Previous Article in Journal
Accurate Implementation of Rotating Magneto-Hydrodynamics in a Channel Geometry Using an Influence Matrix Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Convergence of Slow–Fast Stochastic Reaction–Diffusion–Advection Equation with Hölder-Continuous Coefficients

School of Mathematical Sciences, Tiangong University, Tianjin 300387, China
*
Author to whom correspondence should be addressed.
Mathematics 2025, 13(16), 2550; https://doi.org/10.3390/math13162550 (registering DOI)
Submission received: 15 May 2025 / Revised: 15 July 2025 / Accepted: 7 August 2025 / Published: 8 August 2025

Abstract

This paper investigates a slow–fast stochastic reaction–diffusion–advection equation with Hölder-continuous coefficients, where the irregularity of the coefficients presents significant analytical challenges. Our approach fundamentally relies on techniques from Poisson equations in Hilbert spaces, through which we establish optimal strong convergence rates for the approximation of the averaged solution by the slow component. The key advantage that this paper presents is that the coefficients are merely Hölder continuous yet the optimal rate can still be obtained, which is crucial for subsequent central limit theorems and numerical approximations.

1. Introduction

Many natural systems in physics, chemistry, and biology can be mathematically represented using advection–reaction–diffusion equations. The standard formulation of such equations is given by
X t X ξ ξ + X 3 X X ξ = 0 ,
which can be used to describe the dynamics of spatially distributed biological, chemical, and physical processes, ecological models, genetic models, population distribution, etc.; see [1,2] and the references therein for more studies.
Let T > 0 and D = ( 0 , 1 ) . Consider the following slow–fast stochastic reaction–diffusion–advection equations:
𝜕 X t ε ( ξ ) 𝜕 t = Δ X t ε ( ξ ) + 1 2 𝜕 𝜕 ξ ( X t ε ( ξ ) ) 2 X t ε 3 ( ξ ) + f ( X t ε ( ξ ) , Y t ε ( ξ ) ) + 𝜕 W t 1 ( ξ ) 𝜕 t , 𝜕 Y t ε ( ξ ) 𝜕 t = 1 ε Δ Y t ε ( ξ ) + 1 ε g ( X t ε ( ξ ) , Y t ε ( ξ ) ) + 1 ε 𝜕 W t 2 ( ξ ) 𝜕 t , X t ε ( 0 ) = X t ε ( 1 ) = Y t ε ( 0 ) = Y t ε ( 1 ) = 0 , t ( 0 , T ] , X 0 ε ( ξ ) = x , Y 0 ε ( ξ ) = y , ξ D ,
where f and g are reaction coupling satisfying some appropriate conditions, and W t 1 and W t 2 are mutually independent Q 1 - and Q 2 -Wiener processes, both defined on some probability space ( Ω , F , P ) with a normal filtration { F t } t 0 . The parameter 0 < ε 1 characterizes the time-scale separation between the slow component X t ε and the fast component Y t ε .
Such a multiscale model arises from describing multiscale phenomena and frequently appears in many real-world dynamical systems. Typical examples include but are not limited to phytoplankton dynamics (see, e.g., [3,4]), species competition (see, e.g., [5]), cell modeling (see, e.g., [6]), Hamiltonian systems (see, e.g., [7]), and electronic circuits (see, e.g., [8]), etc.
However, directly analyzing or simulating the multiscale system (1) remains challenging due to the strong time-scale separation between slow and fast modes and their nonlinear coupling effects. To address this, simplified models capturing the long-time system evolution are crucial for practical applications. This reduction methodology originates from the averaging principle, initially developed for deterministic systems by Bogoliubov [9] and later generalized to stochastic differential equations (SDEs) by Khasminskii [10]. Subsequent research has extensively explored averaging techniques for finite-dimensional stochastic systems, as documented in [11,12,13,14,15,16,17] and the references therein. The extension of averaging principles to infinite-dimensional systems presents significant mathematical challenges and has only been systematically developed in recent decades. Cerrai and Freidlin [18] established the averaging principle for stochastic reaction–diffusion equations, focusing on systems where the slow dynamics are deterministic. Later, Cerrai [19,20] extended this result to more general cases; see also [21,22,23,24,25,26,27,28] and the references therein for further developments regarding multiscale stochastic partial differential equations (SPDEs for short).
Note that all the works mentioned above focus on linear or semi-linear SPDEs. However, it is widely recognized that numerous models and equations in physics involve highly nonlinear terms. Consequently, multiscale SPDEs with highly nonlinear terms have garnered increasing attention in this field. Here are some representative references for different classes of nonlinear SPDEs: for parabolic-type SPDEs, we refer to [29,30,31], while hyperbolic-type systems are addressed in [32]; for applications in fluid mechanics with physical motivations, we point to [33,34], etc. We note that Gao [30] studied the strong convergence of the system (1) with Lipschitz continuous coefficients by using the techniques of time discretization and stopping time; however, the optimal convergence rate was not achieved. Very recently, Gao and Sun [31] studied the multiscale stochastic Burgers equation and obtained the optimal strong convergence rate, where the coefficients are assumed to be regular.
In the present paper, we focus our attention on system (1) with Hölder-continuous coefficients. Moreover, we shall establish a stronger convergence result in the averaging principle and provide the optimal convergence rate. To be precise, it is shown that, for every T > 0 , q 1 and γ [ 0 , 1 / 2 ) , there exists a constant C 0 > 0 such that
sup t [ 0 , T ] E ( A ) γ ( X t ε X ¯ t ) q C 0 ε q 2 ,
and also see Theorem 1 below. The main contributions are summarized as follows:
  • Our results significantly relax the regularity assumptions in [30,31], requiring only Hölder continuity in the fast variable while achieving stronger convergence regarding the · ( A ) γ norm with any γ [ 0 , 1 / 2 ) . We note that the established convergence of ( A ) γ X t ε to ( A ) γ X ¯ t proves particularly crucial for the analysis of the central limit theorem, which will be processed in the following work.
  • Our method for establishing strong convergence fundamentally differs from existing approaches (see, e.g., [18,19,20,23,24,28]), where the time discretization procedure is used. Our method is based on the Poisson equation; see Theorem 2.
  • Highly nonlinear terms, incluing the cubic term X t ε 3 ( ξ ) and the advection term 1 2 𝜕 𝜕 ξ ( X t ε ( ξ ) ) 2 , will also cause additional difficulties, which are addressed through some exponential moment estimates.
  • We establish the characteristic 1 2 -order convergence rate, which is known to be sharp for finite-dimensional systems (when γ = 0 ), as demonstrated in [35]. Under the regularity assumptions regarding noise (A2), we show that the averaging convergence is independent of the fast variable’s coefficient regularity.
In the following Section 2, we introduce the main theorem of this paper; in Section 3 and Section 4, we use the Poisson equation and exponential estimation technique to prove the result of strong convergence.
Notations. To end this section, we introduce some notations: p [ 1 , ] , and let L p : = L p ( D ) denote the Banach space of functions on D equipped with the L p -norm · L p . For p = 2 , we denote the Hilbert space L 2 ( D ) endowed with scalar product · , · and norm · .   p , q [ 2 , ) , and we denote by L ( L p , L q ) the space of all the bounded linear operators from L p to L q . When p = q = 2 , we write L ( H ) = L ( H , H ) for simplicity. An operator Q L ( H ) is called Hilbert–Schmidt if
Q L 2 ( H ) 2 : = T r ( Q Q ) < + .
The space of all the Hilbert–Schmidt operators on H is denoted by L 2 ( H ) .
Next, we define the Gâteaux and Fréchet derivatives of ϕ with respect to the x variable. A mapping ϕ : H × H H ^ , where H ^ is another Hilbert space, is Gâteaux differentiable at x if there exists a bounded linear operator D x ϕ ( x , y ) L ( H , H ^ ) satisfying
lim τ 0 ϕ ( x + τ h , y ) ϕ ( x , y ) τ = D x ϕ ( x , y ) . h , h H .
The mapping ϕ is Fréchet differentiable at x if it is Gâteaux differentiable and additionally satisfies the stronger condition:
lim h 0 ϕ ( x + h , y ) ϕ ( x , y ) D x ϕ ( x , y ) . h H ^ h = 0 .
For k 2 , we inductively define the k times Gâteaux derivative D x k ϕ ( x , y ) as an element of L k ( H , H ^ ) : = L ( H , L ( k 1 ) ( H , H ^ ) ) , endowed with the operator norm
D x k ϕ ( x , y ) L k ( H , H ^ ) : = sup h 1 1 , h 2 1 , , h k 1 , h 1 D x k ϕ ( x , y ) . ( h 1 , h 2 , , h k ) , h H ^ .
The above definitions of derivatives with respect to the variable x can be analogously defined for the y variable: D y ϕ ( x , y ) L ( H , H ^ ) , and for k 2 , D y k ϕ ( x , y ) L k ( H , H ^ ) : = L ( H , L ( k 1 ) ( H , H ^ ) ) .
Finally, for ϕ L ( H × H , H ^ ) with norm ϕ L ( H ^ ) : = sup ( x , y ) H × H ϕ ( x , y ) H ^ < , we introduce several differentiable function spaces of importance for our analysis:
(1)
ϕ C b k , 0 ( H × H , H ^ ) , k N : ϕ is k times Gâteaux differentiable at any x H with bounded derivatives.
(2)
ϕ C b 0 , k ( H × H , H ^ ) , k N : ϕ is k times Gâteaux differentiable at any y H with bounded derivatives.
(3)
ϕ C b 0 , k ( H × H , H ^ ) , k N : ϕ is k times Fréchet differentiable at any y H with bounded derivatives.
(4)
ϕ C b k , θ ( H × H , H ^ ) , θ ( 0 , 1 ) : ϕ C b k , 0 ( H × H , H ^ ) , satisfying ϕ ( x , y 1 ) ϕ ( x , y 2 ) H ^ C 0 y 1 y 2 θ .
When H ^ = R , we omit the symbol H ^ in the preceding notations for simplicity.
Throughout this paper, C, with or without subscripts, denotes a positive constant. Its value may vary from line to line, and its dependence on parameters will be clear from the context.

