In the present paper, reaction–diffusion systems (RD-systems) with rapidly oscillating coefficients and righthand sides in equations and in boundary conditions were considered in domains with locally periodic oscillating (wavering) boundary. We proved a weak convergence of the trajectory attractors of the given systems to the trajectory attractors of the limit (homogenized) RD-systems in domain independent of the small parameter, characterizing the oscillation rate. We consider the critical case in which the type of boundary condition was preserved. For this aim, we used the approach of Chepyzhov and Vishik concerning trajectory attractors of evolutionary equations. Also, we applied the homogenization (averaging) method and asymptotic analysis to derive the limit (averaged) system and to prove the convergence. Defining the appropriate axillary functional spaces with weak topology, we proved the existence of trajectory attractors for these systems. Then, we formulated the main theorem and proved it with the help of auxiliary lemmata.
Citation: Gaziz F. Azhmoldaev, Kuanysh A. Bekmaganbetov, Gregory A. Chechkin, Vladimir V. Chepyzhov. Homogenization of attractors to reaction–diffusion equations in domains with rapidly oscillating boundary: Critical case[J]. Networks and Heterogeneous Media, 2024, 19(3): 1381-1401. doi: 10.3934/nhm.2024059
[1] | T. A. Shaposhnikova, M. N. Zubova . Homogenization problem for a parabolic variational inequality with constraints on subsets situated on the boundary of the domain. Networks and Heterogeneous Media, 2008, 3(3): 675-689. doi: 10.3934/nhm.2008.3.675 |
[2] | Patrick Henning, Mario Ohlberger . The heterogeneous multiscale finite element method for advection-diffusion problems with rapidly oscillating coefficients and large expected drift. Networks and Heterogeneous Media, 2010, 5(4): 711-744. doi: 10.3934/nhm.2010.5.711 |
[3] | Hakima Bessaih, Yalchin Efendiev, Florin Maris . Homogenization of the evolution Stokes equation in a perforated domain with a stochastic Fourier boundary condition. Networks and Heterogeneous Media, 2015, 10(2): 343-367. doi: 10.3934/nhm.2015.10.343 |
[4] | Erik Grandelius, Kenneth H. Karlsen . The cardiac bidomain model and homogenization. Networks and Heterogeneous Media, 2019, 14(1): 173-204. doi: 10.3934/nhm.2019009 |
[5] | Alexander Mielke, Sina Reichelt, Marita Thomas . Two-scale homogenization of nonlinear reaction-diffusion systems with slow diffusion. Networks and Heterogeneous Media, 2014, 9(2): 353-382. doi: 10.3934/nhm.2014.9.353 |
[6] | Iryna Pankratova, Andrey Piatnitski . Homogenization of convection-diffusion equation in infinite cylinder. Networks and Heterogeneous Media, 2011, 6(1): 111-126. doi: 10.3934/nhm.2011.6.111 |
[7] | Narcisa Apreutesei, Vitaly Volpert . Reaction-diffusion waves with nonlinear boundary conditions. Networks and Heterogeneous Media, 2013, 8(1): 23-35. doi: 10.3934/nhm.2013.8.23 |
[8] | Avner Friedman . PDE problems arising in mathematical biology. Networks and Heterogeneous Media, 2012, 7(4): 691-703. doi: 10.3934/nhm.2012.7.691 |
[9] | Shin-Ichiro Ei, Toshio Ishimoto . Effect of boundary conditions on the dynamics of a pulse solution for reaction-diffusion systems. Networks and Heterogeneous Media, 2013, 8(1): 191-209. doi: 10.3934/nhm.2013.8.191 |
[10] | François Hamel, James Nolen, Jean-Michel Roquejoffre, Lenya Ryzhik . A short proof of the logarithmic Bramson correction in Fisher-KPP equations. Networks and Heterogeneous Media, 2013, 8(1): 275-289. doi: 10.3934/nhm.2013.8.275 |
In the present paper, reaction–diffusion systems (RD-systems) with rapidly oscillating coefficients and righthand sides in equations and in boundary conditions were considered in domains with locally periodic oscillating (wavering) boundary. We proved a weak convergence of the trajectory attractors of the given systems to the trajectory attractors of the limit (homogenized) RD-systems in domain independent of the small parameter, characterizing the oscillation rate. We consider the critical case in which the type of boundary condition was preserved. For this aim, we used the approach of Chepyzhov and Vishik concerning trajectory attractors of evolutionary equations. Also, we applied the homogenization (averaging) method and asymptotic analysis to derive the limit (averaged) system and to prove the convergence. Defining the appropriate axillary functional spaces with weak topology, we proved the existence of trajectory attractors for these systems. Then, we formulated the main theorem and proved it with the help of auxiliary lemmata.
In the paper, one can find the homogenization problem for reaction–diffusion (RD) equations in domains with very rapidly wavering boundary (for detailed geometric settings [1]). We prove the existence of trajectory attractors and also obtain the convergence of the attractors as the small parameter, characterizing the oscillations, goes to zero, i.e., we prove the Hausdorff convergence of attractors as the small parameter goes to zero. Thus, we construct the limit attractor and prove the convergence of the attractors of the given problem to the attractor of the limit problem. In many pure mathematical papers, one can find the asymptotic methods applying to problems in domains with wavering (rough) boundaries (see, for example, rapidly oscillating boundaries in [1,2,3,4,5], fractal boundaries in [6], diffusivity through rough boundaries in [7], rapidly oscillating type of boundary conditions on oscillating (wavering) boundaries in [8,9], boundaries with many thin rods in [10,11,12,13]). We want to mention here the basic frameworks [14,15,16,17,18] where one can find the detailed bibliography.
Concerning attractors, see, for instance, [19,20,21] and the references in these monographs. Homogenization of attractors were studied in [21,22,23,24] and applications of this theory were investigated in [25,26,27,28].
In this paper, we proved the weak convergence of the trajectory attractor Aϵ to the RD-systems in domains with wavering boundary, as ϵ→0, to the trajectory attractors ¯A of homogenized systems in some natural functional space. Here, the small parameter ϵ characterizes the period and the amplitude of the oscillations. The parameter ϵ is included also in a Fourier condition on a part of the boundary, and we consider the case when the type of this condition is preserved (critical case).
Note that the subcritical case (the case of the Neumann homogenized condition) and supercritical (the case of the Dirichlet homogenized condition) are also interesting, but we suppose to study them in independent papers.
Section 2 is devoted to basic settings. In Section 3, one can find the framework of the theory of attractors. In Section 4, we describe the limiting (homogenized) RD-system and its trajectory attractor. Section 5 contains auxiliary results, and Section 6 is connected with the proof of the main result.
Suppose that D is a bounded domain in Rd, d≥2, with smooth boundary ∂D=Γ1∪Γ2, where D lies in a semi-space {xd>0} and Γ1⊂{x:xd=0}. Given a smooth nonpositive 1-periodic in the ˜y function F(˜x,˜y), ˜x=(x1,...,xd−1),˜y=(y1,...,yd−1), we define the domain Dϵ as follows: ∂Dϵ=Γϵ1∪Γ2, where we set Γϵ1={x=(˜x,xd):(˜x,0)∈Γ1,xd=ϵαF(˜x,˜x/ϵ)}, 0<α<1, i.e., we add the thin oscillating layer Πϵ={x=(˜x,xd):(˜x,0)∈Γ1,xd∈[0,ϵαF(˜x,˜x/ϵ))} to the domain D. Usually, we assume that F(˜x,˜y) is compactly supported on Γ1 uniformly in ˜y. Consider the problem
{∂uϵ∂t=AΔuϵ−a(x,xϵ)f(uϵ)+h(x,xϵ),x∈Dϵ,t>0,∂uϵ∂ν+ϵβp(˜x,˜xϵ)uϵ=ϵ1−αg(˜x,˜xϵ),x=(˜x,xd)∈Γϵ1,t>0,uϵ=0,x∈Γ2,t>0,uϵ=U(x),x∈Dϵ,t=0, | (2.1) |
where uϵ=uϵ(x,t)=(u1,…,un)⊤ is an unknown vector function, the nonlinear function f=(f1,…,fn)⊤ is given, h=(h1,…,hn)⊤ is the known righthand side function, and A is an n×d-matrix with constant coefficients having a positive symmetrical part: 12(A+A⊤)≥ϖI,ϖ>0 (where I is the unit matrix with dimension n). We assume that p(˜x,˜y)=diag{p1,…,pn}, g(˜x,˜y)=(g1,…,gn)⊤ are continuous, 1-periodic in ˜y, and pi(˜x,˜y), i=1,…n, are positive. Here, ∂∂ν is the co-normal derivative of the function, i.e., ∂∂ν:=d∑k,j=1Akj∂∂xkNj and N=(N1,…,Nd) is the outward normal vector to the boundary of the domain with unit length. We denote the maximum of p on Γ1 by pmax.
