In this paper, we examined the transport-diffusion equation in Rd, where the diffusion is represented by the Laplace operator multiplied by a function κ(t) dependent on time. We transformed the equation using the inverse function of s(t)=∫t0κ(t′)dt′. This transformation allowed us to construct a family of approximate solutions by using the heat kernel and translation corresponding to the transport in each step of time discretization. We proved the uniform convergence of these approximate solutions and their first and second derivatives with respect to the spatial variables. We also showed that the limit function satisfies the transport-diffusion equation in the space Rd.
Citation: Lynda Taleb, Rabah Gherdaoui. Approximation by the heat kernel of the solution to the transport-diffusion equation with the time-dependent diffusion coefficient[J]. AIMS Mathematics, 2025, 10(2): 2392-2412. doi: 10.3934/math.2025111
[1] | Yingchao Zhang, Yingzhen Lin . An ε-approximation solution of time-fractional diffusion equations based on Legendre polynomials. AIMS Mathematics, 2024, 9(6): 16773-16789. doi: 10.3934/math.2024813 |
[2] | Zihan Yue, Wei Jiang, Boying Wu, Biao Zhang . A meshless method based on the Laplace transform for multi-term time-space fractional diffusion equation. AIMS Mathematics, 2024, 9(3): 7040-7062. doi: 10.3934/math.2024343 |
[3] | Jing Wang, Qiaozhen Ma, Wenxue Zhou . Attractor of the nonclassical diffusion equation with memory on time- dependent space. AIMS Mathematics, 2023, 8(6): 14820-14841. doi: 10.3934/math.2023757 |
[4] | Madeeha Tahir, Ayesha Naz, Muhammad Imran, Hasan Waqas, Ali Akgül, Hussein Shanak, Rabab Jarrar, Jihad Asad . Activation energy impact on unsteady Bio-convection nanomaterial flow over porous surface. AIMS Mathematics, 2022, 7(11): 19822-19845. doi: 10.3934/math.20221086 |
[5] | Abdon Atangana, Mekkaoui Toufik . A piecewise heat equation with constant and variable order coefficients: A new approach to capture crossover behaviors in heat diffusion. AIMS Mathematics, 2022, 7(5): 8374-8389. doi: 10.3934/math.2022467 |
[6] | Fethi Bouzeffour . Inversion formulas for space-fractional Bessel heat diffusion through Tikhonov regularization. AIMS Mathematics, 2024, 9(8): 20826-20842. doi: 10.3934/math.20241013 |
[7] | Ke Li, Yongqin Xie, Yong Ren, Jun Li . Pullback attractors for the nonclassical diffusion equations with memory in time-dependent spaces. AIMS Mathematics, 2023, 8(12): 30537-30561. doi: 10.3934/math.20231561 |
[8] | Siyu Tian, Boyu Liu, Wenyan Wang . A novel numerical scheme for reproducing kernel space of 2D fractional diffusion equations. AIMS Mathematics, 2023, 8(12): 29058-29072. doi: 10.3934/math.20231488 |
[9] | M. Sivakumar, M. Mallikarjuna, R. Senthamarai . A kinetic non-steady state analysis of immobilized enzyme systems with external mass transfer resistance. AIMS Mathematics, 2024, 9(7): 18083-18102. doi: 10.3934/math.2024882 |
[10] | Xiangtuan Xiong, Wanxia Shi, Xuemin Xue . Determination of three parameters in a time-space fractional diffusion equation. AIMS Mathematics, 2021, 6(6): 5909-5923. doi: 10.3934/math.2021350 |
In this paper, we examined the transport-diffusion equation in Rd, where the diffusion is represented by the Laplace operator multiplied by a function κ(t) dependent on time. We transformed the equation using the inverse function of s(t)=∫t0κ(t′)dt′. This transformation allowed us to construct a family of approximate solutions by using the heat kernel and translation corresponding to the transport in each step of time discretization. We proved the uniform convergence of these approximate solutions and their first and second derivatives with respect to the spatial variables. We also showed that the limit function satisfies the transport-diffusion equation in the space Rd.
As a parabolic equation, the transport-diffusion equation was treated in many studies and well-consolidated methods are known [7]. Concerning the behavior of the solution of the transport-diffusion equation when the diffusion coefficient tends to zero, results have been obtained using the stochastic representation of the solution of a parabolic equation [4]. However, these results are expressed in the language of probability theory, and it might not be straightforward to translate them into the language of mathematical analysis. Furthermore, in this framework, the treatment of non-linear terms is not easy, see [9,10,11].
In recent years, a method inspired by the idea of the stochastic representation of the solution but formulated in the language of mathematical analysis without using probability concepts has been proposed. First, the convergence of approximate solutions constructed by the heat kernel at each discrete time step toward the solution of the transport-diffusion equation in Rd with a constant diffusion coefficient was demonstrated [12,13]. Then, this result was generalized to the case of the equation considered in the half-space with homogeneous Dirichlet boundary condition [2,5] and with homogeneous Neumann boundary condition []. Furthermore, using the approximate solutions of this type, the convergence of the solution of the transport-diffusion equation in Rd toward that of the transport equation were shown [1,3]. On the other hand, in [8], approximate solutions for the transport-diffusion equation in Rd and their limit function are considered, and it was proved that the limit function belongs to the Hölder space corresponding to the regularity of the given functions and satisfies the equation.
In this paper, we consider the transport-diffusion equation in Rd where the diffusion is represented by the Laplace operator multiplied by a function κ(t). More precisely, we consider the equation
∂tu(t,x)+v(t,x)⋅∇u(t,x)=κ(t)Δu(t,x)+f(t,x,u(t,x)). |
Here and throughout, v⋅∇=∑di=1vi∂xi, and Δ=∑di=1∂2xi.
Let us recall that in [12] and [13], the family of approximate solutions was defined on the discretized time family {t[n]k}∞k=0, n=1,2,⋯,
0=t[n]0<t[n]1<⋯<t[n]k<t[n]k+1<⋯,t[n]k=k2−n, |
using the heat kernel, i.e., the fundamental solution of the heat equation
Θn(x)=1(4πκδn)d/2exp(−|x|24κδn),δn=12n=t[n]k−t[n]k−1, |
over each interval [t[n]k−1,t[n]k] of the time discretization. The specific property Θn(x)=(Θn+1∗Θn+1)(x) of Gaussian functions was the technical basis in the demonstration of the convergence of the approximate solutions. But, if κ varies with the time t, this technique cannot be applied directly.
To overcome this difficulty arising from the non-constancy of the coefficient κ(t), we transform the equation using the inverse function of the function s(t)=∫t0κ(t′)dt′. This allows us to construct a family of approximate solutions similarly to the research in [12,13]. However, to demonstrate their convergence, it is essential to examine carefully the inequalities used, in which the consequences of the transformation of the equation also step in. In what follows, we aim to identify a reasonably large class of functions κ(t) for which we can obtain the convergence of the approximate solutions to the solution of the transport-diffusion equation.