2. Assumptions and Main Results

Let H : = L 2 ( 0 , 1 ) be the usual space of square-integrable functions with scalar product and norm denoted by · , · and · , respectively. For k N , W k , p ( 0 , 1 ) is the Sobolev space of all functions in L p ( 0 , 1 ) whose differentials belong to L p ( 0 , 1 ) up to the order k. The usual Sobolev space W k , p ( 0 , 1 ) can be extended to the W s , p ( 0 , 1 ) for s R . Set H k : = W k , 2 ( 0 , 1 ) and denote by H 0 1 the subspace of H 1 of all functions whose trace at 0 and 1 vanishes.
Let
A x : = 𝜕 2 𝜕 ξ 2 x , x D ( A ) = H 2 H 0 1 .
It is well-established that the Hilbert space H admits a complete orthonormal basis { e n } n N such that
A e n = λ n e n ,
with 0 < λ 1 λ 2 λ n . For α R , denote by H α : = D ( ( A ) α 2 ) the Hilbert space equipped with the scalar product
x , y α : = ( A ) α 2 x , ( A ) α 2 y = n = 1 λ n α x , e n y , e n , x , y H α ,
and norm
x α : = n = 1 λ n α x , e n 2 1 2 , x H α .
It is easy to deduce · 0 = · , and the following inequalities hold:
e t A x α 2 C α 1 , α 2 t α 2 α 1 2 e λ 1 t 2 x α 1 , x H α 2 , α 1 α 2 , t > 0 ;
e t A x x C t α 2 x α , x H α , α 0 , t 0 .
Define the bilinear operator B ( x , y ) : H ( 0 , 1 ) × H 0 1 ( 0 , 1 ) H 0 1 ( 0 , 1 ) by
B ( x , y ) = x . 𝜕 ξ y ,
and the trilinear operator b ( x , y , z ) : H ( 0 , 1 ) × H 0 1 ( 0 , 1 ) × H ( 0 , 1 ) R by
b ( x , y , z ) = 0 1 x ( ξ ) . 𝜕 ξ y ( ξ ) . z ( ξ ) d ξ .
For convenience, let B ( x ) = B ( x , x ) , where x H 0 1 .
For the drift coefficients f and g, we define the Nemytskii operators F , G : H × H H by
F ( x , y ) ( ξ ) : = f ( x ( ξ ) , y ( ξ ) ) , G ( x , y ) ( ξ ) : = g ( x ( ξ ) , y ( ξ ) ) , ξ ( 0 , 1 ) .
Assume that
(A1): for some θ > 0 ,
f , g C b 2 , θ ( R × R , R ) .
Then, we have that (see, e.g., [36])
F , G C b 1 , θ ( H × H , H ) .
Moreover, if x , y H and p , r 1 , r 2 [ 1 , ] satisfy 1 p = 1 r 1 + 1 r 2 , there exists a constant C > 0 such that
D x 2 F ( x , y ) . ( h 1 , h 2 ) L p C h 1 L r 1 h 2 L r 2
and
D x 2 G ( x , y ) . ( h 1 , h 2 ) L p C h 1 L r 1 h 2 L r 2 .
To provide precise results, let E ( x ) = x 3 and write the system (1) in the following abstract formulation in H:
d X t ε = A X t ε d t + B ( X t ε ) d t + E ( X t ε ) d t + F ( X t ε , Y t ε ) d t + d W t 1 , X 0 ε = x , d Y t ε = 1 ε A Y t ε d t + 1 ε G ( X t ε , Y t ε ) d t + 1 ε d W t 2 , Y 0 ε = y .
In present paper, we assume that system (8) is well-posed (see Remark 1 for an explanation). We further assume that
(A2):  Q i are nonnegative symmetric operators, which have the same eigenfunctions { e n } n N as the operator A, i.e.,
Q i e n = β i , n e n , β i , n > 0 , n N ,
and
T r ( ( A ) Q 1 ) < + and T r ( Q 2 ) : = n N β 2 , n < + ,
and, for any T > 0 and i = 1 , 2 ,
0 T Γ i , t 1 + ϑ 2 d t < + ,
where
Γ 1 , t : = sup n 1 2 λ n β 1 , n ( e 2 λ n t 1 ) and Γ 2 , t : = sup n 1 2 λ n β 2 , n ( e 2 λ n t 1 ) ,
λ n is given by (2), and ϑ max ( θ , 1 θ ) with θ is the parameter in the assumption (5). Condition (11) is crucial to prove Theorem 2 below; see Lemma 9 in [37].
Given x H , consider the following frozen equation:
d Y t x = A Y t x d t + G ( x , Y t x ) d t + d W t 2 , Y 0 x = y H 1 .
Under conditions (5), (9), (10), and (11), Equation (12) admits a unique strong solution Y t x (see, e.g., [37]), which processes a unique invariant measure μ x ( d y ) (see, e.g., Theorem 4 and Proposition 4 in [38]). Thus, we shall have the averaged equation for system (8):
d X ¯ t = A X ¯ t d t + B ( X ¯ t ) d t + E ( X ¯ t ) d t + F ¯ ( X ¯ t ) d t + d W t 1 , X ¯ 0 = x H ,
where
F ¯ ( x ) : = H F ( x , y ) μ x ( d y ) ,
By Theorem 2 below, we have F ¯ C b 1 ( H , H ) . Thus, Equation (13) admits a unique solution X ¯ t . The following is the main result.
Theorem 1 (Strong convergence).
Let x H l H 0 1 , l ( 1 / 2 , 1 ] and y H . Assume that (A1)(A2) hold. Then, for any q 1 , T > 0 and γ [ 0 , 1 / 2 ) , we have
sup t [ 0 , T ] E ( A ) γ ( X t ε X ¯ t ) q C ε q 2 ,
where X ¯ t is the unique solution of Equation (13), and C = C ( T , x , y ) > 0 is a constant independent of ε.
Remark 1. 
Note that the strong well-posedness of parabolic SPDEs with Hölder-continuous coefficients has been established in [37] through the application of Zvonkin’s transformation. To be precise, based on the associated infinite-dimensional Kolmogorov equation, we systematically eliminate the irregular drift components through an exact decomposition. This transformation yields a modified system with good Lipschitz properties. Similar techniques could potentially be adapted to prove the well-posedness of SPDE (8) under our assumptions. However, such an investigation falls outside the scope of the present work, and we therefore refrain from pursuing this direction here.
Remark 2. 
In contrast to previous works [30,31] that required smoother coefficients, our results hold under the weaker assumption of Hölder-continuous coefficients in the fast variable. Moreover, we establish the stronger convergence in the · ( A ) γ norm for any γ [ 0 , 1 / 2 ) .