The function a(x,y)∈C(¯Dϵ×Rd) is such that 0<a1≤a(x,y)≤a2 with some coefficient a1, a2. We assume that function aϵ(x)=a(x,xϵ) has an average ¯a(x) when ϵ→0+ in space L∞,∗w(D), that is,
∫Da(x,xϵ)φ(x)dx→∫D¯a(x)φ(x)dx(ϵ→0+), | (2.2) |
for each φ∈L1(D).
Denote by V (respectively, Vϵ) the Sobolev space H1(D,Γ2) (respectively, H1(Dϵ,Γ2)), i.e., the space of functions from the Sobolev space H1(D) (respectively, H1(Dϵ)) with zero trace on Γ2. We also denote by V′ (respectively, V′ϵ) the dual space for V (respectively, Vϵ), i.e., the space of linear bounded functionals on V (respectively Vϵ). For vector function h(x,y), assume that for any ϵ>0, function hiϵ(x)=hi(x,xϵ)∈L2(Dϵ) and it has an average ¯hi(x) in space L2(Dϵ) for ϵ→0+, that is,
hi(x,xϵ)⇀¯hi(x)(ϵ→0+)weakly inL2(Dϵ), |
or
∫Dhi(x,xϵ)φ(x)dx→∫D¯hi(x)φ(x)dx(ϵ→0+), | (2.3) |
for each function φ∈L2(D) and i=1,…,n.
From the condition (2.3), it follows that the norms of functions hiϵ(x) are bounded uniformly in ϵ, in the space L2(Dϵ), i.e.,
‖hiϵ(x)‖L2(Dϵ)≤M0,∀ϵ∈(0,1]. | (2.4) |
We suppose that the nonlinearity f(w) is continuous, i.e., f(w)∈C(Rn;Rn), and this function satisfies
n∑k=1|fk(w)|pk(pk−1)≤M0(n∑k=1|wk|pk+1),2≤p1≤…≤pn−1≤pn, | (2.5) |
n∑k=1γk|wk|pk−M1≤n∑k=1fk(w)wk,∀w∈Rn, | (2.6) |
for γk>0 for any k=1,…,n. The inequality (2.5) is due to the fact that in real RD-systems, the functions fk(w) are polynomials with possibly different degrees. The inequality (2.6) is called the dissipativity condition for the RD-system (2.1). In a simple model case pk≡p for each k=1,…,n, bounds (2.5) and (2.6) are reduced to the following:
|f(w)|≤M0(|w|p−1+1),γ|w|p−M1≤f(w)w,∀w∈Rn. | (2.7) |
Note that the fulfillment of the Lipschitz condition for the function f(w) in the variable w is not supposed.
Remark 2.1. Using the presented methods, it is also possible to study systems in which nonlinear terms look as follows: m∑k=1ak(x,xϵ)fk(w), where ak are matrices of the elements of which admit averaging and fk(w) are polynomial vectors of w, which satisfy conditions of the form (2.5) and (2.6). For brevity, we study the case m=1 and a1(x,xϵ)=a(x,xϵ)I, where I is the identity matrix.
Denote
G(˜x)=∫[0,1)d−1√|∇˜yF(˜x,˜y)|2g(˜x,˜y)d˜y, | (2.8) |
P(˜x)=∫[0,1)d−1√|∇˜yF(˜x,˜y)|2p(˜x,˜y)d˜y. | (2.9) |
Note that P(˜x) is positive due to the positiveness of p. We have the convergences (see [1] and Section 5 of this paper)
ϵ1−α∫Γϵ1gi(˜x,˜xϵ)⋅υ(˜x,ϵαF(˜x,˜xϵ))ds→∫Γ1Gi(˜x)⋅υ(x)ds, | (2.10) |
and
ϵ1−α∫Γϵ1pi(˜x,˜xϵ)υ(˜x,ϵαF(˜x,˜xϵ))ds→∫Γ1Pi(˜x)υ(x)ds, | (2.11) |
for each υ∈H1(Dϵ) by ϵ→0. Here, ds is the element of (d−1)-dimensional measure on the hypersurface.
In the further analysis we use the following notation for the spaces U:=[L2(D)]n, Uϵ:=[L2(Dϵ)]n, W:=[H1(D,Γ2)]n, Wϵ:=[H1(Dϵ;Γ2)]n. The norms in our spaces are defined in the following way:
‖v‖2:=∫Dn∑i=1|vi(x)|2dx,‖v‖2ϵ:=∫Dϵn∑i=1|vi(x)|2dx,‖v‖21:=∫Dn∑i=1|∇vi(x)|2dx,‖v‖21,ϵ:=∫Dϵn∑i=1|∇vi(x)|2dx. |
Denote by W′ the dual space to the space W, and by W′ϵ the dual space to the space Wϵ.
Let qk=pk(pk−1) for any k=1,…,n. We use the notation p=(p1,…,pn) and q=(q1,…,qn), and define spaces
Vp:=Lp1(D)×…×Lpn(D),Vp,ϵ:=Lp1(Dϵ)×…×Lpn(Dϵ), |
Vp(R+;Vp):=Lp1(R+;Lp1(D))×…×Lpn(R+;Lpn(D)), |
Vp(R+;Vp,ϵ):=Lp1(R+;Lp1(Dϵ))×…×Lpn(R+;Lpn(Dϵ)). |
As in [21,29], we investigate weak (generalized) solutions of the problem (2.1), that is, functions
uϵ(x,t)∈Vloc∞(R+;Uϵ)∩Vloc2(R+;Wϵ)∩Vlocp(R+;Vp,ϵ), |
which satisfy the Eq (2.1) in the distributional sense (the sense of generalized functions), that is, the integral identity
−∫Dϵ×R+uϵ⋅∂ψ∂t dxdt+∫Dϵ×R+A∇uϵ⋅∇ψ dxdt+∫Dϵ×R+aϵ(x)f(uϵ)⋅ψ dxdt+ϵβ∫Γϵ1×R+p(˜x,˜xϵ)uϵ⋅ψdsdt=∫Dϵ×R+hϵ(x)⋅ψ dxdt+ϵ1−α∫Γϵ1×R+g(˜x,˜xϵ)⋅ψdsdt, | (2.12) |
for each function ψ∈C∞0(R+;Wϵ∩Vp,ϵ). Here, z1⋅z2 denotes the scalar product of vectors z1,z2∈Rn.
If uϵ(x,t)∈Vp(0,M;Vp,ϵ), then from the condition (2.5) it follows that f(u(x,t))∈Vq(0,M;Vq,ϵ). At the same time, if uϵ(x,t)∈V2(0,M;Wϵ), then AΔuϵ(x,t)+hϵ(x)∈V2(0,M;W′ϵ). Therefore, for an arbitrary generalized solution uϵ(x,s) to problem (2.1), it satisfies
∂uϵ(x,t)∂t∈Vq(0,M;Vq,ϵ)+V2(0,M;W′ϵ). |
Now, applying the Sobolev theorems, we get the following:
Vq(0,M;Vq,ϵ)+V2(0,M;W′ϵ)⊂Vq(0,M;U−rϵ). |
Here U−rϵ:=H−r1(Dϵ)×…×H−rn(Dϵ), r=(r1,…,rn) and ri=max{1,d(1/qi−1/2)} for i=1,…,n, where H−r(Dϵ) denotes the space dual to the Sobolev space Hr(Dϵ) with superscript r>0 in the domain Dϵ.
Therefore, for all generalized (weak) solution uϵ(x,t) to problem (2.1), time derivative ∂uϵ(x,t)∂t belongs to Vq(0,M;U−rϵ).
Remark 2.2. Existence of a generalized solution u(x,t) to problem (2.1) for any initial data U∈Uϵ and fixed ϵ, can be proved in the standard way (see, for instance, [20], [29]). This solution may not be unique, since the function f(v) satisfies only the conditions (2.5) and (2.6) and it is not assumed that the Lipschitz condition is satisfied with respect to v.