Let us consider the equation for the unknown function u(t,x):R+×Rd→R
∂tu(t,x)+v(t,x)⋅∇u(t,x)=κ(t)Δu(t,x)+f(t,x,u(t,x)), t>0, x∈Rd, | (2.1) |
where κ(t):R+→R+, v(t,x):R+×Rd→Rd, and f(t,x,u):R+×Rd×R→R are given functions. Eq (2.1) will be envisaged with the initial condition
u(0,x)=u0(x)x∈Rd. | (2.2) |
For the function κ(t), we assume that
κ(t)>0a.e. in R+, | (2.3) |
and for any sequence {[an,bn[}∞n=1 of disjoint intervals contained in R+, we have
∀ε>0 ∃δ>0 such that, if ∞∑n=1∫bnanκ(t)dt≤δ then ∞∑n=1(bn−an)≤ε, | (2.4) |
κ(⋅)∈L1loc(R+). | (2.5) |
The conditions (2.3) and (2.5) imply that the function
s(t)=∫t0κ(t′)dt′ | (2.6) |
is invertible. We also note that the condition (2.4) implies that the inverse function of s(t), denoted by t(s), is absolutely continuous. As for the conditions on the functions v(t,x) and f(t,x,u), we will specify them in the statement of the theorem.
Next, we will consider a family of approximate solutions for Eq (2.1). For their definition, we will use the notation
δn=2−n,n=1,2,⋯, | (2.7) |
and for each n, the function
Θn(x)=1(4πδn)d/2exp(−|x|24δn),x∈Rd. | (2.8) |
We also introduce the class of functions
Λ={φ:D→R, continuous,∞∑n=1λτ,n(φ)<∞,∀τ>0}, | (2.9) |
where D=R+×Rd or D=R+×Rd×R and
λτ,n(φ)=sup{|φ(r1,x)−φ(r2,x)|:r1,r2∈[0,τ],x∈Rd,|r1−r2|≤δn} | (2.10) |
if D=R+×Rd, and
λτ,n(φ)=sup{|φ(r1,x,u)−φ(r2,x,u)|:r1,r2∈[0,τ],x∈Rd,u∈R,|r1−r2|≤δn} | (2.11) |
if D=R+×Rd×R.
In this paper, we use the notations
Dνx=∂|ν|∂xν11⋯∂xνdd,Dνx,u=∂|ν|∂xν11⋯∂xνdd∂uνd+1, |
where
|ν|=d∑j=1νj if ν=(ν1,⋯,νd)∈Nd, |
|ν|=d+1∑j=1νj if ν=(ν1,⋯,νd,νd+1)∈Nd+1. |
Furthermore, we denote by Cb(D) the class of continuous and bounded functions defined on the domain D.
The general result of the present work is the following theorem.
Theorem 1. Suppose that the function κ(t) satisfies conditions (2.3)–(2.5), and the functions v(t,x) and f(t,x,u) (with the function κ(t)) satisfy the conditions:
1κ(t)Dνxv(t,x)∈Cb([0,τ]×Rd))∀ν∈Nd, |ν|≤3, ∀τ>0, | (2.12) |
1κ(t(s))Dνxv(t(s),x)∈Λ∀ν∈Nd, |ν|≤2, | (2.13) |
1κ(t)Dνx,uf(t,x,u)1+|u|∈Cb([0,τ]×Rd×R)∀ν∈Nd+1, |ν|≤3, ∀τ>0, | (2.14) |
1κ(t(s))Dνx,uf(t(s),x,u)∈Λ∀ν∈Nd+1, |ν|≤2, | (2.15) |
Dνxu0(x)∈Cb(Rd)∀ν∈Nd, |ν|≤3. | (2.16) |
(In (2.13) and (2.15), the functions are considered as functions of (s,x) and (s,x,u), respectively.)
If we define
s[n]k=kδn,t[n]k=t(s[n]k),n=1,2,…,k=0,1,2,…, | (2.17) |
then, for any τ>0, the functions u[n](t,x) defined by
u[n](t[n]0,x)=u0(x), | (2.18) |
u[n](t[n]k,x)=∫RdΘn(y)u[n](t[n]k−1,x−δn1κ(t[n]k)v(t[n]k,x)−y)dy+δn1κ(t[n]k−1)f(t[n]k−1,x,u[n](t[n]k−1,x)),k=1,2,⋯, | (2.19) |
u[n](t,x)=s[n]k−s(t)δnu[n](t[n]k−1,x)+s(t)−s[n]k−1δnu[n](t[n]k,x)for t[n]k−1≤t≤t[n]k | (2.20) |
(with s(t) and s[n]k defined by (2.6) and (2.17)), and their first and second derivatives with respect to x∈Rd, converge uniformly on [0,τ]×Rd toward a function u(t,x) and its first and second derivatives with respect to x∈Rd, and the limit function u(t,x) satisfy Eq (2.1) and the initial condition (2.2) in the sense of integral equality:
−∫∞0u(t,x)ddtφ(t)dt−u0(x)φ(0)+∫∞0v(t,x)⋅∇u(t,x)φ(t)dt=∫∞0(κ(t)Δu(t,x)+f(t,x,u))φ(t)dt | (2.21) |
for any φ(⋅)∈C1(R+) such that φ(t)=0fort≥τ1 with τ1>0.
Since the conditions (2.13) and (2.15) are formulated through the inverse function t(s), to have a more concrete idea, we mention a class of functions (κ(t),v(t,x),f(t,x,u)) that satisfies the conditions (2.13) and (2.15).
Lemma 2. Suppose that the function κ(t) satisfies the conditions (2.3) and (2.5), and for each τ>0, there exists a number α=α(τ), 0<α<1, and a constant C1=C1(τ) such that
t2−t1≤C1(∫t2t1κ(t)dt)α∀t1,t2∈[0,τ], t1<t2. | (3.1) |
Furthermore, suppose that for each τ>0, there exist numbers β=β(τ) and γ=γ(τ) such that 0<β<1 and 0<γ<1, and the functions v(t,x) and f(t,x,u) satisfy the relations:
sup0≤t1<t2≤τ,x∈Rd|Dνxv(t2,x)−Dνxv(t2,x)|(t2−t1)β<∞,∀ν∈Nd, |ν|≤2, | (3.2) |
sup0≤t1<t2≤τ,x∈Rd,u∈R|Dνx,uf(t2,x,u)−Dνxf(t2,x,u)|(t2−t1)γ<∞,∀ν∈Nd+1, |ν|≤2. | (3.3) |
Then the functions v(t(s),x) and f(t(s),x,u) (with the function κ(t(s))) satisfy the conditions (2.13) and (2.15).
Proof. Fix τ>0. From (3.1) and (3.2), we deduce that there exists a constant C2=C2(τ) such that
λτ,n(1κ(t(⋅))Dνxv(t(⋅),⋅))≤C2sup0≤t1<t2≤τ,s(t2)−s(t1)≤δn|t2−t1|β≤C2Cβ1(δn)βα. |
Similarly, from (3.1) and (3.3), we deduce that there exists a constant C3=C3(τ) such that
λτ,n(1κ(t(⋯))Dνxf(t(⋯),⋅,⋅))≤C3sup0≤t1<t2≤τ,s(t2)−s(t1)≤δn|t2−t1|γ≤C3Cγ1(δn)γα. |
Since 0<βα<1 and 0<γα<1, it follows that
∞∑n=1λτ,n(1κ(t(⋅))Dνxv(t(⋅),⋅))≤C2Cβ1∞∑n=1(δn)βα<∞,∞∑n=1λτ,n(1κ(t(⋅))Dνx,uf(t(⋅),⋅,⋅))≤C3Cγ1∞∑n=1(δn)γα<∞, |
which means that the functions v(t(s),x) and f(t(s),x,u) (with the function κ(t(s))) satisfy conditions (2.13) and (2.15). The lemma is proved.