3. Preliminaries and A-Priori Estimates

3.1. Poisson Equation in the Hilbert Space

Consider the following Poisson equation in the Hilbert space H × H :
L 2 ( x , y ) ψ ( x , y ) = ϕ ( x , y ) , y H ,
where L 2 ( x , y ) is defined by
L 2 φ ( x , y ) : = L 2 ( x , y ) φ ( x , y ) : = A y + G ( x , y ) , D y φ ( x , y ) + 1 2 T r D y 2 φ ( x , y ) Q 2 , φ C b 0 , 2 ( H × H ) ,
and x H is a parameter, and ϕ : H × H H is measurable. Since Equation (16) is on the whole space, we define the following centering condition:
0 t ϕ ( x , y ) μ x ( d y ) = 0 , x H ,
which arises naturally in our analysis and serves a role analogous to the centering condition in classical central limit theorems [39,40]. Furthermore, we assume that
( A ϕ ): ϕ ( x , y ) C b 1 , θ ( H × H , H ) , and, for any k H κ , x , h , y H ,
D x 2 ϕ ( x , y ) . ( h , k ) L p C h L r 1 k L r 2 .
The following results can be proved similarly to in Theorem 3.2 and Lemma 3.7 from [27]; we omit the details here.
Theorem 2. 
Assume that (9)–(11), (5), and (7) hold. Then, ϕ : H × H H satisfies the centering condition (17) and ( A ϕ ), and Equation (16) admits a unique solution ψ C b 0 , 2 ( H × H , H ) C b 1 , 0 ( H × H , H ) , which is given by
ψ ( x , y ) = 0 E ϕ ( x , Y t x ( y ) ) d t ,
where Y t x ( y ) satisfies the frozen Equation (12), and k H κ and x , y , h H ,
D x 2 ψ ( x , y ) . ( h , k ) C h L r 1 k L r 2 .
Furthermore, assume that F satisfies (5) and (6), and let F ¯ be defined by (14). Then, we have F ¯ C b 1 ( H , H ) . Moreover, for any k H κ , x , h H ,
D x 2 F ¯ ( x ) . ( h , k ) C h k κ ,
where C > 0 is a constant.

3.2. The A-Priori Estimates

To handle the analytical difficulties arising from the unbounded operators in our system, we employ a Galerkin approximation scheme to reduce the infinite-dimensional problem to finite-dimensional subspaces. This standard approach, whose detailed implementation can be found in Section 4 in [23], provides a rigorous foundation for our subsequent analysis. For notational simplicity and to maintain focus on the essential aspects of our arguments, we suppress explicit reference to the approximation procedure in this subsection and Section 4. We emphasize that all the estimates and bounds presented in what follows are to be understood as holding uniformly with respect to the approximation parameter. This uniformity ensures the validity of our results when passing to the infinite-dimensional limit.
Firstly, we provide the following estimates for the mild solution of the system (8).
Lemma 1. 
Let x , y H , T > 0 and q 1 . Then, the solution ( X t ε , Y t ε ) of system (8) satisfies
sup ε ( 0 , 1 ) sup t [ 0 , T ] E X t ε q C q , T ( 1 + x q ) ,
sup ε ( 0 , 1 ) sup t [ 0 , T ] E Y t ε q C q , T ( 1 + y q ) ,
where C q , T > 0 is a constant.
Proof. 
Based on the boundedness of F and G and Lemma 5.2 from [30], Lemma 1 can be proved. □
Lemma 2. 
Let T > 0 , η ( 1 , 3 / 2 ) , x H l with l [ 0 , η ] , and y H . Then, there exists m > 1 such that, for every q 1 and t ( 0 , T ] , we have
E X t ε η q 1 q C q , η , l , T t η l 2 ( e λ 1 t 2 x l + x m + 1 ) ,
where C q , η , l , T > 0 is a constant.
Proof. 
The proof is in Appendix A. □
Remark 3. 
By Lemma 2 and the interpolation inequality, we have that, for every T > 0 , p 1 , η 1 ( 0 , 1 ] , η ( 1 , 3 / 2 ) , and x H l with l [ 0 , 1 ] , there exists m > 1 such that, for every t ( 0 , T ] ,
E X s ε η 1 q 1 q C t ( η l ) η 1 2 η ( e λ 1 η 1 t 2 η x l η 1 η + x m η 1 η + 1 ) .
To derive the estimate for E X t ε 2 , we need the following regularity results for X t ε and Y t ε with respect to the time variable.
Lemma 3. 
(i) Let T > 0 , α ( 1 / 2 , 2 ] , x H l with l ( 1 / 2 , α ] and y H . Then, for any q 1 and 0 < s t T , we have
E X t ε X s ε q 1 q C q , α , l , T ( ( t s ) α 2 s α l 2 e λ 1 s 2 x l + ( t s ) 1 2 ) ;
where C q , α , l , T > 0 is a constant.
(ii) Let T > 0 , β [ 0 , 1 ] , x H , y H l with l [ 0 , β ] . Then, for any q 1 and 0 < s t T , we have
E Y t ε Y s ε q 1 q C q , β , l , T ( ( t s ) β / 2 s β / 2 l / 2 ε l / 2 e λ 1 s 2 ε y l + ( t s ) β / 2 ε β / 2 ) ;
where C q , β , l , T > 0 is a constant.
Proof. 
The proof is in Appendix A. □
By exactly the same arguments as in Lemma 3.6 in [31], we further have
Lemma 4. 
Let T > 0 , η ( 1 , 3 / 2 ) , x H l with l ( 1 / 2 , η ) and y H . Then, there exists m > 1 such that, for any q 1 and 0 < s t T , we have
E X t ε X s ε 1 q 1 q C q , η , l , T ( t s ) η 1 2 s η l 2 ( 1 + e λ 1 s 2 x l + x m ) ,
Now, we provide the following uniform estimate for X t ε 2 .
Lemma 5. 
Let T , β > 0 , η ( 1 , 3 / 2 ) , x H l with l ( 1 / 2 , 1 ] and y H . Then, there exists m > 1 such that, for any q 1 and t ( 0 , T ] , we have
E X t ε 2 q 1 q C q , β , l , T ( t 1 + 3 l η 2 η + ε β 2 ) ( 1 + x l + x m + y ) ,
where C q , β , l , T > 0 is a constant.
Proof. 
The proof is in Appendix A. □