The next lemma is proved in a similar way to the proposition XV.3.1 from [21].
Lemma 2.1. Let uϵ(x,t)∈Vloc2(R+;Wϵ)∩Vlocp(R+;Vp,ϵ) be the generalized solution of problem (2.1). Then,
(i) uϵ∈C(R+;Uϵ);
(ii) function ‖uϵ(⋅,t)‖2 is absolutely continuous on R+, and moreover
12ddt‖uϵ(⋅,t)‖2+∫DϵA∇uϵ(x,t)⋅∇uϵ(x,t)dx+∫Dϵaϵ(x)f(uϵ(x,t))⋅uϵ(x,t)dx+ϵβ∫Γϵ1p(˜x,˜xϵ)uϵ(x,t)⋅uϵ(x,t)ds=∫Dϵhϵ(x)⋅uϵ(x,t)dx+ϵ1−α∫Γϵ1g(˜x,˜xϵ)⋅uϵ(x,t)ds, | (2.13) |
for a. a. t∈R+.
To define the trajectory space T+ϵ for Eq (2.1), we use the general approaches of Section 3, and for every [t0,t1]∈R, we have the Banach spaces
Gt0,t1:=V2(t0,t1;W)∩V∞(t0,t1;U)∩Vp(t0,t1;Vp)∩{v | ∂v∂t∈Vq(t0,t1;U−r)}, |
(sometimes we omit the parameter ϵ for brevity) with the following norm:
‖w‖Gt0,t1:=‖w‖V2(t0,t1;W)+‖w‖Vp(t0,t1;Vp)+‖w‖V∞(0,M;U)+‖∂w∂t‖Vq(t0,t1;U−r). |
Letting Dt0,t1=Vq(t0,t1;U−r), we obtain Gt0,t1⊆Dt0,t1, and for u(t)∈Gt0,t1, we have L(u(t))∈Dt0,t1. One considers now the generalized solutions to Eq (2.1) as solutions of the equation in the general scheme of Section 3.
Consider the following spaces:
Gloc+=Vloc2(R+;W)∩Vlocp(R+;Vp)∩Vloc∞(R+;U)∩{w | ∂w∂t∈Vlocq(R+;U−r)}, |
Glocϵ,+=Vloc2(R+;Wϵ)∩Vlocp(R+;Vp,ϵ)∩Vloc∞(R+;Uϵ)∩{w | ∂w∂t∈Vlocq(R+;U−rϵ)}. |
We introduce the following notation. Let K+ϵ be the set of all generalized solutions to Eq (2.1). For any U∈U, there exists at least one trajectory u(⋅)∈T+ϵ such that u(0)=U(x). Hence, the space T+ϵ to Eq (2.1) is not empty.
It is easy to see that T+ϵ⊂Glocϵ,+ and the space T+ϵ is translation invariant, i.e., if u(t)∈T+ϵ, then u(τ+t)∈T+ϵ for all τ≥0. Hence, S(τ)T+ϵ⊆T+ϵ for all τ≥0.
In the set Gt0,t1 we can introduce metrics ρt0,t1(⋅,⋅) in Gt0,t1 by means of V2(t0,t1;U)–norms. Hence, we obtain the following definition of this metric:
ρt0,t1(v,w)=(t1∫t0‖v(t)−w(t)‖2Udt)1/2∀v(⋅),w(⋅)∈Ft0,t1. |
The topology Θloc+ in Gloc+ is generated by these metrics. Let us recall that {vk}⊂Gloc+ converges to v∈Gloc+ as k→∞ in Θloc+ if ‖vk(⋅)−v(⋅)‖V2(t0,t1;U)→0(k→∞) for all [t0,t1]⊂R+. Bearing in mind Eq (3.2), we conclude that the topology Θloc+ is metrizable. We consider this topology in T+ϵ of Eq (2.1). Similarly, we define the topology Θlocϵ,+ in Glocϵ,+.
Consider the semigroup of translation {S(τ)} on T+ϵ, S(τ):T+ϵ→T+ϵ, τ≥0. This semigroup {S(τ)} acting on T+ϵ, is continuous in the topology Θlocϵ,+.
Using the scheme from Section 3, one can define bounded sets in T+ϵ by means of the Banach space Gbϵ,+. We naturally get
Gbϵ,+=Vb2(R+;Wϵ)∩Vbp(R+;Vp,ϵ)∩V∞(R+;Uϵ)∩{w | ∂w∂t∈Vbq(R+;U−rϵ)}, |
and the space Gbϵ,+ is a subspace of Glocϵ,+.
Suppose that Tϵ is the kernel to Eq (2.1), i.e., we have the set of all generalized complete bounded solutions u(t),t∈R, to our RD-system. We consider solutions bounded in
Gbϵ=Vb2(R;Wϵ)∩Vbp(R;Vp,ϵ)∩V∞(R;Uϵ)∩{w | ∂w∂t∈Vbq(R;U−rϵ)}. |
Proposition 2.1. Problem (2.1) has the trajectory attractors Aϵ in the topological space Θlocϵ,+. The set Aϵ is bounded in Gbϵ,+ and compact in Θlocϵ,+. In addition, Aϵ=Π+Kϵ, and the kernel Kϵ is nonempty and bounded in Gbϵ. Recall that the spaces Gbϵ,+ and Θlocϵ,+ depend on ϵ.
To prove this proposition, we use the approach of the proof from [21]. To prove the existence of an absorbing set (bounded in Fbϵ,+ and compact in Θlocϵ,+), one can use Lemma 2.1 similar to [21].
It is easy to verify that Aϵ⊂B0(R) for all ϵ∈(0,1). Here, B0(R) is a ball in Gbϵ,+ with a sufficiently large radius R. Due to Lemma 3.1, we have
B0(R)⋐Vloc2(R+;U1−ηϵ), | (2.14) |
B0(R)⋐Cloc(R+;U−ηϵ),0<η≤1. | (2.15) |
Bearing in mind Eqs (2.14) and (2.15), the attraction to the constructed trajectory attractor can be strengthened.
Corollary 2.1. For any bounded in Gbϵ,+ set B⊂T+ϵ we get
distV2(0,M;U1−ηϵ)(Π0,MS(τ)B,Π0,MTϵ)→0,distC([0,M];U−ηϵ)(Π0,MS(τ)B,Π0,MTϵ)→0(τ→∞), |
where M is a positive constant.
Recall that D⊂Dϵ and D lies in the positive half-space {xd>0}. Therefore, for each function u(x,t) of the variable x∈Dϵ that belongs to the space Fbϵ,+, its restriction to the domain D belongs to the space Fb+ and, moreover,
‖u‖Gb+≤‖u‖Gbϵ,+. |
Using this observation, we have:
Corollary 2.2. The trajectory attractors Aϵ are uniformly (with respect to ϵ∈(0,1)) bounded in Fb+. It should be noted that the kernels Kϵ are uniformly bounded in the space Gb. We mean that they are uniformly bounded with respect to ϵ∈(0,1).
The section is devoted to the trajectory attractors to autonomous evolutionary equations (see details in [21]).
Consider an autonomous equation of the form
∂u∂t=L(u),t≥0. | (3.1) |
Here, L(⋅):Υ1→Υ0 is a nonlinear mapping, Υ1,Υ0 are Banach spaces, and Υ1⊆Υ0. As an example, one can consider L(u)=AΔu−a(⋅)f(u)+h(⋅).
We study generalized solutions u(t) to Eq (3.1) as functions of t∈R+ as an object. The set of solutions of Eq (3.1) is said to be a trajectory space T+ of Eq (3.1). Now, we give a detailed description of T+.
Consider solutions u(t) of Eq (3.1) on [t0,t1]⊂R. We consider solutions to problem (3.1) in a Banach space Gt0,t1. The space Gt0,t1 is a set f(s),s∈[t0,t1] satisfying f(t)∈Υ for a.a. t∈[t0,t1], where Υ is a Banach space, satisfying Υ1⊆Υ⊆Υ0.
We consider Gt0,t1 as the intersection of spaces C([t0,t1];E), or Lp(t0,t1;E), forp∈[1,∞]. Suppose that Rt0,t1Gτ0,τ1⊆Gt0,t1 and
‖Rt0,t1f‖Gt0,t1≤C(t0,t1,τ0,τ1)‖f‖Gτ0,τ1∀f∈Gτ1,τ2. |
Here, [t0,t1]⊆[τ0,τ1] and we denote by Rt0,t1 the restriction operator onto [t0,t1], where C(t0,t1,τ0,τ1) is independent of f.
