagumo, J.S.; Arimoto, S.; Yoshizawa, S. An active pulse transmission line simulating nerve axon. Proc. IRE 1962, 50, 2061–2070. [Google Scholar] [CrossRef]
  20. Bethuel, F.; Brezis, H.; Hélein, F. Ginzburg–Landau Vortices; Birkhä: Basel, Switzerland; Boston, MA, USA, 1994. [Google Scholar]
  21. Fisher, R.A. The wave of advance of advantageous genes. Ann. Eugen. 1936, 7, 355–369. [Google Scholar] [CrossRef] [Green Version]
  22. Malflict, W. Solitary wave solutions of nonlinear wave equations. Am. J. Phys. 1992, 60, 650–654. [Google Scholar] [CrossRef]
  23. Aronson, D.J.; Weinberg, H.F. Nonlinear Diffusion in Population Genetics Combustion and Never Pulse Propagation; Springer: New York, NY, USA, 1988. [Google Scholar]
  24. Frank, D.A. Diffusion and Heat Exchange in Chemical Kinetics; Princeton University Press: Princeton, NJ, USA, 1955. [Google Scholar]
  25. Tuckwell, H.C. Introduction to Theoretical Neurobiology; Cambridge University Press: Cambridge, CA, USA, 1988. [Google Scholar]
  26. Canosa, J. Diffusion in nonlinear multiplication media. J. Math. Phys. 1969, 186, 2–9. [Google Scholar]
  27. Mohammed, W.W. Amplitude equation for the stochastic reaction-diffusion equations with random Neumann boundary conditions. Math. Methods Appl. Sci. 2015, 38, 48–4878. [Google Scholar] [CrossRef]
  28. Blömker, D.; Mohammed, W.W. Amplitude equations for SPDEs with cubic nonlinearities. Stochastic Int. J. Probability Stoch. Process. 2013, 85, 181–215. [Google Scholar] [CrossRef] [Green Version]
  29. Mohammed, W.W.; Blömker, D. Fast-diffusion limit with large noise for systems of stochastic reaction-diffusion equations. J. Stoch. Anal. Appl. 2016, 34, 961–978. [Google Scholar] [CrossRef] [Green Version]
  30. Mohammed, W.W.; Blömker, D.; Klepel, K. Multi-Scale analysis of SPDEs with degenerate additive noise. J. Evol. Equ. 2014, 14, 273–298. [Google Scholar] [CrossRef] [Green Version]
  31. Bulut, H.; Atas, S.S.; Baskonus, H.M. Some novel exponential function structures to the Cahn-Allen equation. Cogent Phys. 2016, 3. [Google Scholar] [CrossRef]
  32. Jeong, D.; Kim, J. An explicit hybrid finite difference scheme for the Allen–Cahn equation. J. Comput. Appl. Math. 2018, 340, 247–255. [Google Scholar] [CrossRef]
  33. Mohammed, W.W.; Ahmad, H.; Hamza, A.E.; ALy, E.S.; El-Morshedy, M.; Elabbasy, E.M. The exact solutions of the stochastic Ginzburg–Landau equation. Results Phys. 2021, 23, 103988. [Google Scholar] [CrossRef]
  34. Mohammed, W.W.; Blömker, D. Fast-diffusion limit for reaction-diffusion equations with multiplicative noise. J. Math. Anal. Appl. 2021, 496, 124808. [Google Scholar] [CrossRef]
  35. Liu, F.; Anh, V.; Turner, I.; Zhuang, P. Numerical simulation for solute transport in fractal porous media. ANZIAM J. 2004, 45, 461–473. [Google Scholar] [CrossRef] [Green Version]
  36. Ilic, M.; Liu, F.; Turner, I.; Anh, V. Numerical approximation of a fractional-in-space diffusion equation. Fract. Calc. Appl. Anal. 2005, 8, 323–341. [Google Scholar]
  37. Ilic, M.; Liu, F.; Turner, I.; Anh, V. Numerical approximation of a fractional-in-space diffusion equation (II)—With nonhomogeneous boundary conditions. Fract. Calc. Appl. Anal. 2006, 9, 333–349. [Google Scholar]
  38. Shen, S.; Liu, F. Error analysis of an explicit finite difference approximation for the space fractional diffusion. ANZIAM J. 2005, 46, 871–887. [Google Scholar] [CrossRef] [Green Version]
  39. Tadjeran, C.; Meerschaert, M.M.; Scheffler, H.P. A second-order accurate numerical approximation for the fractional diffusion equation. J. Comput. Phys. 2006, 213, 205–213. [Google Scholar] [CrossRef]
  40. Dehghan, M.; Manafian, J.; Saadatmandi, A. Solving nonlinear fractional partial differential equations using the homotopy analysis method. Numer. Methods Partial. Differ. Equ. 2010, 26, 448–479. [Google Scholar] [CrossRef]
  41. Saadatmandi, A.; Dehghan, M. A new operational matrix for solving fractional-order differential equations. Comput. Math. Appl. 2010, 59, 1326–1336. [Google Scholar] [CrossRef] [Green Version]
  42. Da Prato, G.; Zabczyk, J. Stochastic equations in infinite dimensions. In Encyclopedia of Mathematics and Its Applications; Cambridge University Press: Cambridge, UK, 1992; Volume 44. [Google Scholar]
  43. Liu, K. Stability of Infinite Dimensional Stochastic Differential Equations with Applications; Chapman and Hall/CRC Monographs: Boca Raton, FL, USA, 2006; Volume 135. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohammed, W.W.; Iqbal, N.; Botmart, T. Additive Noise Effects on the Stabilization of Fractional-Space Diffusion Equation Solutions. Mathematics 2022, 10, 130. https://doi.org/10.3390/math10010130

AMA Style

Mohammed WW, Iqbal N, Botmart T. Additive Noise Effects on the Stabilization of Fractional-Space Diffusion Equation Solutions. Mathematics. 2022; 10(1):130. https://doi.org/10.3390/math10010130

Chicago/Turabian Style

Mohammed, Wael W., Naveed Iqbal, and Thongchai Botmart. 2022. "Additive Noise Effects on the Stabilization of Fractional-Space Diffusion Equation Solutions" Mathematics 10, no. 1: 130. https://doi.org/10.3390/math10010130

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

Article Metrics

Back to TopTop