4. Strong Convergence in the Averaging Principle

Let
L 1 ( x , y ) φ ( x , y ) : = A x + B ( x ) + E ( x ) + F ( x , y ) , D x φ ( x , y ) + 1 2 T r D x 2 φ ( x , y ) Q 1 , φ C b 2 , 0 ( H × H ) .
We begin by deriving crucial strong fluctuation estimates for integral functionals of the coupled process ( X r ε , Y r ε ) over arbitrary time intervals [ s , t ] . These estimates serve as fundamental tools for establishing our main convergence result in Theorem 1.
Lemma 6 (Strong fluctuation estimate).
Let T > 0 , x H l , l ( 1 / 2 , 1 ] , y H . Assume that (A1)(A2) hold. Then, for any q 1 , γ [ 0 , l / 2 ) , 0 s t T , and ϕ C b 2 , θ ( H × H , H ) satisfying (17), we have
E s t ( A ) γ e ( t r ) A ϕ ( X r ε , Y r ε ) d r q C q , γ , T ( t s ) ( l / 2 γ ) q ε q / 2 ,
where C q , γ , T > 0 is a constant.
Proof. 
Consider the Poisson equation
L 2 ( x , y ) ψ ( x , y ) = ϕ ( x , y ) ,
and let
ψ t , γ ( r , x , y ) : = ( A ) γ e ( t r ) A ψ ( x , y ) .
By the definition of L 2 , we have
L 2 ( x , y ) ψ t , γ ( r , x , y ) = ( A ) γ e ( t r ) A ϕ ( x , y ) .
Applying Itô’s formula, it follows that
ψ t , γ ( t , X t ε , Y t ε ) = ψ t , γ ( s , X s ε , Y s ε ) + s t ( 𝜕 r + L 1 ) ψ t , γ ( r , X r ε , Y r ε ) d r + 1 ε s t L 2 ψ t , γ ( r , X r ε , Y r ε ) d r + M t , s 1 + 1 ε M t , s 2 ,
where M t , s 1 and M t , s 2 are defined by
M t , s 1 : = s t D x ψ t , γ ( r , X r ε , Y r ε ) d W r 1 a n d M t , s 2 : = s t D y ψ t , γ ( r , X r ε , Y r ε ) d W r 2 .
Multiplying (25) by ε and applying (24), we get
s t ( A ) γ e ( t r ) A ϕ ( X r ε , Y r ε ) d r = s t L 2 ψ t , γ ( r , X r ε , Y r ε ) d r = ε ψ t , γ ( s , X s ε , Y s ε ) ψ t , γ ( t , X t ε , Y t ε ) + ε s t ( 𝜕 r + L 1 ) ψ t , γ ( r , X r ε , Y r ε ) d r + ε M t , s 1 + ε 1 / 2 M t , s 2 .
Note that
s t 𝜕 r ψ t , γ ( r , X r ε , Y r ε ) d r = s t 𝜕 r ψ t , γ ( r , X t ε , Y t ε ) d r + s t 𝜕 r ψ t , γ ( r , X r ε , Y r ε ) ψ t , γ ( r , X t ε , Y t ε ) d r = ψ t , γ ( t , X t ε , Y t ε ) ψ ˜ t , γ ( s , X t ε , Y t ε ) + s t 𝜕 r ψ t , γ ( r , X r ε , Y r ε ) ψ t , γ ( r , X t ε , Y t ε ) d r ,
and
𝜕 r ψ t , γ ( r , x , y ) = ( A ) 1 + γ e ( t r ) A ψ ( x , y ) .
Consequently, we further get
s t ( A ) γ e ( t r ) A ϕ ( X r ε , Y r ε ) d r = ε ( A ) γ e ( t s ) A ψ ( X s ε , Y s ε ) ψ ( X t ε , Y t ε ) + ε s t ( A ) 1 + γ e ( t r ) A ψ ( X r ε , Y r ε ) ψ ( X t ε , Y t ε ) d r + ε s t L 1 ψ t , γ ( r , X r ε , Y r ε ) d r + ε M t , s 1 + ε 1 / 2 M t , s 2 .
Thus, 0 s t T and q 1 , so we get
E s t ( A ) γ e ( t r ) A ϕ ( X r ε , Y r ε ) d r q C q ( ε q E ( A ) γ e ( t s ) A ψ ( X s ε , Y s ε ) ψ ( X t ε , Y t ε ) q + E ε s t ( A ) 1 + γ e ( t r ) A ψ ( X r ε , Y r ε ) ψ ( X t ε , Y t ε ) d r q + E ε s t L 1 ψ t , γ ( r , X r ε , Y r ε ) d r q + ε q E M t , s 1 q + ε q / 2 E M t , s 2 q ) = : i = 1 5 J i ( t , s , ε ) .
According to Theorem 2, Lemma 1, (20), and (21), we obtain
J 1 ( t , s , ε ) C 1 ε q ( t s ) γ q E 1 + X t ε p + X s ε p + Y t ε p + Y s ε p 2 q 1 / 2 · E X t ε X s ε 2 q + E Y t ε Y s ε 2 q 1 / 2 C 1 ( t s ) ( l / 2 γ ) q ε ( 1 l / 2 ) q C 1 ( t s ) ( l / 2 γ ) q ε q / 2 .
For the second term, by Minkowski’s inequality, we also get
J 2 ( t , s , ε ) C 2 ε q ( s t ( t r ) 1 γ [ E X t ε X r ε 2 q 1 / 2 q + E Y t ε Y r ε 2 q 1 / 2 q ] d r ) q C 2 ε q s t ( t r ) 1 γ ( t r ) l / 2 ε l / 2 d r q C 2 ( t s ) ( l / 2 γ ) q ε ( 1 l / 2 ) q C 2 ( t s ) ( l / 2 γ ) q ε q / 2 .
For the third term, according to definition (23), we deduce that
J 3 ( t , s , ε ) = ε q E s t ( A ) γ e ( t r ) A L 1 ψ ( X r ε , Y r ε ) d r q C 3 ε q E s t ( A ) γ e ( t r ) A A X r ε , D x ψ ( X r ε , Y r ε ) d r q + C 3 ε q E s t ( A ) γ e ( t r ) A B ( X r ε ) , D x ψ ( X r ε , Y r ε ) d r q + C 3 ε q E s t ( A ) γ e ( t r ) A E ( X r ε ) , D x ψ ( X r ε , Y r ε ) d r q + C 3 ε q E s t ( A ) γ e ( t r ) A F ( X r ε , Y r ε ) , D x ψ ( X r ε , Y r ε ) d r q + C 3 ε q E s t ( A ) γ e ( t r ) A · 1 2 T r D x 2 ψ ( X r ε , Y r ε ) Q 1 d r q = : i = 1 5 J 3 , i ( t , s , ε ) .
According to Theorem 2, Lemmas 1 and 5, and Minkowski’s inequality, we obtain
J 3 , 1 ( t , s , ε ) C 3 , 1 ε q E s t ( t s ) γ A X r ε , D x ψ ( X r ε , Y r ε ) d r q C 3 , 1 ε q s t ( t r ) γ E A X r ε q 1 q d r q C 3 , 1 ε q s t ( t r ) γ r 1 + 3 l η 2 η + ε l 2 d r q C 3 , 1 ( t s ) ( l 2 γ ) q ε ( 1 l 2 ) q C 3 , 1 ( t s ) ( l 2 γ ) q ε q 2 .
By Theorem 2, (19), and Minkowski’s inequality, we have
J 3 , 2 ( t , s , ε ) C 3 , 2 ε q E s t ( t r ) γ B ( X r ε ) , D x ψ ( X r ε , Y r ε ) d r q C 3 , 2 ε q s t ( t r ) γ E X r ε 1 2 q 1 / q d r q C 3 , 2 ε q s t ( t r ) γ r ( η l ) / η d r q C 3 , 2 ( t s ) ( l / 2 γ ) q ε q .
Using Theorem 2, (19), Lemma 1, the Gagliardo–Nirenberg inequality, and Minkowski’s inequality, we obtain
J 3 , 3 ( t , s , ε ) C 3 , 3 ε q E s t ( t r ) γ E ( X r ε ) , D x φ ( X r ε , Y r ε ) d r q C 3 , 3 ε q E s t ( t r ) γ X r ε 3 d r q C 3 , 3 ε q E s t ( t r ) γ X r ε L 6 3 d r q C 3 , 3 ε q s t ( t r ) γ E X r ε 1 3 q 1 / q d r q C 3 , 3 ε q s t ( t r ) γ r 3 ( η l ) / 2 η d r q C 3 , 3 ( t s ) ( l / 2 γ ) q ε q .
According to Theorem 2 and Lemma 1, we obtain
J 3 , 4 ( t , s , ε ) C 3 , 4 ε q E s t ( t r ) γ F ( X r ε , Y r ε ) , D x ψ ( X r ε , Y r ε ) d r q C 3 , 4 ( t s ) ( l / 2 γ ) q ε q .
For the last term, by (18), Lemma 1, and assumption (A2), we get
J 3 , 5 ( t , s , ε ) C 3 , 5 ε q E s t ( t r ) γ · 1 2 T r ( Q 1 ) D x 2 ψ ( X r ε , Y r ε ) d r q C 3 , 5 ε q s t ( t r ) γ E ( 1 + X r ε q + Y r ε q ) 1 / q d r q C 3 , 5 ( t s ) ( l / 2 γ ) q ε q .
According to Theorem 2, Burkholder–Davis–Gundy’s inequality, and assumption (A2), we have
J 4 ( t , s , ε ) C 4 ε q s t E ( A ) γ e ( t r ) A D x ψ ( X r ε , Y r ε ) Q 1 1 / 2 L 2 ( H ) 2 d r q / 2 C 4 ( t s ) ( 1 / 2 γ ) q ε q C 4 ( t s ) ( l / 2 γ ) q ε q ,
and similarly
J 5 ( t , s , ε ) C 5 ( t s ) ( 1 / 2 γ ) q ε q C 5 ( t s ) ( l / 2 γ ) q ε q .
Thus, the proof is finished. □
Now, we are in a position to provide the following:
Proof of Theorem 1. 
Let T > 0 . By (8) and (13), t [ 0 , T ] , so we have
( A ) γ ( X t ε X ¯ t ) = 0 t ( A ) γ e ( t s ) A B ( X s ε ) B ( X ¯ s ) d s + 0 t ( A ) γ e ( t s ) A E ( X s ε ) E ( X ¯ s ) d s + 0 t ( A ) γ e ( t s ) A F ¯ ( X s ε ) F ¯ ( X ¯ s ) d s + 0 t ( A ) γ e ( t s ) A δ F ( X s ε , Y s ε ) d s ,
where
δ F ( x , y ) : = F ( x , y ) F ¯ ( x ) .
Thus, q 1 , and we have
( A ) γ ( X t ε X ¯ t ) q C q 0 t ( A ) γ e ( t s ) A B ( X s ε ) B ( X ¯ s ) d s q + C q 0 t ( A ) γ e ( t s ) A E ( X s ε ) E ( X ¯ s ) d s q + C q 0 t ( A ) γ e ( t s ) A F ¯ ( X s ε ) F ¯ ( X ¯ s ) d s q + C q 0 t ( A ) γ e ( t s ) A δ F ( X s ε , Y s ε ) d s q = : i = 1 4 Q i ( t , ε ) .
For the first term, by Corollary 2.2. in [41], we have
Q 1 ( t , ε ) C 1 0 t ( t s ) ( γ + 1 2 ) B ( X s ε ) B ( X ¯ s ) 1 d s q C 1 0 t ( t s ) ( γ + 1 2 ) X s ε X ¯ s X s ε 1 + X ¯ s 1 d s q C 1 0 t ( t s ) ( γ + 1 2 ) · q q 1 d s q 1 · 0 t X s ε X ¯ s q X s ε 1 + X ¯ s 1 q d s C 1 0 t X s ε X ¯ s q ( X s ε 1 q + X ¯ s 1 q ) d s ,
where we hold that, if q > 2 , q ( 2 γ + 1 ) 2 ( q 1 ) > 1 , and then
0 t ( t s ) q ( 2 γ + 1 ) 2 ( q 1 ) d s C T .
According to Hölder’s inequality, we deduce
Q 2 ( t , ε ) C 2 0 t ( A ) γ e ( t s ) A E ( X s ε ) E ( X ¯ s ) d s q C 2 0 t ( t s ) γ X s ε 3 + X ¯ s 3 d s q C 2 0 t ( t s ) γ X s ε X ¯ s X s ε 2 + X s ε X ¯ s + X ¯ s 2 L d s q C 2 0 t ( t s ) γ X s ε X ¯ s X s ε 1 2 + X ¯ s 1 2 d s q C 2 0 t X s ε X ¯ s q X s ε 1 2 q + X ¯ s 1 2 q d s .
For the third term, by Theorem 2, it is evident that
Q 3 ( t , ε ) C 3 0 t ( t s ) γ F ¯ ( X s ε ) F ¯ ( X ¯ s ) d s q C 3 0 t ( t s ) γ X s ε X ¯ s d s q C 3 0 t X s ε X ¯ s q d s .
Combining the above computations, we have
( A ) γ ( X t ε X ¯ t ) q C 4 0 t X s ε X ¯ s q 1 + X s ε 1 2 q + X ¯ s 1 2 q d s + C 4 0 t ( A ) γ e ( t s ) A δ F ( X s ε , Y s ε ) d s q .
Then, by Gronwall’s inequality, we get
E ( A ) γ ( X t ε X ¯ t ) q C 5 E 0 t ( A ) γ e ( t s ) A δ F ( X s ε , Y s ε ) d s 2 q 1 / 2 · E e C 6 0 t ( 1 + X s ε 1 2 q + X ¯ s 1 2 q ) d s 1 / 2 .
According to Lemma 1.1 in [42], one can check that
sup t [ 0 , T ] E e C 0 t ( X s ε 1 2 q + X ¯ s 1 2 q ) d s C T .
Since δ F ( x , y ) satisfies the centering condition (17), using Lemma 6, we can directly obtain
E 0 t ( A ) γ e ( t s ) A δ F ( X s ε , Y s ε ) d s 2 q 1 / 2 C 7 ε q / 2 .
Thus, we obtain the desired result. □