Denote by S(τ) for τ∈R the translation S(τ)f(t)=f(τ+t). If the variable t of f(⋅) belongs to the segment [t0,t1], then the variable t of S(τ)f(⋅) belongs to [t0−τ,t1−τ] for τ∈R. Suppose that S(τ) is an isomorphism from Ft0,t1 to Ft0−τ,t1−τ and ‖S(τ)f‖Gt0−τ,t1−τ=‖f‖Gt0,t1 ∀f∈Gt0,t1.
Suppose that if f(t)∈Gt0,t1, then L(f(t))∈Dt0,t1, where Dt0,t1 is a Banach space, which is larger, Gt0,t1⊆Dt0,t1. The derivative ∂f(t)∂t is a distribution with values in Υ0,∂f∂t∈D′((t0,t1);Υ0), and we suppose that Dt0,t1⊆D′((t0,t1);Υ0) for all (t0,t1)⊂R. A function u(t)∈Gt0,t1 is a solution of Eq (3.1) if ∂u∂t(t)=L(u(t)) in the sense of D′((t0,t1);Υ0).
Consider the space
Gloc+={f(t),t∈R+|Rt0,t1f(t)∈Gt0,t1,∀[t0,t1]⊂R+}. |
For instance, if Gt0,t1=C([t0,t1];E), then Gloc+=C(R+;E), and if Gt0,t1=Lp(t0,t1;E), then Gloc+=Llocp(R+;E).
A function u(t)∈Gloc+ is a solution of Eq (3.1) if Rt0,t1u(t)∈Gt0,t1, and u(t) is a solution to Eq (3.1) for any [t0,t1]⊂R+.
Let T+ be a set of solutions to Eq (3.1) from Gloc+. Note that T+ in general is not the set of all solutions from Gloc+. The set T+ consists on elements, which are trajectories, and the set T+ is the trajectory space of the Eq (3.1).
Suppose that the trajectory space T+ is translation invariant, i.e., if u(t)∈T+, then u(τ+t)∈T+ for every τ≥0.
Consider the translations S(τ) in Gloc+: S(τ)f(t)=f(τ+t), τ≥0. It is easy to see that the map {S(τ),τ≥0} forms a semigroup in Gloc+:S(τ1)S(τ2)=S(τ1+τ2) for τ1,τ2≥0, and, in addition, S(0) is the identity operator. The semigroup {S(τ),τ≥0} maps the trajectory space T+ to itself: S(τ)T+⊆T+ for all τ≥0.
We investigate attracting properties of the translation semigroup {S(τ)} acting on the trajectory space T+⊂Gloc+. Next step is to get a topology in Gloc+.
Assume that some metrics ρt0,t1(⋅,⋅) are defined on Gt0,t1 for any [t0,t1]⊂R. Suppose that
ρt0,t1(Rt0,t1f,Rt0,t1g)≤D(t0,t1,τ0,τ1)ρτ0,τ1(f,g)for all f,g∈Gτ0,τ1, [t0,t1]⊆[τ0,τ1], |
ρt0−τ,t1−τ(S(τ)f,S(τ)g)=ρt0,t1(f,g)∀f,g∈Gt0,t1, [t0,t1]⊂R, τ∈R. |
Now, we denote by Θt0,t1 the metric spaces on Gt0,t1. For instance, ρt0,t1 is the metric defined by the norm ‖⋅‖Gt0,t1 of Gt0,t1.
The projective limit of the spaces Θt0,t1 defines the topology Θloc+ in Gloc+, that is, by definition, a sequence {fk(t)}⊂Gloc+ goes to f(t)∈Gloc+ as k→∞ in Θloc+ if ρt0,t1(Rt0,t1fk,Rt0,t1f)→0 as k→∞ for all [t0,t1]⊂R+. It is possible to show that the topology Θloc+ is metrizable. For this aim we use, for instance, the Fréchet metric
ρ+(f1,f2):=∑k∈N2−kρ0,k(f1,f2)1+ρ0,k(f1,f2). | (3.2) |
We define the Banach space
Gb+:={f(t)∈Gloc+|‖f‖Fb+<+∞}, |
with
‖f‖Gb+:=supτ≥0‖R0,1f(τ+t)‖G0,1. |
We recall that Gb+⊆Θloc+. For our Banach space Gb+, we need only the fact that it should define bounded subsets in the trajectory space T+.
Assume that T+⊆Gb+.
Definition 3.1. A set Ξ⊆Θloc+ is said to be the attracting set of {S(τ)} acting on T+ in the topology Θloc+ if for any bounded in Fb+ set B⊆T+, the set Ξ attracts S(τ)B as τ→+∞ in the topology Θloc+, i.e., for any ϵ-neighbourhood Oϵ(Ξ) in Θloc+ there is τ1≥0 such that S(τ)B⊆Oϵ(Ξ) for all τ≥τ1.
It is easy to see that the attracting property of Ξ can be reformulated equivalently: we have
distΘ0,M(R0,MS(τ)B,R0,MΞ)⟶0(τ→+∞). |
Here,
distM(X,Y):=supx∈XdistM(x,Y)=supx∈Xinfy∈YρM(x,y), |
is the Hausdorff semi-distance from a set X to a set Y in a metric space M. We recall that the Hausdorff semi-distance is not symmetric, for any B⊆T+ bounded in Gb+ and for all M>0.
Definition 3.2. ([21]). A set A⊆T+ is said to be the trajectory attractor of the semigroup {S(τ)} on T+ in the topology Θloc+ if
(i) A is compact in Θloc+ and bounded in Fb+,
(ii) the set A is invariant: S(τ)A=A for all τ≥0,
(iii) the set A is an attracting for {S(τ)} on T+ in the topology Θloc+, i.e., for every M>0, we have
distΘ0,M(R0,MS(τ)B,R0,MA)→0(τ→+∞). |
Let us give the main assertion on the trajectory attractor for Eq (3.1).
Theorem 3.1. ([20,21]). Let the trajectory space T+ corresponding to Eq (3.1) be contained in Gb+. We also assume that our semigroup {S(t)} has an attracting set Ξ⊆T+ which is bounded in Gb+ and compact in Θloc+. Then, the semigroup {S(τ),τ≥0} acting on T+ has the trajectory attractor A⊆Ξ. The set A is compact in Θloc+ and bounded in Gb+.
Let us describe in detail, i.e., in terms of complete trajectories of the equation, the structure of the trajectory attractor A to Eq (3.1). We study Eq (3.1) on the time axis
∂u∂t=L(u),t∈R. | (3.3) |
Note that the trajectory space T+ of Eq (3.3) on R+ has been defined. We need this notion on the entire R. If a function f(t),t∈R, is defined on the entire axis, then S(τ)f(t)=f(τ+t) are also defined for τ<0. A function u(t),t∈R, is a complete trajectory of Eq (3.3) if R+u(τ+t)∈T+ for all τ∈R. Here, R+=R0,∞ denotes the restriction operator to R+.
We have Gloc+,Gb+, and Θloc+. Let us define spaces Gloc,Gb, and Θloc in the same way
Gloc:={f(t),t∈R|Rt0,t1f(s)∈Gt0,t1∀[t0,t1]⊆R}, |
Gb:={f(t)∈Gloc|‖f‖Gb<+∞}, |
where
‖f‖Gb:=suph∈R‖R0,1f(τ+t)‖G0,1. | (3.4) |
Note that our topological space Θloc coincides (the coincidence as a set) with Gloc and, by definition, fk(t)→f(t)(k→∞) in Θloc if Rt0,t1fk(t)→Rt0,t1f(t)(k→∞) in Θt0,t1 for each [t0,t1]⊆R.
Definition 3.3. The kernel T in Gb of Eq (3.3) is the collection of all complete trajectories u(t),t∈R, of Eq (3.3), bounded in Gb w.r.t. the norm Eq (3.4), i.e.,
‖R0,1u(τ+t)‖G0,1≤Cu∀τ∈R. |
Theorem 3.2. Suppose the assumptions of the previous theorem hold. Then, A=R+T and the set T is bounded in Gb and compact in Θloc.
To prove this assertion, one can use the approach from [21].