With Lemma 2 proved, under the assumption of the validity of Theorem 1, we can state the following result, which immediately follows from Theorem 1 and Lemma 2.
Corollary 3. Suppose that the functions κ(t), v(t,x), f(t,x,u), and u0(x) satisfy the conditions (2.3)–(2.5), (2.12), (2.14), (2.16), and (3.1)–(3.3). Then, for any τ>0, the functions u[n](t,x) defined by (2.18)–(2.20) (with s(t) and s[n]k defined by (2.6) and (2.17)) and their first and second derivatives with respect to x∈Rd converge uniformly in [0,τ]×Rd to a function u(t,x) and its first and second derivatives with respect to x∈Rd, and the limit function u(t,x) satisfies equation (2.1) and the initial condition (2.2) in the sense of integral equality (2.21).
We will prove Theorem 1 by transforming Eq (2.1) into an equation with the diffusion coefficient κ=1. Theorem 1 will result from the theorem for the equation with the constant diffusion coefficient κ, which we will prove subsequently.
Consider the equation
∂tu(t,x)+v(t,x)⋅∇u(t,x)=Δu(t,x)+f(t,x,u(t,x)),t>0, x∈Rd, | (4.1) |
and the initial condition
u(0,x)=u0(x),x∈Rd. | (4.2) |
The first term on the right-hand side of (4.1) can be κΔu(t,x) instead of Δu(t,x), but this generalization is almost immediate. Therefore, here we consider the equation in the form of (4.1). For this problem, we have the following theorem.
Theorem 4. Suppose that
Dνxv(t,x)∈Cb([0,τ]×Rd))∀ν∈Nd, |ν|≤3, ∀τ>0, | (4.3) |
Dνxv(t,x)∈Λ∀ν∈Nd, |ν|≤2, | (4.4) |
Dνx,uf(t,x,u)1+|u|∈Cb([0,τ]×Rd)∀ν∈Nd+1, |ν|≤3, ∀τ>0, | (4.5) |
Dνx,uf(t,x,u)∈Λ∀ν∈Nd+1, |ν|≤2, | (4.6) |
Dνxu0(x)∈Cb(Rd)∀ν∈Nd, |ν|≤3, | (4.7) |
where Λ is the function class defined by (2.9). Let us also define
t[n]k=kδn, δn=2−n. | (4.8) |
Then, for any τ>0, the functions u[n](t,x) defined by
u[n](t[n]0,x)=u0(x),u[n](t[n]k,x)=∫RdΘn(y)u[n](t[n]k−1,x−δnv(t[n]k,x)−y)dy | (4.9) |
+δnf(t[n]k−1,x,u[n](t[n]k−1,x)),k=1,2,⋯, | (4.10) |
u[n](t,x)=t[n]k−tδnu[n](t[n]k−1,x)+t−t[n]k−1δnu[n](t[n]k,x)for t[n]k−1≤t≤t[n]k, | (4.11) |
and their first and second derivatives with respect to x∈Rd, converge uniformly in [0,τ]×Rd to a function u(t,x) and its first and second derivatives with respect to x∈Rd, and the limit function u(t,x) satisfies Eq (4.1) and the initial condition in the sense of the integral equality
−∫∞0u(t,x)ddtφ(t)dt−u0(x)φ(0)+∫∞0v(t,x)⋅∇u(t,x)φ(t)dt=∫∞0(Δu(t,x)+f(t,x,u))φ(t)dt | (4.12) |
for every φ(⋅)∈C1(R+) such that φ(t)=0 for t≥τ1 with a τ1>0.
The proof of Theorem 4 follows the scheme developed in [12,13]. The demonstration of the estimates of the approximate solutions and their derivatives with respect to x∈Rd does not differ from that presented in [12,13] only for simple modifications. However, to prove the convergence of the approximate solutions, we need to specify the consequence of the conditions (4.4) and (4.6), which was not explicitly exposed in [12,13]. Therefore, before we begin the essential part of the proof of Theorem 4, let us revisit the estimates of the approximate solutions.
Lemma 5. Assume that the hypotheses of Theorem 4 hold. Let u[n](t,x) be the functions defined by (2.18)–(2.20). Then, there exist functions Φ0(t), Φ1(t), Φ2(t), and Φ3(t) that are continuous on R+, increasing, independent of n, and such that we have
supx∈Rd|u[n](t,x)|≤Φ0(t), | (4.13) |
∑|ν|=1supx∈Rd|Dνxu[n](t,x)|≤Φ1(t), | (4.14) |
∑|ν|=2supx∈Rd|Dνxu[n](t,x)|≤Φ2(t), | (4.15) |
∑|ν|=3supx∈Rd|Dνxu[n](t,x)|≤Φ3(t) | (4.16) |
for all t≥0 and for all n∈N∖{0}.
Proof. The existence of Φ0(t) satisfying (4.13) can be established similarly to Lemma 5 in [5] and Lemma 1 in [2]. We note that if we fix τ>0, according to condition (4.5), the function Dνx,uf(t,x,u) with |ν|≤3 is continuous and bounded in [0,τ]×Rd. Hence, the lemma can be demonstrated in a manner entirely analogous to [12] and [13].
Having recalled the necessary estimates of the approximate solutions and their derivatives, we now proceed with the proof of Theorem 4.
Proof. We will demonstrate it in stages: Step 1 – Convergence of the approximate solutions.
Step 2 – Convergence of their first derivative. Step 3 – Convergence of their second derivatives.
Step 4 – Passage to the limit.
To simplify the presentation, we introduce the fellowing notations:
λτ,n(v)=max|ν|≤2λτ,n(1κ(t(⋅))Dνxv(t(⋅),⋅)),λτ,n(f)=max|ν|≤2λτ,n(1κ(t(⋯))Dνxf(t(⋯),⋅,⋅)),¯λτ,n=max(λτ,n(v),λτ,n(f)) |
(for the symbol λτ,n(⋅), see (2.10) and (2.11)). Moreover, in different inequalities, we simply denote C (or C′) as constants that may depend on τ but do not depend on n, when it is not necessary to specify them.
Step 1 –Convergence of the approximate solutions: We will prove that, for any τ>0, the functions u[n](t,x) converge uniformly to a function u(t,x) on [0,τ]×Rd as n→∞.