5. Example

Consider a practical application: let f ( x , y ) = x + y , g ( x , y ) = y in (1), and then we have
𝜕 X t ε ( ξ ) 𝜕 t = Δ X t ε ( ξ ) + 1 2 𝜕 𝜕 ξ ( X t ε ( ξ ) ) 2 X t ε 3 ( ξ ) + X t ε ( ξ ) + Y t ε ( ξ ) + 𝜕 W t 1 ( ξ ) 𝜕 t , 𝜕 Y t ε ( ξ ) 𝜕 t = 1 ε Δ Y t ε ( ξ ) 1 ε Y t ε ( ξ ) + 1 ε 𝜕 W t 2 ( ξ ) 𝜕 t , X t ε ( 0 ) = X t ε ( 1 ) = Y t ε ( 0 ) = Y t ε ( 1 ) = 0 , t ( 0 , T ] , X 0 ε ( ξ ) = x , Y 0 ε ( ξ ) = y , ξ ( 0 , 1 ) ,
where
  • X t ε represents the concentration of a slowly varying chemical substance;
  • Y t ε represents the concentration of a rapidly varying chemical substance;
  • the nonlinear term X t ε 3 represents self-inhibition;
  • The term 1 2 𝜕 𝜕 ξ ( X t ε ( ξ ) ) 2 represents a convective effect;
  • The functions f ( x , y ) = x + y and g ( x , y ) = y represent the reaction coupling.
It is easy to check that f ( x , y ) = x + y and g ( x , y ) = y satisfy the Lipschitz condition and our assumption (A1). Note that, in this special case, the frozen equation is
𝜕 Y t ( ξ ) 𝜕 t = Δ Y t ( ξ ) Y t ( ξ ) + 𝜕 W t 2 ( ξ ) 𝜕 t ,
which has a unique invariant measure μ . Moreover, we can have exactly
f ¯ ( x ) = R ( x + y ) μ ( d y ) = x + 0 = x .
Thus, by our result regarding Theorem 1, we determine that X t ε converges strongly (with the optimal convergence rate 1 / 2 ) to X ¯ t , thereby satisfying
𝜕 X ¯ t ( ξ ) 𝜕 t = Δ X ¯ t ( ξ ) + 1 2 𝜕 𝜕 ξ ( X ¯ ( ξ ) ) 2 X ¯ t 3 ( ξ ) + X ¯ t ( ξ ) + 𝜕 W t 1 ( ξ ) 𝜕 t .
From the above example, we can see that, as ε approaches zero, the original fast–slow system (26) converges strongly (with the optimal convergence rate 1 / 2 ) to an averaged system (28), which no longer depends on the fast variable Y t ε . Intuitively, the averaged system is much simpler than the original one. When ε is small, directly simulating the original system numerically requires extremely small time steps. In contrast, simulating the effective averaged equation significantly improves computational efficiency; see [43] and the references therein for the Heterogeneous Multiscale Methods (simulating the effective averaged equation instead of the slow–fast system) and studies on numerical analysis regarding slow–fast SPDEs based on the optimal strong convergence in the averaging principle. The main application of the optimal rates of convergence obtained in the present manuscript is the construction of efficient numerical schemes for the slow–fast system based on HMMs. However, obtaining the explicit invariant measure μ x for the averaged Equation (13) is highly challenging, primarily because of the complexity in the dynamics of the frozen Equation (12). Hence, the explicit form of the averaged Equation (13) is generally inaccessible, making it difficult for us to intuitively see how the approximate solution of the averaged equation approximates the solution of the original system. We chose Example (26) because it directly validates our hypothesis and yields the explicit form of the averaged Equation (27) thanks to the explicit form of the invariant measure μ of the frozen Equation (27). For the averaged Equation (27), we shall simulate it via the general solver for SPDEs. This clearly demonstrates that the averaged equation is significantly simpler, both in terms of numerical simulation and theoretical analysis, compared to the original fast–slow system (26).

6. Conclusions

This work establishes a framework for analyzing multiscale stochastic reaction–diffusion–advection systems with Hölder-continuous coefficients. Compared with the existing work [30], we obtain the optimal strong convergence rate in the averaging principle, which is primarily motivated by developing efficient numerical schemes using Heterogeneous Multiscale Methods (HMMs), as detailed in [43] and the references therein. Thus, the optimal convergence rate established in our work is crucial for numerical approximation, which shall be studied in future work.

Author Contributions

Methodology, L.Y. and L.L.; writing—original draft preparation, L.Y. and L.L.; writing—review and editing, L.Y. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NNSF OF China (Nos. 12301179 and 12471468).