Now we are going to prove that a ball in the space Gb+ is compact in our topological space Θloc+. For this aim we use the next lemma. Assume that Υ0 and Υ1 are Banach spaces and Υ1⊂Υ0. We consider the spaces
Wp1,p0(0,M;Υ1,Υ0)={ξ(t),t∈0,M|ξ(⋅)∈Lp1(0,M;Υ1),ξ′(⋅)∈Lp0(0,M;Υ0)}, |
W∞,p0(0,M;Υ1,Υ0)={ξ(t),t∈0,M|ξ(⋅)∈L∞(0,M;Υ1),ξ′(⋅)∈Lp0(0,M;Υ0)}, |
(p1≥1, p0>1) with the norms
‖ξ‖Wp1,p0:=(M∫0‖ξ(t)‖p1Υ1dt)1/p1+(M∫0‖ξ′(t)‖p0Υ0dt)1/p0, |
‖ξ‖W∞,p0:=ess sup{‖ξ(t)‖Υ1|t∈[0,M]}+(M∫0‖ξ′(t)‖p0Υ0dt)1/p0. |
Lemma 3.1. ([30]). Suppose that Υ1⋐Υ⊂Υ0. Then, we have compact embeddings
Wp1,p0(0,T;Υ1,Υ0)⋐Lp1(0,T;Υ),W∞,p0(0,T;Υ1,Υ0)⋐C([0,T];Υ). |
In this paper we investigate evolutionary equations and their attractors that depend on a small parameter ϵ>0.
Definition 3.4. The trajectory attractors Aϵ tend to the trajectory attractor ¯A as ϵ→0 in the topology Θloc+ if for every vicinity O(¯A) in Θloc+ there exists an ϵ1≥0 such that Aϵ⊆O(¯A) for all ϵ<ϵ1, i.e., for every M>0, we have
distΘ0,M(R0,MAϵ,R0,M¯A)→0(ϵ→0). |
In the next sections, we study the behaviour of the problem (2.1) as ϵ→0 in the critical case β=1−α. We have the following "formal" limit problem with an inhomogeneous Fourier boundary condition
{∂u0∂t=AΔu0−¯a(x)f(u0)+¯h(x),x∈D,t>0,∂u0∂ν+P(˜x)u0=G(˜x),x=(˜x,0)∈Γ1,t>0,u0=0,x∈Γ2,t>0,u0=U(x),x∈D,t=0. | (4.1) |
Here, ¯a(x) and ¯h(x) are defined in Eqs (2.2) and (2.3), respectively, and G(˜x) and P(˜x) were defined in Eqs (2.8) and (2.9).
As before, we consider generalized solutions of the problem (4.1), that is, functions
u0(x,t)∈Vloc∞(R+;U)∩Vloc2(R+;W)∩Vlocp(R+;Vp), |
which obey the integral identity
−∫D×R+u0⋅∂ξ∂t dxdt+∫D×R+A∇u0⋅∇ξ dxdt+∫D×R+ˉa(x)f(u0)⋅ξ dxdt+∫Γ1×R+P(˜x)u0⋅ξdsdt=∫D×R+ˉh(x)⋅ξ dxdt+∫Γ1×R+G(˜x)⋅ξdsdt, | (4.2) |
for each function ξ∈C∞0(R+;W∩Vp). For each u(x,t) to Eq (4.1), we have that ∂u0(x,t)∂t∈Vq(0,M;U−r) (see Section 2). Recall that the "limit" domain D in Eqs (4.1) and (4.2) is independent of ϵ and its boundary contains the plain part Γ1.
Similar to Eq (2.1), for any initial data U∈U, the problem (4.1) has at least one generalized solution (see Remark 2.2). Lemma 2.1 also holds true for the problem (4.1) with replacing the ϵ-depending coefficients a,h,p, and g by the corresponding averaged coefficients ¯a(x),¯h(x),P(˜x), and G(˜x).
As usual, let ¯T+ be the the trajectory space for Eq (4.1) (the set of all generalized solutions) that belongs to the corresponding spaces Gloc+ and Gb+ (see Section 3). Recall that ¯T+⊂Gloc+ and the space ¯T+ is translation invariant with respect to translation semigroup {S(τ)}, that is, S(τ)¯T+⊆¯T+ for all τ≥0. We now construct the trajectory attractor in the topology Θloc+ for the problem (4.1) (see Sections 2 and 3).
Similar to Proposition 2.1, we have:
Proposition 4.1. Problem (4.1) has the trajectory attractor ¯A in the topological space Θloc+. The set ¯A is bounded in Gb+ and compact in Θloc+. Moreover,
¯A=R+¯K, |
and the kernel ¯K of the problem (4.1) is nonempty and bounded in Gb.
We also have ¯A⊂B0(R), where B0(R) is a ball in Fb+ with a sufficiently large radius R. Finally, the analog of Corollary 2.1 holds for the trajectory attractor ¯A.
Corollary 4.1. For any bounded in Gb+ set B⊂¯T+, we have
distV2(0,M;U1−η)(R0,MS(τ)B,R0,M¯T)→0,distC([0,M];U−ηϵ)(R0,MS(τ)B,R0,M¯T)→0(τ→∞),∀M>0. |
Next lemmata are proved in [1].
Lemma 5.1. The convergence
v(˜x,ϵαF(˜x,˜xϵ))→v(˜x,0)asϵ→0, | (5.1) |
is strongly in [L2(Γ1)]n and the inequality
‖v‖[L2(Rϵ)]n≤C1√ϵα‖v‖Wϵ, | (5.2) |
is true for any v∈Wϵ.
Let us consider auxiliary elliptic problems
{AΔvϵ+h(x,xϵ)=0,x∈Dϵ,∂vϵ∂ν+ϵβp(˜x,˜xϵ)vϵ=ϵ1−αg(˜x,˜xϵ),x=(˜x,xd)∈Γϵ1,vϵ=0,x∈Γ2, | (5.3) |
and
{AΔv0+¯h(x)=0,x∈D,∂v0∂ν+P(˜x)v0=G(˜x),x=(˜x,0)∈Γ1,v0=0,x∈Γ2, | (5.4) |
where ¯h(x) is defined in Eq (2.3), and G(˜x) and P(˜x) are defined in Eqs (2.8) and (2.9).
Lemma 5.2. For all v∈Wϵ, the convergence
|ϵ1−α∫Γϵ1g(˜x,˜xϵ)⋅v(˜x,ϵαF(˜x,˜xϵ))ds−∫Γ1G(˜x)⋅v(x)ds|→0, | (5.5) |
is valid as ϵ→0.
Lemma 5.3. The convergence
|ϵ1−α∫Γϵ1p(˜x,˜xϵ)v0(˜x,ϵαF(˜x,˜xϵ))⋅v(˜x,ϵαF(˜x,˜xϵ))ds−∫Γ1P(˜x)v0(x)⋅v(x)ds|→0, | (5.6) |
takes place as ϵ→0. Here, v0 is a solution to Eq (5.4) and v∈Wϵ.
Remark 5.1. Due to the smoothness of the boundary ∂D, the solution v0 belongs to H2(D) [31], and, hence, can be continued on Rϵ to belong to H2(Dϵ) [32].
Lemma 5.4. Let β=1−α and F(˜x,˜y), g(˜x,˜y), p(˜x,˜y) be periodic in y smooth functions, A be a given matrix, and h(x,xϵ) be a righthand function which satisfies the conditions (2.3) and (2.4). Suppose that F(˜x,˜y) is compactly supported in x∈Γ1 uniformly in y. Then, for all ϵ>0, the existence and uniqueness of the solution to problem (5.3) follows, and the strong convergence
vϵ→v0, | (5.7) |
in W as ϵ→0 is valid.