Let us first examine the difference between u[n+1](t[n+1]2k,x) and u[n](t[n]k,x) for n=1,2,⋯ and k=0,1,⋯ (recalling that t[n+1]2k=t[n]k (see (4.8))). Applying the definition (4.10) twice and using the notation
ξ[n′]k′(x,y)=x−δn′v(t[n′]k′,x)−y | (5.1) |
for k′=k+1 (or 2k+1 or 2k+2) and n′=n (or n+1), we have
u[n+1](t[n+1]2k+2,x)=I[n+1]2k+J[n+1]a,2k+J[n+1]b,2k, | (5.2) |
where
I[n+1]2k=∫Rd∫RdΘn+1(y1)Θn+1(y2)u[n+1](t[n+1]2k,ξ∗(y1,y2))dy1dy2, | (5.3) |
ξ∗(y1,y2)=ξ[n+1]2k+2(x,y1)−δn+1v(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))−y2,J[n+1]a,2k=δn+1∫RdΘn+1(y)×f(t[n+1]2k,ξ[n+1]2k+2(x,y),u[n+1](t[n+1]2k,ξ[n+1]2k+2(x,y)))dy, | (5.4) |
J[n+1]b,2k=δn+1f(t[n+1]2k+1,x,U1+U2), | (5.5) |
U1=∫RdΘn+1(y)u[n+1](t[n+1]2k,ξ[n+1]2k+1(x,y))dy, |
U2=δn+1f(t[n+1]2k,x,u[n+1](t[n+1]2k,x)). |
Therefore, recalling once again the definition (4.10) and the relation δn=2δn+1, we have
u[n+1](t[n+1]2k+2,x)−u[n](t[n]k+1,x)=D(0)+D(a)+D(b), | (5.6) |
where
D(0)=I[n+1]2k−∫RdΘn(y)u[n](t[n]k,ξ[n]k+1(x,y))dy,D(a)=J[n+1]a,2k−δn+1f(t[n]k,x,u[n](t[n]k,x)),D(b)=J[n+1]b,2k−δn+1f(t[n]k,x,u[n](t[n]k,x)). |
To estimate D(0), we note that, thanks to the well-known property of Gaussian functions, we have
∫RdΘn(z)u[n](t[n]k,ξ[n]k+1(x,z))dz=∫Rd∫RdΘn+1(y1)Θn+1(y2)u[n](t[n]k,ξ[n]k+1(x,y1+y2))dy1dy2 |
and thus
D(0)=∫Rd∫RdΘn+1(y1)Θn+1(y2)×(u[n+1](t[n+1]2k,ξ∗(y1,y2))−u[n](t[n]k,ξ[n]k+1(x,y1+y2)))dy1dy2. |
Now, based on the definition of λτ,n+1(⋅), we have
|v(t[n+1]2k+1,ξ)−v(t[n]k+1,ξ)|=|v(t[n+1]2k+1,ξ)−v(t[n+1]2k+1+δn+1,ξ)|≤λτ,n+1(v), |
and recalling the expression for ξ∗(y1,y2) and ξ[n]k+1(x,y1+y2), we obtain
|ξ∗(y1,y2)−ξ[n]k+1(x,y1+y2)|=δn+1|v(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))−v(t[n]k+1,x)|≤δn+1(λτ,n+1(v)+sup|∇v|(δn+1sup|v|+|y1|)). |
Note that we simply write sup|∇v| instead of sup(t,y)∈[0,τ]×Rd|∇v(t,x)|, wherever this abbreviated notation does not cause ambiguity. It follows that
|u[n+1](t[n+1]2k,ξ∗(y1,y2))−u[n](t[n]k,ξ[n]k+1(x,y1+y2))|≤supy∈Rd|u[n+1](t[n+1]2k,y)−u[n](t[n]k,y)|+sup|∇u[n+1]|(δn+1λτ,n+1(v)+δn+1sup|∇v|(δn+1sup|v|+|y1|)). | (5.7) |
Therefore, considering Lemma 5 and the relation ∫RdΘn+1(y1)|y1|dy1=2√δn+1√π, we deduce that there exists a constant K1 such that
|D(0)|≤supy∈Rd|u[n+1](t[n+1]2k,y)−u[n](t[n]k,y)|+K1δn+1(λτ,n+1(v)+δ1/2n+1)sup|∇u[n+1]|. | (5.8) |
Regarding D(a), using the inequality
|f(t[n+1]2k,ξ[n+1]2k+2(x,y),u[n+1](t[n+1]2k,ξ[n+1]2k+2(x,y)))−f(t[n]k,x,u[n](t[n]k,x))|≤(sup|∇xf|+sup|∂uf|sup|∇u[n+1]|)(δn+1sup|v|+|y|)+sup|∂uf||u[n+1](t[n+1]2k,x)−u[n](t[n]k,x)|, |
we easily obtain
|D(a)|≤K2(δ2n+1+δ3/2n+1)+K2δn+1|u[n+1](t[n+1]2k,x)−u[n](t[n]k,x)| | (5.9) |
with a constant K2 independent of n.
As for D(b), recalling the expression of U1 and U2 and considering the relation
|f(t[n+1]2k+1,x,u[n+1](t[n+1]2k,x))−f(t[n+1]2k,x,u[n+1](t[n+1]2k,x))|≤λτ,n+1(f)(see(2.11)), |
we have
|f(t[n+1]2k+1,x,U1+U2)−f(t[n]k,x,u[n](t[n]k,x))|≤sup|∂uf|sup|∇u[n+1]|(δn+1sup|v|+|y|)+δn+1sup|f|+λτ,n+1(f)+sup|∂uf||u[n+1](t[n+1]2k,x)−u[n](t[n]k,x)|. |
We deduce that there exists a constant K3 independent of n such that
|D(b)|≤K3(δ2n+1+δn+1(λτ,n+1(f)+δ1/2n+1))+K3δn+1|u[n+1](t[n+1]2k,x)−u[n](t[n]k,x)|. | (5.10) |
Finally, from the relations (5.2), (5.6), and (5.8)–(5.10), we deduce that there exists a constant K4 independent of n such that
|u[n+1](t[n+1]2k+2,x)−u[n](t[n]k+1,x)|≤(1+K4δn+1)supy∈Rd|u[n+1](t[n+1]2k,y)−u[n](t[n]k,y)|+K4δn+1(¯λτ,n+1+δ1/2n+1). |
Therefore, if we set
Yk=supx∈Rd|u[n+1](t[n+1]2k,x)−u[n](t[n]k,x)|, | (5.11) |
we have
Yk+1≤(1+K4δn+1)Yk+K4δn+1(¯λτ,n+1+δ1/2n+1), | (5.12) |
where, considering the relation Y0=0, we obtain
Yk≤δn+1(¯λτ,n+1+δ1/2n+1)K4k∑j=0(1+K4δn+1)k−j≤(¯λτ,n+1+δ1/2n+1)etK4, | (5.13) |
or
supx∈Rd|u[n+1](t,x)−u[n](t,x)|≤(¯λτ,n+1+δ1/2n+1)etK4for 0≤t≤τ. | (5.14) |
As under the assumptions (4.4) and (4.6) (also refer to (2.9)), we have
∞∑n=1(¯λτ,n+1+δ1/2n+1)<∞, |
and considering also the independence of K4 from n, it is evident that the inequality (5.14) and the definition (2.20) imply that the sequence u[n](t,x) converges uniformly on [0,τ]×Rd as n→∞.
Step 2 –Convergence of the first derivatives of the approximate solutions – We will demonstrate that, for any τ>0, the functions ∂xiu[n](t,x), i=1,⋯,d, converge to ∂xiu(t,x) (where u(t,x) is the limit function obtained in Step 1) uniformly on [0,τ]×Rd as n→∞.
We put
w[1,n]i,k(x)=∂∂xiu[n](t[n]k,x),i=1,⋯,d, | (5.15) |
and we will examine w[1,n+1]i,2k+2(x)−w[1,n]i,k+1(x). To simplify the notation, let us introduce abbreviated notations:
u[n]k(x)=u[n](t[n]k,x),f′i,k(x,u[n](t[n]k,x))=∂f(t[n]k,x,u)∂xi|u=u[n](t[n]k,x), |
and similarly for f′u,k(x,u[n](t[n]k,x)). Additionally, denote by ξ[n′]k′,j(x,y) the j-th component of the vector ξ[n′]k′(x,y) (see (5.1)).