Data Availability Statement

The original contributions presented in this study are included in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Proof of Lemma 2. 
Recall that
X t ε = e t A x + 0 t e ( t s ) A B ( X s ε ) d s + 0 t e ( t s ) A E ( X s ε ) d s + 0 t e ( t s ) A F ( X s ε , Y s ε ) d s + 0 t e ( t s ) A d W s 1 .
Using (3), we have
E e t A x η C t η l 2 e λ 1 t 2 x l .
According to (3) and Lemma 2.1 in [44], for any η ( 1 , 3 / 2 ) , we have
0 t e ( t s ) A B ( X s ε ) d s η 0 t e ( t s ) A B ( X s ε ) η d s C 0 t ( t s ) η ( τ 3 ) 2 B ( X s ε ) τ 3 d s C 0 t ( t s ) η + τ 3 2 X s ε τ 1 X s ε τ 2 + 1 d s ,
where we choose positive constants τ 1 and τ 2 such that they are small enough to also satisfy 1 + τ 1 + τ 2 ( 1 , η ) , and then find a proper τ 3 [ 1 / 2 , 2 η ) that satisfies τ 1 + τ 2 + τ 3 > 1 / 2 . According to interpolation inequality, for any 0 < τ 1 < η , we obtain
X s ε τ 1 C X s ε 1 τ 1 η X s ε η τ 1 η ,
and, for any 0 < τ 2 + 1 < η , we have
X s ε τ 2 + 1 C X s ε 1 τ 2 + 1 η X s ε η τ 2 + 1 η .
By Minkowski’s inequality, (A1), and (A2), there exists m > 1 such that we have
E 0 t e ( t s ) A B ( X s ε ) d s η q E 0 t ( t s ) η + τ 3 2 X s ε 2 τ 1 + τ 2 + 1 η X s ε η τ 1 + τ 2 + 1 η d s q C E 0 t ( t s ) η + τ 3 2 · q q 1 d s q 1 · 0 t X s ε q · ( 2 τ 1 + τ 2 + 1 η ) X s ε η q ( τ 1 + τ 2 + 1 ) η d s C E 0 t X s ε q · ( 2 τ 1 + τ 2 + 1 η ) X s ε η q ( τ 1 + τ 2 + 1 ) η d s C 0 t E X s ε q ( 2 η τ 1 τ 2 1 ) η τ 1 τ 2 1 d s + 0 t E X s ε η q d s C ( 1 + x m q ) + C 0 t E X s ε η q d s .
According to Gagliardo–Nirenberg inequality, we obtain
0 t e ( t s ) A E ( X s ε ) d s η 0 t e ( t s ) A E ( X s ε ) η d s C 0 t ( t s ) 1 2 · 2 η + 5 4 X s ε 3 2 η 5 4 d s C 0 t ( t s ) 2 η + 5 8 X s ε L 3 3 d s C 0 t ( t s ) 2 η + 5 8 X s ε η 1 2 X s ε 5 2 d s ,
so, for the third term, we have
E 0 t e ( t s ) A E ( X s ε ) d s η q C E 0 t ( t s ) 2 η + 5 8 X s ε η 1 2 X s ε 5 2 d s q C E 0 t ( t s ) 2 η + 5 8 · q q 1 d s q 1 · 0 t X s ε η q 2 X s ε 5 q 2 d s C E 0 t X s ε η q 2 X s ε 5 q 2 d s C 0 t E X s ε η q d s + 0 t E X s ε 5 q d s C ( 1 + x 5 q ) + C 0 t E X s ε η q d s .
According to (3) and Minkowski’s inequality, we get
E 0 t e ( t s ) A F ( X s ε , Y s ε ) d s η q E 0 t e ( t s ) A F ( X s ε , Y s ε ) η d s q C 0 t ( t s ) η / 2 E F ( X s ε , Y s ε ) q 1 / q d s q C .
Using the assumption (A2), we have
E 0 t e ( t s ) A d W s 1 η q = E 0 t ( A ) η / 2 e ( t s ) A d W s 1 q C 0 t ( A ) η / 2 e ( t s ) A Q 1 1 / 2 L 2 ( H ) 2 d s q / 2 C .
Hence, according to the above estimates, for some m > 1 , we have
E X s ε η q C 0 t E X s ε η q d s + C t ( η l ) q 2 e λ 1 q t 2 x l q + x m q + 1 C 0 t E X s ε η q d s + C t ( η l ) q 2 e λ 1 q t 2 x l q + x m q + 1 .
Applying Gronwall’s inequality, we obtain the desired result. □
Proof of Lemma 3. 
The estimate (21) comes from Lemma 4.3 in [27]. To prove (20), note that
X t ε X s ε = ( e t A e s A ) x + s t e ( t r ) A B ( X r ε ) d r + 0 s ( e ( t r ) A e ( s r ) A ) B ( X r ε ) d r + s t e ( t r ) A E ( X r ε ) d r + 0 s ( e ( t r ) A e ( s r ) A ) E ( X r ε ) d r + s t e ( t r ) A F ( X r ε , Y r ε ) d r + 0 s ( e ( t r ) A e ( s r ) A ) F ( X r ε , Y r ε ) d r + s t e ( t r ) A d W r 1 + 0 s ( e ( t r ) A e ( s r ) A ) d W r 1 = : i = 1 9 X i ( t , s ) .
Applying (3) and (4), we have
E X 1 ( t , s ) C 1 ( t s ) α 2 s α l 2 e λ 1 s 2 x l .
Using (3), (19), Minkowski’s inequality, and Corollary 2.1 in [41], for any q > 1 , we obtain
E X 2 ( t , s ) q C 2 s t E B ( X r ε ) q 1 q d r q C 2 s t E X r ε 1 2 q 1 q d r q C 2 s t r η l η e λ 1 r 2 η x l 1 η + x m η + 1 2 d r q C 2 ( t s ) q .
For the third term, using (3), (4), (19), Minkowski’s inequality, and Corollary 2.1 in [41], we have
E X 3 ( t , s ) q 0 s E ( e ( t s ) A I ) e ( s r ) A B ( X r ε ) q 1 q d r q C 3 ( t s ) q 2 0 s E ( A ) 1 2 e ( s r ) A B ( X r ε ) q 1 q d r q C 3 ( t s ) q 2 0 s ( s r ) 1 2 E X r ε 1 2 q 1 q d r q C 3 ( t s ) q 2 0 s ( s r ) 1 2 r η l η e λ 1 r 2 η x l 1 η + x m η + 1 2 d r q C 3 ( t s ) q 2 .
Applying the Gagliardo–Nirenberg inequality, Minkowski’s inequality, (3), and (19), for l > 1 / 2 , we obtain
E X 4 ( t , s ) q C 4 s t E X r ε L 6 3 q 1 q d r q C 4 s t E X r ε 1 3 q 1 q d r q C 4 s t r 3 ( η l ) 2 η e λ 1 r 2 η x l 1 η + x m η + 1 3 d r q C 4 ( t s ) q .
Using (3), (4), (19), and Minkowski’s inequality, for l > 1 / 2 , we have