Proof. The existence and uniqueness of vϵ (v0) are due to the positiveness of p(˜x,˜xϵ) (P(˜x)) and the Lax-Milgram lemma (see [33]). Then, according to Eqs (2.1) and (4.1),
∫DϵA∇(v0−vϵ)⋅∇wdx+ϵ1−α∫Γϵ1p(v0−vϵ)⋅wds=∫DA∇v0⋅∇wdx−∫Dϵh⋅wdx−ϵ1−α∫Γϵ1g⋅wds+ϵ1−α∫Γϵ1pv0⋅wds=∫DA∇v0⋅∇wdx−∫Dϵh⋅wdx−ϵ1−α∫Γϵ1g⋅wds+∫RϵA∇v0⋅∇wdx+ϵ1−α∫Γϵ1pv0⋅wds=∫RϵA∇v0⋅∇wdx−ϵ1−α∫Γϵ1g⋅wds+∫Γ1G(˜x)⋅wds−∫Rϵh⋅wdx+ϵ1−α∫Γϵ1pv0⋅wds−∫Γ1P(˜x)v0⋅wds. |
Let us estimate all the terms in the righthand side of the last relation. By Eq (5.2), considering the smoothness of v0, we have
|∫RϵA∇v0⋅∇wdx|≤‖A‖‖∇v0‖[L2(Rϵ)]n‖w‖Wϵ≤C2√ϵα‖v0‖Wϵ‖w‖Wϵ, |
and
|∫Rϵh⋅wdx|≤‖h‖[L2(Rϵ)]n‖w‖[L2(Rϵ)]n≤C3√ϵα‖h‖Uϵ‖w‖Wϵ. |
Then, according to Lemmas 5.2 and 5.3, the inequalities
|ϵ1−α∫Γϵ1g⋅wds−∫Γ1G⋅wds|≤C4(ϵ1−α+√ϵα)‖w‖Wϵ, |
and
|ϵ1−α∫Γϵ1pv0⋅wds−∫Γ1Pv0⋅wds|≤C5(ϵ1−α+√ϵα)‖v0‖Wϵ‖w‖Wϵ, |
hold. With the help of these inequalities, we obtain
|∫DϵA∇(v0−vϵ)⋅∇wdx+∫Γϵ1p(v0−vϵ)⋅wds|≤C6(ϵ1−α+√ϵα)‖w‖Wϵ. |
Substituting w=v0−vϵ and using Lemma 5.3 and the Friedrichs type inequality (see [34,35,36]), we obtain Eq (5.7). The lemma is proved.
Lemma 5.5. 1) All solutions uϵ(t) to Eq (2.1) satisfy
‖uϵ(t)‖2ϵ≤‖uϵ(0)‖2ϵe−ϰ1t+R21, | (5.8) |
ϖ∫t+1t‖uϵ(s)‖2ϵ,1ds+2a0N∑i=1γit+1∫t‖uiϵ(s)‖piLpi(Dϵ)ds++2pmaxϵ1−αt+1∫t‖uϵ(s)‖2V2(Γϵ1)ds≤‖uϵ(t)‖2ϵ+R22, | (5.9) |
where ϰ1>0 is a constant independent of ϵ. Positive values R1 and R2 depend on M0 (see Eq (2.4)) and are independent of uϵ(0) and ϵ.
2) All solutions u(t) to Eq (4.1) satisfy the same inequalities (5.8) and (5.9) with the norms in the function spaces over the domain D instead Dϵ.
Proof. We give a brief outline of the proof (see the details in [21]).
In the righthand side of Eq (2.13), the integral over the part of the boundary Γϵ1 is nonnegative because of the positiveness of the matrix p. We integrate Eq (2.13) with respect to t. Then, to estimate the terms
ϵ1−α∫Γϵ1g⋅wds andϵ1−α∫Γϵ1puϵ⋅wds, |
we use the Cauchy inequality and the compactness of embedding V2(Γϵ1)⋐Wϵ. For other terms, we use a standard procedure (see [21]). The lemma is proved.
Here, we formulate the main result concerning the limiting behavior of the trajectory attractors Aϵ of the systems (2.1) as ϵ→0 in the critical case β=1−α.
Theorem 6.1. The following limit holds in the topological space Θloc+
Aϵ→¯Aasϵ→0+. | (6.1) |
Moreover,
Kϵ→¯Kasϵ→0+inΘloc. | (6.2) |
Proof. It is easy to see that Eq (6.2) implies Eq (6.1). Hence, it is sufficient to prove Eq (6.2), i.e., for every neighborhood O(¯K) in Θloc, there exists ϵ1=ϵ1(O)>0 such that
Kϵ⊂O(¯K) for ϵ<ϵ1. | (6.3) |
Assume that Eq (6.3) is not true. Then, there exists a neighborhood O′(¯K) in Θloc, a sequence ϵk→0+(k→∞), and a sequence uϵk(⋅)=uϵk(t)∈Kϵk, such that
uϵk∉O′(¯K) for all k∈N. |
The function uϵk(x,t),t∈R is a solution to
{∂uϵk∂t=AΔuϵk−a(x,xϵk)f(uϵk)+h(x,xϵk),x∈Dϵk,∂uϵk∂ν+ϵβkp(˜x,˜xϵk)uϵk=ϵ1−αkg(˜x,˜xϵk),x∈Γϵk1,uϵk=0,x∈Γ2, | (6.4) |
where β=1−α. To get the uniform in ϵ estimate of the solution, we use Lemma 5.5 (see also Corollary 4.1). By means of Eqs (5.8) and (5.9), we obtain that the sequence {uϵk(x,t)} is bounded in Gb, i.e.,
‖uϵk‖Gb=supt∈R‖uϵk(t)‖+supt∈R(t+1∫t‖uϵk(ϑ)‖21dϑ)1/2+supt∈R(t+1∫t‖uϵk(ϑ)‖pVpdϑ)1/p+ϵβsupt∈Rt+1∫t∫Γϵ1p(˜x,˜xϵ)uϵ(x,ϑ)⋅uϵ(x,ϑ)dsdϑ+supt∈R(t+1∫t‖∂uϵk∂t(ϑ)‖qU−rdϑ)1/q≤C ∀k∈N. | (6.5) |
Note that here, β=1−α. The constant C is independent of ϵ. Consequently, there exists a subsequence {uϵ′k(x,t)}⊂{uϵk(x,t)}, such that
uϵ′k(x,t)→¯u(x,t)as k→∞ in Θloc. |
Here, ¯u(x,t)∈Gb and ¯u(t) satisfies Eq (6.5) with the same constant C. Because of Eq (6.5), we get
uϵ′k(x,t)⇀¯u(x,t)(k→∞), |
weakly in Vloc2(R;W), weakly in Vlocp(R;Vp), star-weakly in Vloc∞(R+;U) and
∂uϵ′k(x,t)∂t⇀∂¯u(x,t)∂t(k→∞), |
weakly in Vlocq,w(R;U−r). We claim that ¯u(x,t)∈¯K. We have ‖¯u‖Gb≤C. Hence, we have to verify that ¯u(x,t)=u0(x,t), i.e. it is a generalized solution to Eq (4.1).
Using Eqs (6.5) and (2.3), we find that
∂uϵk∂t−AΔuϵk−hϵk(x)⟶∂ˉu∂t−AΔˉu−¯h(x) as k→∞, | (6.6) |
in the space D′(R;U−r), since the derivative operators are continuous in the space of distributions.
Let us prove that
a(x,xϵk)f(uϵk)⇀ˉa(x)f(ˉu) as k→∞, | (6.7) |
weakly in Vlocq,w(R;Vq). We fix any number M>0. The sequence {uϵk(x,t)} is bounded in Vp(−M,M;Vp) (see Eq (6.5)). Then, due to Eq (2.5), the sequence {f(uϵk(t))} is bounded in Vq(−M,M;Vq). Because {uϵk(x,t)} is bounded in V2(−M,M;W) and {∂uϵk∂t(t)} is bounded in Vq(−M,M;U−r), we may assume that
uϵk(x,t)→ˉu(x,t) as k→∞ in V2(−M,M;V2)=V2(D×]−M,M[), |
hence,
uϵk(x,t)→ˉu(x,t) as k→∞ for almost all (x,t)∈D×]−M,M[. |
Because the function f(w) is continuous in w∈R, we conclude that
f(uϵk(x,t))→f(ˉu(x,t)) as k→∞ for almost all (x,t)∈D×]−M,M[. | (6.8) |
We have
a(x,xϵk)f(uϵk)−ˉa(x)f(ˉu)= a(x,xϵk)(f(uϵk)−f(ˉu))+(a(x,xϵk)−ˉa(x))f(ˉu). | (6.9) |
Let us show that both terms in the righthand side of Eq (6.9) tend to zero as k→∞ weakly in Vq(−M,M;Vq)=Vq(D×]−M,M[). First, the sequence a(x,xϵk)(f(uϵk)−f(ˉu)) goes to zero as k→∞ for almost all (x,t)∈D×]−M,M[ (see Eq (6.8)). Applying Lemma 1.3 from [37], we conclude that
a(x,xϵk)(f(uϵk)−f(ˉu))⇀0 as k→∞, |
weakly in Vq(D×]−M,M[). Second, the sequence (a(x,xϵk)−ˉa(x))f(ˉu) also goes to zero a k→∞ weakly in Vq(D×]−M,M[), since a(x,xϵk)⇀ˉa(x) as k→∞ star-weakly in V∞,∗w(−M,M;V2) and f(ˉu)∈Vq(D×]−M,M[). Thus, Eq (6.7) is proved.