By taking the derivative of both sides of equation (4.10) with respect to xi, we obtain
w[1,n]i,k(x)=∫RdΘn(y)[w[1,n]i,k−1(ξ(x,y))−δnd∑j=1∂vj(t[n]k,x)∂xiw[1,n]j,k−1(ξ(x,y))]dy+δnf′u,k−1(x,u[n](t[n]k−1,x))w[1,n]i,k−1(x)+δnf′i,k−1(x,u[n](t[n]k−1,x)). | (5.16) |
On the other hand, by deriving with respect to xi the right-hand side of (5.2), which corresponds to the terms given in (5.3)–(5.5), we get
w[1,n+1]i,2k+2(x)=∫Rd∫RdΘn+1(y1)Θn+1(y2)[w[1,n+1]i,2k(ξ∗)−δn+1d∑j=1w[1,n+1]j,2k(ξ∗)(∂xivj(t[n+1]2k+2,x)+d∑l=1∂xlvj(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))∂xiξ[n+1]2k+2,l(x,y1))]dy1dy2+δn+1∫RdΘn+1(y1)d∑j=1[f′j,2k(ξ[n+1]2k+2(x,y1),u[n+1]2k(ξ[n+1]2k+2(x,y1)))+f′u,2k(ξ[n+1]2k+2(x,y1),u[n+1]2k(ξ[n+1]2k+2(x,y1)))w[1,n+1]j,2k(ξ[n+1]2k+2(x,y1))]∂xiξ[n+1]2k+2,j(x,y1)dy1+δn+1f′i,2k+1(x,u[n+1]2k+1(x))+δn+1f′u,2k+1(x,u[n+1]2k+1(x))w[n+1]i,2k+1(x), | (5.17) |
where ξ∗=ξ∗(y1,y2) (see (5.3)).
From (5.16) and (5.17), it follows that
w[1,n+1]i,2k+2(x)−w[1,n]i,k+1(x)=7∑p=1Zp, | (5.18) |
where
Z1=∫Rd∫RdΘn+1(y1)Θn+1(y2)×(w[1,n+1]i,2k(ξ∗)−w[1,n]i,k(ξ[n]k+1(x,y1+y2)))dy1dy2, | (5.19) |
Z2=δn+1∫Rd∫RdΘn+1(y1)Θn+1(y2)d∑j=1ζ2,j(x,y1,y2)dy1dy2,ζ2,j(x,y1,y2)=−∂xivj(t[n+1]2k+2,x)w[1,n+1]j,2k(ξ∗)+∂xivj(t[n]k+1,x)w[1,n]j,k(ξ[n]k+1(x,y1+y2)), | (5.20) |
Z3=δn+1∫Rd∫RdΘn+1(y1)Θn+1(y2)d∑j=1ζ3,j(x,y1,y2)dy1dy2,ζ3,j(x,y1,y2)=−d∑l=1∂xlvj(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))∂xiξ[n+1]2k+2,l(x,y1)w[1,n+1]j,2k(ξ∗)+∂xivj(t[n]k+1,x)w[1,n]j,k(ξ[n]k+1(x,y1+y2)), | (5.21) |
Z4=δn+1∫RdΘn+1(y1)d∑j=1ζ4,j(x,y1)dy1−δn+1f′i,k(x,u[n]k(x)),ζ4,j(x,y1)=f′j,2k(ξ[n+1]2k+2(x,y1),u[n+1]2k(ξ[n+1]2k+2(x,y1)))∂xiξ[n+1]2k+2,j(x,y1), | (5.22) |
Z5=δn+1∫RdΘn+1(y1)d∑j=1ζ5,j(x,y1)dy1−δn+1f′u,k(x,u[n]k(x))w[1,n]i,k(x),ζ5,j(x,y1)=f′u,2k(ξ[n+1]2k+2(x,y1),u[n+1]2k(ξ[n+1]2k+2(x,y1)))×w[1,n+1]j,2k(ξ[n+1]2k+2(x,y1))∂xiξ[n+1]2k+2,j(x,y1), | (5.23) |
Z6=δn+1(f′i,2k+1(x,u[n+1]2k+1(x))−f′i,k(x,u[n]k(x))), | (5.24) |
Z7=δn+1(f′u,2k+1(x,u[n+1]2k+1(x))w[1,n+1]i,2k+1(x)−f′u,k(x,u[n]k(x))w[1,n]i,k(x)). | (5.25) |
As
ξ∗−ξ[n]k+1(x,y1+y2)=δn+1(v(t[n]k+1,x)−v(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))), |
similarly to (5.7), we have
|w[1,n+1]i,2k(ξ∗)−w[1,n]i,k(ξ[n]k+1(x,y1+y2))|≤supy∈Rd|w[1,n+1]i,2k(y)−w[1,n]i,k(y)|+sup|∇w[1,n+1]i,2k|×(Cδn+1λτ,n+1(v)+δn+1sup|∇v|(δn+1sup|v|+|y1|)). | (5.26) |
Recall that according to Lemma 5, |∇w[1,n+1]i,2k| is bounded by a constant independent of n. Thus, similarly to obtaining (5.8), we get
|Z1|≤supy∈Rd|w[1,n+1]i,2k(y)−w[1,n]i,k(y)|+C′δn+1(λτ,n+1(v)+δ1/2n+1). | (5.27) |
Let us introduce the notation
Y[1]k=d∑i=1supx∈Rd|w[1,n+1]i,2k(x)−w[1,n]i,k(x)|. | (5.28) |
Recalling that t[n+1]2k+2=t[n]k+1 and using the inequality (5.26), which, written with j instead of i, is valid for all j∈{1,⋯,d}, we obtain
|Z2|≤δn+1C(Y[1]k+Cδn+1(λτ,n+1(v)+δ1/2n+1)). | (5.29) |
Regarding ζ3,j(x,y1,y2) that appears under the integration sign in (5.21), considering the relation
∂xiξ[n+1]2k+2,l(x,y1)=δil−δn+1∂xivl(t[n+1]2k+2,x), | (5.30) |
where δil is the Kronecker symbol, we have
ζ3,j(x,y1,y2)=δn+1d∑l=1∂xlvj(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))∂xivl(t[n+1]2k+2,x)w[1,n+1]j,2k(ξ∗)+(∂xivj(t[n]k+1,ξ[n+1]2k+2(x,y1))−∂xivj(t[n+1]2k+1,ξ[n+1]2k+2(x,y1)))w[1,n+1]j,2k(ξ∗)+(∂xivj(t[n]k+1,x)−∂xivj(t[n]k+1,ξ[n+1]2k+2(x,y1)))w[1,n+1]j,2k(ξ∗)+∂xivj(t[n]k+1,x)(w[1,n+1]j,2k(ξ[n]k+1(x,y1+y2))−w[1,n+1]j,2k(ξ∗))+∂xivj(t[n]k+1,x)(w[1,n]j,k(ξ[n]k+1(x,y1+y2))−w[1,n+1]j,2k(ξ[n]k+1(x,y1+y2))). | (5.31) |
As
|∂xivj(t[n]k+1,ξ)−∂xivj(t[n+1]2k+1,ξ)|≤λτ,n+1(v) (ξ∈Rd),|∂xivj(t[n]k+1,x)−∂xivj(t[n]k+1,ξ[n+1]2k+2(x,y1))|≤sup|∇∂xivj|(δn+1sup|v|+|y1|), |
given (5.26) and the values uniformly bounded by the hypotheses and Lemma 5, we have
|Z3|≤Cδn+1(λτ,n+1(v)+δ1/2n+1)+δn+1CY[1]k. | (5.32) |
Regarding Z4, Z5, Z6, and Z7, recalling the notation convention f′⋅,2k(x,u)=f′⋅,k(x,u) and taking into account the relation
|f′⋅,k(x(1),u(1))−f′⋅,k(x(2),u(2))|≤C(|x(1)−x(2)|+|u(1)−u(2)|) |
and relations (4.6) and (5.30), we have
|Z4+Z5+Z6+Z7|≤Cδn+1(λτ,n+1(f)+δ1/2n+1)+δn+1C(Yk+Y[1]k), | (5.33) |
where Yk is defined in (5.11).