E X 5 ( t , s ) q 0 s E ( e ( t s ) A I ) e ( s r ) A X r ε 3 q 1 q d r q C 5 ( t s ) q 2 0 s E ( A ) 1 2 e ( s r ) A X r ε 3 q 1 q d r q C 5 ( t s ) q 2 0 s ( s r ) 1 2 E X r ε 1 3 q 1 q d r q C 5 ( t s ) q 2 0 s ( s r ) 1 2 r 3 ( η l ) 2 η e λ 1 r 2 η x l 1 η + x m η + 1 3 d r q C 5 ( t s ) q 2 .
By Minkowski’s inequality, we deduce that
E X 6 ( t , s ) q s t E e ( t r ) A F ( X r ε , Y r ε ) q 1 / q d r q C 6 ( t s ) q .
Similarly, using (3), (4), and Minkowski’s inequality, we have
E X 7 ( t , s ) q 0 s E ( e ( t s ) A I ) e ( s r ) A F ( X r ε , Y r ε ) q 1 / q d r q C 7 ( t s ) q / 2 0 s E ( A ) 1 / 2 e ( s r ) A F ( X r ε , Y r ε ) q 1 / q d r q C 7 ( t s ) q / 2 .
According to Burkholder–Davis–Gundy’s inequality and the assumption (A2), we further get
E X 8 ( t , s ) q C 8 s t e ( t r ) A Q 1 1 / 2 L 2 ( H ) 2 d r q / 2 C 8 ( t s ) q / 2 ,
and
E X 9 ( t , s ) q C 9 ( t s ) q / 2 0 s ( A ) 1 / 2 e ( s r ) A Q 1 1 / 2 L 2 ( H ) 2 d r q / 2 C 9 ( t s ) q / 2 .
Combining the above computations, we obtain the desired result (20). □
Proof of Lemma 5. 
We rewrite
X t ε = e t A x + 0 t e ( t s ) A B ( X s ε ) d s + 0 t e ( t s ) A E ( X s ε ) d s + 0 t e ( t s ) A F ( X s ε , Y s ε ) d s + 0 t e ( t s ) A d W s 1 = e t A x + 0 t e ( t s ) A B ( X t ε ) d s + 0 t e ( t s ) A ( B ( X s ε ) B ( X t ε ) ) d s + 0 t e ( t s ) A E ( X t ε ) d s + 0 t e ( t s ) A ( E ( X s ε ) E ( X t ε ) ) d s + 0 t e ( t s ) A F ( X t ε , Y t ε ) d s + 0 t e ( t s ) A ( F ( X s ε , Y s ε ) F ( X t ε , Y t ε ) ) d s + 0 t e ( t s ) A d W s 1 .
Take the expectation on both sides and we get
E X t ε 2 q C q e t A x 2 + C q E 0 t e ( t s ) A B ( X t ε ) d s 2 q + C q E 0 t e ( t s ) A ( B ( X s ε ) B ( X t ε ) ) d s 2 q + C q E 0 t e ( t s ) A E ( X t ε ) d s 2 q + C q E 0 t e ( t s ) A ( E ( X s ε ) E ( X t ε ) ) d s 2 q + C q E 0 t e ( t s ) A F ( X t ε , Y t ε ) d s 2 q + C q E 0 t e ( t s ) A ( F ( X s ε , Y s ε ) F ( X t ε , Y t ε ) ) d s 2 q + C q E 0 t e ( t s ) A d W s 1 2 q = : i = 1 8 R i ( T , ε ) .
From (3), we can infer
R 1 ( T , ε ) C 1 t 1 + l 2 e λ 1 t 2 x l .
According to (19) and Lemma 2, we have
R 2 ( T , ε ) = E 0 t ( A ) e ( t s ) A B ( X t ε ) d s q = E ( e t A I ) B ( X t ε ) q C 2 E X t ε 1 2 X t ε η q C 2 t η l 4 η t η l 2 e λ 1 t 4 η x l 1 2 η + x m 2 η + 1 e λ 1 t 2 x l + x m + 1 q C 2 t q + q ( 3 l η ) 2 η x l + x m + 1 2 q .
For the third term, we rewrite
0 t e ( t s ) A ( B ( X s ε ) B ( X t ε ) ) d s 2 = 0 t ( A ) e ( t s ) A ( B ( X s ε X t ε , X t ε ) + B ( X s ε , X s ε X t ε ) ) d s C 3 0 t ( t s ) 1 B ( X s ε X t ε , X t ε ) d s + C 3 0 t ( t s ) 1 B ( X s ε , X s ε X t ε ) d s .
From Minkowski’s inequality, interpolation inequality, (22), and (19), it follows that, for small enough δ > 0 ,
R 3 ( T , ε ) C 3 0 t ( t s ) 1 E X s ε X t ε 1 2 1 2 E X t ε 1 2 1 2 d s q + C 3 0 t ( t s ) 1 E X s ε 1 / 2 + δ 2 1 2 E X s ε X t ε 1 2 1 2 d s q C 3 0 t ( t s ) η 3 2 t 1 2 s η l 2 e λ 1 s 2 x l + x m + 1 d s q + C 3 0 t ( t s ) η 3 2 s 1 4 δ 2 η l 2 e λ 1 s 2 x l + x m + 1 d s q C 3 t q + q ( 3 l η ) 2 η x l + x m + 1 q .
By (4) and (19), we obtain
R 4 ( T , ε ) = E 0 t ( A ) e ( t s ) A ( X t ε 3 ) d s q C 4 E ( e t A I ) ( X t ε 3 ) q C 4 E X t ε 1 3 q C 4 t 3 q ( η l ) 2 η e λ 1 t 2 η x l 1 η + x m η + 1 3 q .
With Hölder’s inequality, Minkowski’s inequality, Lemma 3, and (19), we have
R 5 ( T , ε ) = E 0 t ( A ) e ( t s ) A ( X t ε 3 X s ε 3 ) d s q C 5 E 0 t ( t s ) 1 X t ε X s ε ( X t ε 1 2 + X s ε 1 2 ) d s q C 5 0 t ( t s ) 1 E X t ε X s ε 2 1 / 2 E ( X t ε 1 2 + X s ε 1 2 ) 2 1 / 2 d s q C 5 ( 0 t ( t s ) 1 ( t s ) α 2 s α l 2 e λ 1 s 2 x l + ( t s ) 1 2 ( t η l η ( e λ 1 t 2 η x l 1 η + x m η + 1 ) 2 + s η l η ( e λ 1 s 2 η x l 1 η + x m η + 1 ) 2 ) d s ) q C 5 t q + q ( 3 l η ) 2 η x l + x m + 1 2 q .
Using (4), Minkowski’s inequality, and Lemma 1, we have
R 6 ( t , ε ) C 6 E ( e t A I ) F ( X t ε , Y t ε ) q C 6 .
Furthermore, by applying (20) and (21) with l = 0 , we obtain, for β ( 0 , 1 ] ,
E R 7 ( t , ε ) q C 7 ( 0 t ( t s ) 1 [ E X t ε X s ε 2 q 1 2 q + E Y t ε Y s ε 2 θ q 1 2 q ] d s ) q C 7 ( y + x + 1 ) 0 t ( t s ) θ β / 2 1 1 s θ β / 2 + 1 ε θ β / 2 d s q C 7 ε q β / 2 ( y + x + 1 ) .
For the last term, by the assumption (A2), we have
R 8 ( T , ε ) = E 0 t ( A ) e ( t s ) A d W s 1 q C 8 E 0 t ( A ) e ( t s ) A Q 1 1 / 2 L 2 ( H ) 2 d s q / 2 C 8 .
Combining the above computations, we obtain the desired result. □