Now, let us show that
ϵ1−αkp(˜x,˜xϵk)uϵk⇀P(˜x)ˉuask→+∞, | (6.10) |
weakly in V2(Γ1×]−M,M[). Indeed, we have
ϵ1−αkp(˜x,˜xϵk)uϵk(˜x,ϵαkF(˜x,˜xϵk))−P(˜x)ˉu(x)=ϵ1−αkp(˜x,˜xϵk)(uϵk(˜x,ϵαkF(˜x,˜xϵk))−ˉu(˜x,ϵαkF(˜x,˜xϵk)))+ϵ1−αkp(˜x,˜xϵk)ˉu(˜x,ϵαkF(˜x,˜xϵk))−P(˜x)ˉu(˜x,0). |
We have
ϵ1−αkp(˜x,˜xϵk)(uϵk(˜x,ϵαkF(˜x,˜xϵk),t)−ˉu(˜x,ϵαkF(˜x,˜xϵk),t))⇀0ask→+∞, |
weakly in V2(Γ1×]−M,M[) due to Lemma 5.1. We state that
ϵ1−αkp(˜x,˜xϵk)ˉu(˜x,ϵαkF(˜x,˜xϵk),t)−P(˜x)ˉu(˜x,0,t)⇀0ask→+∞, | (6.11) |
weakly in V2(Γ1×]−M,M[). Indeed, due to Lemma 5.5, both terms are bounded V2(Γ1×]−M,M[). Also, one can see that this convergence due to Eq (2.11) is almost everywhere in ]−M,M[. Using Lemma 1.3 from [37], we get the weak convergence Eq (6.11) and, hence, we obtain Eq (6.10).
In an analogous way, we act with the terms with g(˜x,˜xϵk) and G(˜x), using Lemma 5.2.
Hence, for ¯u(x,t)=u0(x,t), we have
−M∫−M∫Dϵkuϵk⋅∂ξ∂t dxdt+M∫−M∫DϵkA∇uϵk⋅∇ξ dxdt+M∫−M∫Dϵkaϵk(x)f(uϵk)⋅ξ dxdt+ |
ϵkβM∫−M∫Γϵk1p(˜x,˜xϵk)uϵk⋅ξdsdt⟶−M∫−M∫Du0⋅∂ξ∂t dxdt+ |
M∫−M∫DA∇u0⋅∇ξ dxdt+∫D¯a(x)f(u0)⋅ξ dxdt+M∫−M∫Γ1P(˜x)u0⋅ξdsdt, |
as k→∞.
Using Eq (6.8), we pass to the limit in the Eq (6.4) as k→∞ in the space D′(R;U−r) and obtain that the function u0(x,t) satisfies the integral identity Eq (4.2) and, hence, it is a complete trajectory of the Eq (4.1).
Consequently, u0∈¯K. We have proved above that uϵk→u0 as k→∞ in Θloc. Assumption uϵk∉O′(¯K) (see [38]) implies u0∉O′(¯K), and, hence, u0∉¯K. We arrive to the contradiction that completes the proof of the theorem.
Using the compact imbedding Eqs (2.14) and (2.15), we improve the convergence Eq (6.1).
Corollary 6.1. For any 0<η≤1 and for all M>0,
distV2([0,M];U1−η)(R0,MAϵ,R0,M¯A)→0, | (6.12) |
distC([0,M];U−η)(R0,MAϵ,R0,M¯A)→0(ϵ→0+). | (6.13) |
To prove Eqs (6.12) and (6.13), we use the reasoning in proof Theorem 6.1, changing the topological space Θloc by Vloc2(R+;U1−η) or Cloc(R+;U−η).
In conclusion, we consider the case of uniqueness of the Cauchy problem for RD-systems. It is sufficient to suppose that the nonlinear function f(u) in Eq (2.1) satisfies the inequality
−C|w1−w2|2≤(f(w1)−f(w2),w1−w2) for any w1,w2∈Rn, | (6.14) |
(see [21,29]). In [29], it was proved that if Eq (6.14) is true, then Eqs (2.1) and (4.1) generate dynamical semigroups in U, possessing that global attractors Aϵ and ¯A are bounded in W (see [19,20]). Moreover,
Aϵ={u(0)|u∈Aϵ},¯A={u(0)|u∈¯A}. |
The convergence Eq (6.13) gives:
Corollary 6.2. Under the assumption of Theorem 6.1, the limit formula takes place
distU−η(Aϵ,¯A)→0(ϵ→0+). |
In the paper, we consider RD-systems with rapidly oscillating terms in equations and in boundary conditions in domains with locally periodic wavering boundary (rough surface) depending on a small parameter. We define the trajectory attractors of these systems and prove that they weakly converge to the trajectory attractors of the limit (averaged) RD-systems in domain independent of the small parameter.
In this paper we consider the critical case in which the type of boundary condition is preserved under the limit passage.
Defining the appropriate axillary functional spaces with weak topology, we prove the existence of trajectory attractors for these systems. Then, we formulate the main theorem and prove it with the help of auxiliary lemmata.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Methodology, G.A.C. and V.V.C.; Formal analysis, K.A.B., G.A.C., and V.V.C.; Investigation, G.F.A., K.A.B., G.A.C., and V.V.C.; Writing—original draft, G.F.A.; Writing—review & editing, G.F.A., K.A.B., G.A.C., and V.V.C. All authors have read and agreed to the published version of the manuscript.
The work of the G. F. Azhmoldaev and K. A. Bekmaganbetov in Sections 2 and 6 is supported by the Committee of Science of the Ministry of Science and Higher Education of the Republic of Kazakhstan (grant AP26199535). The work of G. A. Chechkin in Section 3 and in Section 6 the work was financially supported by the Ministry of Education and Science of the Russian Federation as part of the program of the Moscow Center for Fundamental and Applied Mathematics under the agreement No 075-15-2022-284. The work of V.V. Chepyzhov in Sections 4 and 5 is partially supported by the Russian Science Foundation (project 23-71-30008).
The authors thank the reviewers for comments and recommendations which allowed to improve the presentation of the results.
The authors declare there is no conflict of interest.