By summing up the inequalities (5.27), (5.29), (5.32), and (5.33), we have
|w[1,n+1]i,2k+2(x)−w[1,n]i,k+1(x)|≤7∑p=1|Zp|≤supy∈Rd|w[1,n+1]i,2k(y)−w[1,n]i,k(y)|+Cδn+1(¯λτ,n+1+δ1/2n+1)+δn+1C(Yk+Y[1]k). |
As this inequality holds for any x∈Rd, summing up this inequality for i=1,⋯,d, we get
Y[1]k+1≤Y[1]k+Cδn+1(¯λτ,n+1+δ1/2n+1)+δn+1C(Yk+Y[1]k). | (5.34) |
If we set
˜Y[1]k=Yk+Y[1]k, | (5.35) |
then, by adding (5.34) and (5.12), we obtain
˜Y[1]k+1≤˜Y[1]k+C′δn+1(¯λτ,n+1+δ1/2n+1)+δn+1C′˜Y[1]k. | (5.36) |
Therefore, similarly to the derivation of (5.13) (and thus (5.14)), we obtain
˜Y[1]k≤(¯λτ,n+1+δ1/2n+1)etC′. | (5.37) |
As, due to (2.9), (4.4), and (4.6), we have ∑∞n=1(¯λτ,n+1+δ1/2n+1)<∞, from inequality (5.37) and definition (2.20), we conclude that the sequence w[1,n]i(t,x)=∂∂xiu[n](t,x) converges uniformly on [0,τ]×Rd.
Step 3 –Convergence of second derivatives of approximate solutions –We will demonstrate that, for any τ>0, the functions ∂2∂xi∂xju[n](t,x), i,j=1,⋯,d, converge to ∂2∂xi∂xju(t,x) (u(t,x) being the limiting function obtained in Step 1) uniformly on [0,τ]×Rd as n→∞.
Let us define
w[2,n]ij,k(x)=∂∂xjw[1,n]i,k(x)=∂2∂xj∂xiu[n](t[n]k,x) | (5.38) |
and initially estimate
w[2,n+1]ij,2k+2(x)−w[2,n]ij,k+1(x)=7∑p=1∂xjZp, | (5.39) |
where Zp, p=1,⋯,7, are the terms defined in (5.19)–(5.25).
Recalling that for the l-th component ξ∗l of ξ∗=ξ∗(y1,y2) (see (5.3)), we have
∂xjξ∗l=δjl−δn+1∂xjvl(t[n+1]2k+2,x)−δn+1∂xjvl(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))+δ2n+1d∑q=1∂xqvl(t[n+1]2k+1,ξ[n+1]2k+2(x,y1))∂xjvq(t[n+1]2k+2,x) | (5.40) |
(see (5.30)), whereas for the l-th component ξ[n]k+1,l(x,y1+y2) of ξ[n]k+1(x,y1+y2), we have
∂xjξ[n]k+1,l(x,y1+y2)=δjl−2δn+1∂xjvl(t[n]k+1,x). | (5.41) |
Using relations (5.40) and (5.41) and the hypotheses on the regularity of v(t,x), reasoning similarly to obtain (5.27), we get
|∂xjZ1|≤(1+Cδn+1)supy∈Rd|w[2,n+1]ij,2k(y)−w[2,n]ij,k(y)|+Cδn+1(λτ,n+1(v)+δ1/2n+1). | (5.42) |
Notice that for p=2 and p=3, we have
∂xjZp=δn+1∫Rd∫RdΘn+1(y1)Θn+1(y2)d∑l=1∂xjζp,l(x,y1,y2)dy1dy2 |
and for p=2, we decompose ζ2,l as
ζ2,l=∂xivl(t[n]k+1,x)(w[1,n]l,k(ξ[n]k+1(x,y1+y2))−w[1,n+1]l,2k(ξ[n]k+1(x,y1+y2)))+∂xivl(t[n]k+1,x)(w[1,n+1]l,2k(ξ[n]k+1(x,y1+y2))−w[1,n+1]l,2k(ξ∗)). |
Recall that a similar decomposition of ζ3,l was made in (5.31). By examining the terms in which ζ2,l and ζ3,l decompose and deriving each term with respect to xj, we can notice, particularly with the help of (5.30) and (5.40), that the absolute value of the derivative of each term is bounded by
δn+1C |
or
λτ,n+1(v)C |
or
C(δn+1+|y1|) |
or
CY[1]k |
or
CY[2]k, |
where Y[1]k is the function defined by (5.28), while Y[2]k is defined by
Y[2]k=d∑l,j=1supx∈Rd|w[2,n+1]lj,2k(x)−w[2,n]lj,k(x)|. | (5.43) |
Using the properties of the integral of Θn+1 over Rd, we deduce that
|∂xjZ2+∂xjZ3|≤Cδn+1(λτ,n+1(v)+δ1/2n+1)+δn+1C(Y[1]k+Y[2]k). | (5.44) |
Regarding ∂xjZp, p=4,⋯,7, applying the considerations used earlier to each term found in the expression of ∂xjZp, we readily obtain the inequality
|∂xj(Z4+Z5+Z6+Z7)|≤Cδn+1(λτ,n+1(f)+δ1/2n+1)+δn+1C(Yk+Y[1]k+Y[2]k), | (5.45) |
where Yk is the function defined by (5.11).
However, we can proceed similarly to the final part of the demonstration in Step 2. That is, by combining the inequalities (5.42), (5.44), and (5.45), considering that the inequality obtained is valid for |w[2,n+1]ij,2k+2(x)−w[2,n]ij,k+1(x)| for all x∈Rd, and summing up the inequality obtained for i=1,⋯,d, we obtain
Y[2]k+1≤Y[2]k+C(δ2n+1+δn+1(¯λτ,n+1+δ1/2n+1))+δn+1C(Yk+Y[1]k+Y[2]k). | (5.46) |
If we set
˜Y[2]k=Yk+Y[1]k+Y[2]k, | (5.47) |
then, by combining (5.12), (5.34), and (5.46), we obtain
˜Y[2]k+1≤˜Y[2]k+C′(δ2n+1+δn+1(¯λτ,n+1+δ1/2n+1))+δn+1C′˜Y[2]k. | (5.48) |
Therefore, we obtain
˜Y[2]k≤(¯λτ,n+1+δ1/2n+1))etC′, | (5.49) |
from which we deduce that the sequence w[1,n]ij(t,x)=∂2∂xi∂xju[n](t,x) converges uniformly on [0,τ]×Rd.