References

  1. Zhang, T.; Yang, Y.; Han, S. Exponential heterogeneous anti-synchronization of multi-variable discrete stochastic inertial neural networks with adaptive corrective parameter. Eng. Appl. Artif. Intell. 2025, 142, 109871. [Google Scholar] [CrossRef]
  2. Cantrell, R.S.; Cosner, C.; Lou, Y. Approximating the ideal free distribution via reaction–diffusion–advection equations. J. Differ. Equ. 2008, 245, 3687–3703. [Google Scholar] [CrossRef]
  3. Du, Y.; Mei, L. On a nonlocal reaction-diffusion-advection equation modelling phytoplankton dynamics. Nonlinearity 2010, 24, 319. [Google Scholar] [CrossRef]
  4. Peng, R.; Zhao, X.Q. A nonlocal and periodic reaction-diffusion-advection model of a single phytoplankton species. J. Math. Bio. 2016, 72, 755–791. [Google Scholar] [CrossRef]
  5. Chen, X.; Hambrock, R.; Lou, Y. Evolution of conditional dispersa: A reaction-diffusion-advection model. J. Math. Biol. 2008, 57, 361–386. [Google Scholar] [CrossRef] [PubMed]
  6. Rubin, J.; Wechselberger, M. Giant squid-hidden canard: The 3D geometry of the Hodgkin-Huxley model. Bio. Cyber. 2007, 97, 5–32. [Google Scholar] [CrossRef]
  7. Huang, Z.L.; Zhu, W.Q. Stochastic averaging of quasi-generalized Hamiltonian systems. Int. J. -Non-Linear Mech. 2009, 44, 71–80. [Google Scholar] [CrossRef]
  8. Roose, D.; Szalai, R. Numerical Continuation Methods for Dynamical Systems; Springer: Dordrech, The Netherlands, 2007. [Google Scholar]
  9. Bogoliubov, N.N.; Mitropolsky, Y.A. Asymptotic Methods in the Theory of Non-Linear Oscillations; Gordon and Breach Science Publishers: New York, NY, USA, 1961. [Google Scholar] [CrossRef]
  10. Khasminskii, R.Z. On stochastic processes defined by differential equations with a small parameter. Theor. Probab. Appl. 1966, 11, 211–228. [Google Scholar] [CrossRef]
  11. Bakhtin, V.; Kifer, Y. Diffusion approximation for slow motion in fully coupled averaging. Probab. Theory Relat. Fields 2004, 129, 157–181. [Google Scholar] [CrossRef]
  12. Gonzales-Gargate, I.I.; Ruffino, P.R. An averaging principle for diffusions in foliated spaces. Ann. Probab. 2016, 44, 567–588. [Google Scholar] [CrossRef]
  13. Hairer, M.; Pardoux, E. Averaging dynamics driven by fractional Brownian motion. Ann. Probab. 2020, 48, 1826–1860. [Google Scholar] [CrossRef]
  14. Khasminskii, R.Z.; Yin, G. On averaging principles: An asymptotic expansion approach. SIAM J. Math. Anal. 2004, 35, 1534–1560. [Google Scholar] [CrossRef]
  15. Li, X.M. An averaging principle for a completely integrable stochastic Hamiltonian system. Nonlinearity 2008, 21, 803–822. [Google Scholar] [CrossRef]
  16. Li, X.M.; Sieber, J. Slow-fast systems with fractional environment and dynamics. Ann. Appl. Probab. 2022, 32, 3964–4003. [Google Scholar] [CrossRef]
  17. Veretennikov, A.Y. On the averaging principle for systems of stochastic differential equations. Math. USSR Sborn. 1991, 69, 271–284. [Google Scholar] [CrossRef]
  18. Cerrai, S.; Freidlin, M. Averaging principle for a class of stochastic reaction-diffusion equations. Probab. Theory Relat. Fields 2009, 144, 137–177. [Google Scholar] [CrossRef]
  19. Cerrai, S. A Khasminskii type averaging principle for stochastic reaction-diffusion equations. Ann. Appl. Probab. 2009, 19, 899–948. [Google Scholar] [CrossRef]
  20. Cerrai, S. Averaging principle for systems of reaction-diffusion equations with polynomial nonlinearities perturbed by multiplicative noise. SIAM J. Math. Anal. 2011, 43, 2482–2518. [Google Scholar] [CrossRef]
  21. Bao, J.; Yin, G.; Yuan, C. Two-time-scale stochastic partial differential equations driven by α-stable noises: Averaging principles. Bernoulli 2017, 23, 645–669. [Google Scholar] [CrossRef]
  22. Bréhier, C.E. Uniform weak error estimates for an asymptotic preserving scheme applied to a class of slow-fast parabolic semilinear SPDEs. J. Comput. Math. 2024, 10, 175–228. [Google Scholar] [CrossRef]
  23. Bréhier, C.E. Strong and weak orders in averaging for SPDEs. Stoch. Process Appl. 2012, 122, 2553–2593. [Google Scholar] [CrossRef]
  24. Cerrai, S.; Lunardi, A. Averaging principle for non-autonomous slow-fast systems of stochastic reaction-diffusion equations: The almost periodic case. SIAM J. Math. Anal. 2017, 49, 2843–2884. [Google Scholar] [CrossRef]
  25. Cerrai, S.; Zhu, Y. Averaging principle for slow-fast systems of stochastic PDEs with rough coefficients. Stoch. Proc. Appl. 2025, 185, 104618. [Google Scholar] [CrossRef]
  26. Han, M.; Pei, B. An averaging principle for stochastic evolution equations with jumps and random time delays. AIMS Math. 2021, 6, 39–51. [Google Scholar] [CrossRef]
  27. Röckner, M.; Xie, L.; Yang, L. Asymptotic behavior of multiscale stochastic partial differential equations with Hölder coefficients. J. Funct. Anal. 2023, 285, 110103. [Google Scholar] [CrossRef]
  28. Wang, W.; Roberts, A.J. Average and deviation for slow-fast stochastic partial differential equations. J. Differ. Equ. 2012, 253, 1265–1286. [Google Scholar] [CrossRef]
  29. Dong, Z.; Sun, X.; Xiao, H.; Zhai, J. Averaging principle for one dimensional stochastic Burgers equation. J. Differ. Equ. 2018, 265, 4749–4797. [Google Scholar] [CrossRef]
  30. Gao, P. Averaging principle for multiscale stochastic reaction-diffusion-advection equations. Math. Method Appl. 2019, 42, 1122–1150. [Google Scholar] [CrossRef]
  31. Gao, P.; Sun, X. Optimal convergence order for multi-scale stochastic Burgers equation. Stochastics Partial. Differ. Equ. Anal. Comput. 2025, 13, 421–464. [Google Scholar] [CrossRef]
  32. Gao, P. Averaging principle for multiscale stochastic Klein-Gordon-heat system. J. Nonlinear Sci. 2019, 29, 1701–1759. [Google Scholar] [CrossRef]
  33. Li, S.; Xie, Y. Averaging principle for stochastic 3D fractional Leray-α model with a fast oscillation. Stoch. Anal. Appl. 2020, 38, 248–276. [Google Scholar] [CrossRef]
  34. Gao, P. Averaging principle for multiscale nonautonomous random 2D Navier-Stokes system. J. Funct. Anal. 2023, 285, 110036. [Google Scholar] [CrossRef]
  35. Liu, D. Strong convergence of principle of averaging for multiscale stochastic dynamical systems. Commun. Math. Sci. 2010, 8, 999–1020. [Google Scholar] [CrossRef]
  36. Bréhier, C.E. Orders of convergence in the averaging principle for SPDEs: The case of a stochastically forced slow component. Stoch. Proc. Appl. 2020, 130, 3325–3368. [Google Scholar] [CrossRef]
  37. Da Prato, G.; Flandoli, F. Pathwise uniqueness for a class of SDEs in Hilbert spaces and applications. J. Funct. Anal. 2010, 259, 243–267. [Google Scholar] [CrossRef]
  38. Chojnowska-Michalik, A.; Goldys, B. Existence, uniqueness and invariant measures for stochastic semilinear equations on Hilbert spaces. Probab. Theory Relat. Fields 1995, 102, 331–356. [Google Scholar] [CrossRef]
  39. Pardoux, E.; Veretennikov, A.Y. On the Poisson equation and diffusion approximation. I. Ann. Probab. 2001, 29, 1061–1085. [Google Scholar] [CrossRef]
  40. Pardoux, E.; Veretennikov, A.Y. On the Poisson equation and diffusion approximation 2. Ann. Probab. 2003, 31, 1166–1192. [Google Scholar] [CrossRef]
  41. Dong, Z.; Xu, T.G. One-dimensional stochastic Burgers equation driven by Lévy processes. J. Func. Anal. 2007, 243, 631–678. [Google Scholar] [CrossRef]
  42. Ge, Y.; Sun, X.; Xie, Y. Optimal convergence rates in the averaging principle for slow-fast SPDEs driven by multiplicative noise. Commun. Math. Stat. 2024, 1–50. [Google Scholar] [CrossRef]
  43. Bréhier, C.E. Analysis of an HMM time-discretization scheme for a system of stochastic PDEs. SIAM J. Numer. Anal. 2013, 51, 1185–1210. [Google Scholar] [CrossRef]
  44. Temam, R. Navier-Stokes Equations and Nonlinear Functional Analysis; Society for Industrial and Applied Mathematics: Philadelphia, PA, USA, 1995. [Google Scholar] [CrossRef]
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

Yang, L.; Liu, L. Optimal Convergence of Slow–Fast Stochastic Reaction–Diffusion–Advection Equation with Hölder-Continuous Coefficients. Mathematics 2025, 13, 2550. https://doi.org/10.3390/math13162550

AMA Style

Yang L, Liu L. Optimal Convergence of Slow–Fast Stochastic Reaction–Diffusion–Advection Equation with Hölder-Continuous Coefficients. Mathematics. 2025; 13(16):2550. https://doi.org/10.3390/math13162550

Chicago/Turabian Style

Yang, Li, and Lin Liu. 2025. "Optimal Convergence of Slow–Fast Stochastic Reaction–Diffusion–Advection Equation with Hölder-Continuous Coefficients" Mathematics 13, no. 16: 2550. https://doi.org/10.3390/math13162550

APA Style

Yang, L., & Liu, L. (2025). Optimal Convergence of Slow–Fast Stochastic Reaction–Diffusion–Advection Equation with Hölder-Continuous Coefficients. Mathematics, 13(16), 2550. https://doi.org/10.3390/math13162550

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