[1] |
G. A. Chechkin, A. Friedman, A. L. Piatnitski, The boundary-value problem in domains with very rapidly oscillating boundary, J. Math. Anal. Appl., 231 (1999), 213–234. https://doi.org/10.1006/jmaa.1998.6226 doi: 10.1006/jmaa.1998.6226
![]() |
[2] | V. A. Marchenko, E. Y. Khruslov, Homogenization of Partial Differential Equations, Springer Science & Business Media, Boston (MA): Birkhäuser, 2006. https://doi.org/10.1007/978-0-8176-4468-0 |
[3] | A. G. Belyaev, A. L. Piatnitski, G. A. Chechkin, Asymptotic behavior of a solution to a boundary value problem in a perforated domain with oscillating boundary, Sib. Math. J., 39 (1998), 621–644. |
[4] | G. A. Chechkin, T. A. Mel'nyk, Homogenization of a boundary-value problem in a thick 3-dimensional multilevel junction, Russ. Acad. Sci. Sbornik Math., 200 (2009), 357–383. |
[5] |
D. Borisov, G. Cardone, L. Faella, C. Perugia, Uniform resolvent convergence for a strip with fast oscillating boundary, J. Differ. Equations, 255 (2013), 4378–4402. https://doi.org/10.1016/j.jde.2013.08.005 doi: 10.1016/j.jde.2013.08.005
![]() |
[6] |
G. A. Chechkin, A. McMillan, R. Jones, D. Peng, A computational study of the influence of surface roughness on material strength, Meccanica, 53 (2018), 2411–2436. https://doi.org/10.1007/s11012-018-0830-6 doi: 10.1007/s11012-018-0830-6
![]() |
[7] |
A. Gaudiello, A. Sili, Homogenization of highly oscillating boundaries with strongly contrasting diffusivity, SIAM J. Math. Anal., 47 (2015), 1671–1692. https://doi.org/10.1137/140987225 doi: 10.1137/140987225
![]() |
[8] |
D. I. Borisov, R. R. Suleimanov, Operator estimates for problems in domains with singular curved boundary: Dirichlet and Neumann conditions, Dokl. Math., 109 (2024), 6–11. https://doi.org/10.1134/S1064562424701758 doi: 10.1134/S1064562424701758
![]() |
[9] |
D. I. Borisov, R. R. Suleimanov, On operator estimates for elliptic operators with mixed boundary conditions in two-dimensional domains with fast oscillating boundaries, Math. Notes, 116 (2024), 163–184. https://doi.org/10.1134/S0001434624070149 doi: 10.1134/S0001434624070149
![]() |
[10] |
Y. Amirat, G. A. Chechkin, R. R. Gadyl'shin, Asymptotics of simple eigenvalues and eigenfunctions for the Laplace operator in a domain with an oscillating boundary, Comput. Math. Math. Phys., 46 (2006), 97–110. https://doi.org/10.1134/S0965542506010118 doi: 10.1134/S0965542506010118
![]() |
[11] |
Y. Amirat, G. A. Chechkin, R. R. Gadyl'shin, Asymptotics for eigenelements of Laplacian in domain with oscillating boundary: Multiple eigenvalues, Appl. Anal., 86 (2007), 873–897. https://doi.org/10.1080/00036810701461238 doi: 10.1080/00036810701461238
![]() |
[12] |
Y. Amirat, G. A. Chechkin, R. R Gadyl'shin, Asymptotics of the solution of a Dirichlet spectral problem in a junction with highly oscillating boundary, C.R. Mec., 336 (2008), 693–698. https://doi.org/10.1016/j.crme.2008.06.008 doi: 10.1016/j.crme.2008.06.008
![]() |
[13] |
Y. Amirat, G. A. Chechkin, R. R. Gadyl'shin, Spectral boundary homogenization in domains with oscillating boundaries, Nonlinear Anal. Real World Appl., 11 (2010), 4492–4499. https://doi.org/10.1016/j.nonrwa.2008.11.023 doi: 10.1016/j.nonrwa.2008.11.023
![]() |
[14] | É. Sanchez-Palencia, Homogenization Techniques for Composite Media, Berlin: Springer–Verlag, 1987. |
[15] | O. A. Oleinik, A. S. Shamaev, G. A. Yosifian, Mathematical Problems in Elasticity and Homogenization, Amsterdam: North–Holland, 1992. |
[16] | V. V. Jikov, S. M. Kozlov, O. S. Oleinik, Homogenization of Differential Operators and Integral Functionals, Berlin: Springer–Verlag, 1994. |
[17] | G. A. Chechkin, A. L. Piatnitski, A. S. Shamaev, Homogenization: Methods and Applications, American Mathematical Society, 2007. |
[18] |
A. G. Belyaev, A. L. Piatnitski, G. A. Chechkin, Averaging in a perforated domain with an oscillating third boundary condition, Sbornik Math., 192 (2001), 933–949. https://doi.org/10.1070/SM2001v192n07ABEH000576 doi: 10.1070/SM2001v192n07ABEH000576
![]() |
[19] |
R. Temam, Infinite-dimensional dynamical systems in mechanics and physics, Springer Science & Business Media, 68 (2012). https://doi.org/10.1007/978-1-4684-0313-8 doi: 10.1007/978-1-4684-0313-8
![]() |
[20] | A. V. Babin, M. I. Vishik, Attractors of Evolution Equations, Amsterdam: North–Holland, 1992. |
[21] | V. V. Chepyzhov, M. I. Vishik, Attractors for Equations of Mathematical Physics, American Mathematical Society, 49 (2002). |
[22] |
J. K. Hale, S. M. V. Lunel, Averaging in infinite dimensions, J. Integr. Equations Appl., 2 (1990), 463–494. https://doi.org/10.1216/jiea/1181075583 doi: 10.1216/jiea/1181075583
![]() |
[23] |
A. A. Ilyin, Averaging principle for dissipative dynamical systems with rapidly oscillating right-hand sides, Sbornik Math., 187 (1996), 635–677. https://doi.org/10.1070/SM1996v187n05ABEH000126 doi: 10.1070/SM1996v187n05ABEH000126
![]() |
[24] |
M. Efendiev, S. Zelik, Attractors of the reaction–diffusion systems with rapidly oscillating coefficients and their homogenization, Ann. I. H. Poincaré Nonlinear Anal., 19 (2002), 961–989. https://doi.org/10.1016/S0294-1449(02)00115-4 doi: 10.1016/S0294-1449(02)00115-4
![]() |
[25] |
K. A. Bekmaganbetov, G. A. Chechkin, V. V. Chepyzhov, Homogenization of random attractors for reaction–diffusion systems, C. R. Mec., 344 (2016), 753–758. https://doi.org/10.1016/j.crme.2016.10.015 doi: 10.1016/j.crme.2016.10.015
![]() |
[26] |
K. A. Bekmaganbetov, G. A. Chechkin, V. V. Chepyzhov, A. Y. Goritsky, Homogenization of trajectory attractors of 3D Navier-Stokes system with randomly oscillating force, Dyn. Syst., 37 (2017), 2375–2393. https://doi.org/10.3934/dcds.2017103 doi: 10.3934/dcds.2017103
![]() |
[27] |
G. A. Chechkin, V. V. Chepyzhov, L. S. Pankratov, Homogenization of trajectory attractors of Ginzburg–Landau equations with randomly oscillating terms, Discrete Contin. Dyn. Syst. Ser. B, 23 (2018), 1133–1154. https://doi.org/10.3934/dcdsb.2018145 doi: 10.3934/dcdsb.2018145
![]() |
[28] |
K. A. Bekmaganbetov, G. A. Chechkin, V. V. Chepyzhov, Weak convergence of attractors of reaction–diffusion systems with randomly oscillating coefficients, Appl. Anal., 98 (2019), 256–271. https://doi.org/10.1080/00036811.2017.1400538 doi: 10.1080/00036811.2017.1400538
![]() |
[29] | V. V. Chepizhov, M. I. Vishik, Trajectory attractors for reaction–diffusion systems, Topol. Methods Nonlinear Anal., 7 (1996), 49–76. |
[30] | F. Boyer, P. Fabrie, Mathematical Tools for the Study of the Incompressible Navier–Stokes Equations and Related Models, New York: Springer, 2013. |
[31] | O. A. Ladyzhenskaya, The Boundary Value Problems of Mathematical Physics, New York: Springer-Verlag, 1985. |
[32] | V. G. Maz'ya, Classes of spaces, measures, and capacities in the theory of spaces of differentiable functions, in Modern problems of Mathematics, Fundamental Investigations (Itogi Nauki i Techniki VINITI AN SSSR), Moskow: Nauka, 26 (1987). |
[33] | P. D. Lax, A. Milgram, Parabolic equations, in Contributions to the Theory of Partial Differential Equations, Princeton: Princeton University Press, 33 (1954), 167–190. |
[34] | V. G. Maz'ya, Sobolev Spaces with Applications to Elliptic Partial Differential Equations, Springer-Verlag, Berlin, 2011. |
[35] |
G. A. Chechkin, Y. O. Koroleva, L. E. Persson, On the precise asymptotics of the constant in Friedrich's inequality for functions vanishing on the part of the boundary with microinhomogeneous structure, J. Inequal. Appl., 2007 (2007), 13. https://doi.org/10.1155/2007/34138 doi: 10.1155/2007/34138
![]() |
[36] |
G. A. Chechkin, Y. O. Koroleva, A. Meidell, L. E. Persson, On the Friedrichs inequality in a domain perforated aperiodically along the boundary. Homogenization procedure. Asymptotics for parabolic problems, Russ. J. Math. Phys., 16 (2009), 1–16. https://doi.org/10.1134/S1061920809010014 doi: 10.1134/S1061920809010014
![]() |
[37] | J. L. Lions, Quelques Methodes de Resolution des Problemes aux Limites Nonlineaires, Paris: Dunod, Gauthier-Villars, 1969. |
[38] |
G. A. Chechkin, A. L. Piatnitski, Homogenization of boundary-value problem in a locally periodic perforated domain, Appl. Anal., 71 (1998), 215–235. https://doi.org/10.1080/000368 19908840714 doi: 10.1080/00036819908840714
![]() |