Step 4 —Convergence to the limit —
First, let us demonstrate that, given τ>0, for t[n]1≤t[n]k≤τ, we have
u[n](t[n]k,x)−u[n](t[n]k−1,x)δn=−v(t,x)⋅∇u[n](t[n]k−1,x)+Δu[n](t[n]k−1,x)+f(t[n]k−1,x,u[n](t[n]k−1,x))+R | (5.50) |
with
|R|≤(δ2n+δ1/2n)C. | (5.51) |
Indeed, according to Taylor's formula, we have
u[n](t[n]k−1,x−δnv(t,x)+y)=u[n](t[n]k−1,x)−δnv(t,x)⋅∇u[n](t[n]k−1,x)+y⋅∇u[n](t[n]k−1,x)+12d∑i,j=1[δ2nvi(x)vj(t,x)−2δnvi(t,x)yj+yiyj]∂2u[n](t[n]k−1,x)∂xi∂xj+16d∑i,j,h=1μiμjμh∂3u[n](t[n]k−1,˜x)∂xi∂xj∂xh, | (5.52) |
where μi=−δnvi−yi (and similarly for μj and μh), while ˜x is a point between x and x−δnv(t[n]k,x)−y.
However, since
∫RdΘn(y)yjdy=0, ∫RdΘn(y)yiyjdy=0 if i≠j, ∫RdΘn(y)y2idy=2δn, |
we have
∫RdΘn(y)y⋅∇u[n](t[n]k−1,x)dy=0,∫RdΘn(y)[12d∑i,j=1(δ2nvi(t,x)vj(t,x)−2δnvi(t,x)yj+yiyj)∂2u[n](t[n]k−1,x)∂xi∂xj]dy=δnΔu[n](t[n]k−1,x)+δ2n12d∑i,j=1vi(t,x)vj(t,x)∂2u[n](t[n]k−1,x)∂xi∂xj. |
On the other hand, since we have |μi|≤δn|v|+|y| and similarly for μj and μh, there exists a constant C such that
|163∑i,j,h=1μiμjμh∂3u[n](t[n]k−1,x)∂xi∂xj∂xh|≤C(δn|v|+|y|)3|∂3u[n](t[n]k−1,x)∂xi∂xj∂xh|. |
Since, according to Lemma 5, the third derivatives of u[n] are uniformly bounded, we have
|∫RdΘn(y)163∑i,j,h=1μiμjμh∂3u[n](t[n]k−1,x)∂xi∂xj∂xhdy|≤(δ3n+δ3/2n)C′. |
We deduce that
u[n](t[n]k,x)−u[n](t[n]k−1,x)=−δnv(t,x)⋅∇u[n](t[n]k−1,x)+δnκΔu[n](t[n]k−1,x)+δnF(t[n]k−1,x,u[n](t[n]k−1,x))+R′ |
where
|R′|≤(δ3n+δ3/2n)C; |
therefore, dividing both sides of this equality by δn, we obtain (5.50) with (5.51).
From (5.50), it follows that there exists a constant L independent of n such that
|u[n](t1,x)−u[n](t2,x)|≤L|t1−t2|∀t1,t2∈[0,τ−1], ∀x∈Rd, | (5.53) |
|u(t1,x)−u(t2,x)|≤L|t1−t2|∀t1,t2∈[0,τ−1], ∀x∈Rd, | (5.54) |
where u(t,x) is the limit function of the sequence u[n](t,x).
Indeed, according to Lemma 5 and the obvious relation δn≤δ1, the absolute value of the right-hand side of (5.50) is bounded by a constant L that does not depend on n. Therefore, the inequality (5.53) follows from the definition (2.20). Furthermore, inequality (5.54) results from (5.53) and the uniform convergence of u[n](t,x) to u(t,x).
Now we are ready to conclude the proof of Theorem 4. Consider the function
ψ[n](t,x)=t[n]k+1−tδn(u[n](t[n]k+1,x)−u[n](t[n]k,x)δn)+t−t[n]kδn(u[n](t[n]k+2,x)−u[n](t[n]k+1,x)δn),for t[n]k<t<t[n]k+1, k=0,1,⋯. | (5.55) |
It is immediately evident that ψ[n](t,x) is continuous concerning t, and according to the definition (2.20), we have
ψ[n](t,x)=(t[n]k+1−t)δn∂u[n](t,x)∂t+(t−t[n]k)δn∂u[n](t+δn,x)∂t,for t[n]k<t<t[n]k+1. |
Furthermore, according to relations (5.50) along with (5.51) and definition in (2.20), we have
ψ[n](t,x)=−v(t,x)⋅∇u[n](t,x)+Δu[n](t,x)+f(t,x,u[n](t,x))+R | (5.56) |
where
|R|≤(δ2n+δ1/2n)C. |
Consider now a function φ(⋅)∈C∞([0,∞[) such that φ(t)=0 for t≥τ1 with τ1>0. By multiplying both sides of (5.56) by the function φ(t) and integrating with respect to t, we obtain
∫∞0ψ[n](t,x)φ(t)dt=∫∞0(−v(t,x)⋅∇u[n](t,x)+Δu[n](t,x)+f(t,x,u[n](t,x))+R)φ(t)dt. | (5.57) |
By virtue of (5.51) (also refer to (5.56)) and what we have proved in Steps 1, 2, and 3, the right-hand side of equality (5.57) tends to
∫∞0(−v(t,x)⋅∇u(t,x)+Δu(t,x)+f(t,x,u(t,x)))φ(t)dt. |
On the other hand, if we set
Ψ[n](t,x)=12(u[n](t[n]0,x)+u[n](t[n]1,x))+∫t0ψ[n](t′,x)dt′, |
by performing integration by parts, the first term of (5.57) transforms into
−∫∞0Ψ[n](t,x)φ′(t)dt−12(u[n](t[n]0,x)+u[n](t[n]1,x))φ(0)≡I. |
Now, by explicit calculation, we observe that
Ψ[n](t,x)=12(u[n](t[n]k,x)+u[n](t[n]k+1,x))−(12−(t[n]k+1−t)22δ2n)u[n](t[n]k,x)+(12−(t[n]k+1−t)22δ2n−(t−t[n]k)22δ2n)u[n](t[n]k+1,x)+(t−t[n]k)22δ2nu[n](t[n]k+2,x) |
for t[n]k<t≤t[n]k+1 and k=0,1,2,⋯. From this expression of Ψ[n](t,x), the uniform convergence of u[n](t,x) to u(t,x) (Step 1), and the relation (5.53), we can conclude that
Ψ[n](t,x)→u(t,x)uniformly on [0,τ]×Rdas n→∞ |
for any τ>0. Furthermore, as u[n](t[n]0,x)=u0(x) for all n and for all x∈Rd, considering (5.53), we have
12(u[n](t[n]0,x)+u[n](t[n]1,x))→u0(x)uniformly on Rdas n→∞. |
Therefore, I tends toward
−∫∞0u(t,x)φ′(t)dt−u0(x)φ(0), |
which gives us (4.12). The proof of Theorem 4 is complete.
Proof. As t(s) is the inverse function of s(t), by virtue of (2.6), we have
dt(s)ds=1ds(t)dt=1κ(t). |
We thus obtain
∂su(t(s),x)=1κ(t)∂tu(t,x), |
allowing us to transform Eq (2.1) into
∂su(t(s),x)+1κ(t(s))v(t(s),x)⋅∇u(t(s),x)=Δu(t(s),x)+1κ(t(s))f(t(s),x,u(t(s),x)). | (6.1) |
If we set
˜v(s,x)=1κ(t(s))v(t(s),x),˜f(s,x,u)=1κ(t(s))f(t(s),x,u), |
according to hypotheses (2.12)–(2.15), the functions ˜v(s,x) and ˜f(s,x,u) satisfy the conditions (4.3)–(4.6) by replacing t with s. Therefore, following Theorem 4, the functions u[n](s,x) defined by (4.9)–(4.11), where s substitutes t, converge uniformly on [0,τ]×Rd for any τ>0, with their first and second derivatives, to a function u(s,x). The limit function u(s,x) satisfies equation (4.1) (with s replacing t) and the initial condition (2.2) in terms of integral equality
−∫∞0u(s,x)dds˜φ(s)ds−u0(x)˜φ(0)+∫∞0˜v(s,x)⋅∇u(s,x)˜φ(s)ds=∫∞0(Δu(s,x)+˜f(s,x,u))˜φ(s)ds | (6.2) |
for any ˜φ(⋅)∈C1([0,∞[) such that ˜φ(s)=0 for s≥τ1,s with τ1,s>0.
Now, let us consider a function φ(⋅)∈C1([0,∞[) with supp(φ(⋅))⊂[0,τ1,t] and its composition φ∘t=φ(t(⋅)) with the function t(s). Then we have
ddsφ(t(s))=ddtφ(t)|t=t(s)⋅dt(s)ds=ddtφ(t)|t=t(s)⋅1κ(t(s)). |
Since t(s) is continuous and ddtφ(t) is also continuous by assumption, the function ddtφ(t)|t=t(s) is continuous in s. On the other hand, according to condition (2.4), the function t(s) is absolutely continuous, so that dt(s)ds belongs to L1loc([0,∞[). Consequently,
ddsφ(t(s))∈L1(]0,s(τ1,t)+1[). |
Hence, there exists a sequence of functions {˜φm}∞m=1 such that
˜φm∈C1(R+),supp(˜φm)⊂[0,s(τ1,t)+1]∀m∈N∖{0}, | (6.3) |
and that
‖˜φm(⋅)−φ(t(⋅))‖W11(]0,s(τ1,t)+1[)→0,|˜φm(s)−φ(0)|→0,for m→∞. | (6.4) |
According to (6.3), we can substitute ˜φ(s)=˜φm(s) into (6.2). Since u∈L∞([0,τ]×Rd) for any τ>0, considering (6.4), we have
∫∞0u(s,x)dds˜φm(s)ds→∫∞0u(s,x)ddsφ(t(s))ds=∫∞0u(t,x)ddtφ(t)dt |
as m→∞. Moreover, it can be easily seen that
∫∞0˜v(s,x)⋅∇u(s,x)˜φm(s)ds→∫∞01κ(t(s))v(t(s),x)⋅∇u(t(s),x)φ(s)ds=∫∞0v(t,x)⋅∇u(t,x)φ(t)dt,∫∞0Δu(s,x)˜φm(s)ds→∫∞0Δu(t(s),x)φ(s)ds=∫∞0κ(t)Δu(t,x)φ(t)dt,∫∞0˜f(s,x,u)˜φm(s)ds→∫∞01κ(t(s))f(t(s),x,u)φ(s)ds=∫∞0f(t,x,u)φ(t)dt |
as m→∞. Hence, it follows that for any function φ(t)∈C1(R+) such that there exists a positive number τ1,t satisfying φ(t)=0 for all t≥τ1,t, it satisfies the relation (2.21). The theorem is proved.
Lynda Taleb, Rabah Gherdaoui: Writing-original draft, Writing-review & editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this paper.
The problem treated in this paper was proposed by Prof. H. Fujita Yashima (ENS Constantine, Algeria; NHSM, Algiers, Algeria). He continuously encouraged the authors with useful suggestions to accomplish this study. They express their gratitude to him.
The authors would like to express their sincere gratitude to the reviewers for their valuable comments and constructive feedback, which have greatly contributed to improving the quality of this paper.
All authors declare no conflicts of interest in this paper.
[1] | L. Ait Mahiout, H. Fujita Yashima, Convergence de la solution d'une équation de transport-diffusion vers la solution d'une équation de transport, Ann. Math. Afr., 10 (2023), 105–124. |
[2] |
M. Aouaouda, A. Ayadi, H. Fujita Yashima, Convergence of approximate solutions by heat kernel for transport-diffusion equation in a half-plane, J. Samara State Tech. Univ., Ser. Phys. Math. Sci, 26 (2022), 222–258. https://doi.org/10.14498/vsgtu1881 doi: 10.14498/vsgtu1881
![]() |
[3] | H. Fujita Yashima, L. Ait-Mahiout, Convergence of solution of transport-diffusion system to that of transport system, Bull. Buryat St. Univ. Math. Inf., 2023 (2023), 22–36. |
[4] | M. I. Freidlin, A. D. Wentzell, Random perturbations of dynamical systems, 3 Eds., Berlin: Springer, 2012. |
[5] |
R. Gherdaoui, L. Taleb, S. Selvaduray, Convergence of the heat kernel approximated solutions of the transport-diffusion equation in the half-space, J. Math. Anal. Appl., 527 (2023), 127507. https://doi.org/10.1016/j.jmaa.2023.127507 doi: 10.1016/j.jmaa.2023.127507
![]() |
[6] |
R. Gherdaoui, S. Selvaduray, H. Fujita Yashima, Convergence of approximate solutions for transport-diffusion equation in the half-space with Neumann condition, Bull. Irkutsk State Uni. Ser. Math., 48 (2024), 64–79. https://doi.org/10.26516/1997-7670.2024.48.64 doi: 10.26516/1997-7670.2024.48.64
![]() |
[7] | O. A. Ladyzhenskaja, V. A. Solonnikov, N. N. Ural'tseva, Linear and quasi-linear equations of parabolic type, Providence: American Mathematical Society, 1968. |
[8] |
A. Nemdili, F. Korichi, H. Fujita-Yashima, Approximation of the solution of transport-diffusion equation in Hölder space, J. Samara State Tech. Uni., Ser. Phys. Math. Sci., 28 (2024), 426–444. https://doi.org/10.14498/vsgtu2097 doi: 10.14498/vsgtu2097
![]() |
[9] |
É. Pardoux, S. Peng, Backward doubly stochastic differential equations and systems of quasi-linear SPDEs, Prob. Th. Rel. Fields, 98 (1994), 209–227. https://doi.org/10.1007/BF01192514 doi: 10.1007/BF01192514
![]() |
[10] | É. Pardoux, A. Răşcanu, Stochastic differential equations, backward SDEs, partial differential equations, Berlin: Springer 2014. |
[11] |
É. Pardoux, A. Y. Veretennikov, Averaging of backward stochastic differential equations, with applications to semi-linear PDE's, Stochastics Stochastics Rep., 60 (1997), 255–270. https://doi.org/10.1080/17442509708834109 doi: 10.1080/17442509708834109
![]() |
[12] | H. Smaali, H. Fujita Yashima, Une généralisation de l'approximation par une moyenne locale de la solution de l'équation de transport-diffusion, Ann. Math. Afr., 9 (2021), 89–108. |
[13] | L. Taleb, S. Selvaduray, H. Fujita Yashima, Approximation par une moyenne locale de la solution de l'équation de transport-diffusion, Ann. Math. Afr., 8 (2020), 53–73. |