Startseite Boundary layer analysis of nonlinear reaction-diffusion equations in a smooth domain
Artikel Open Access

Boundary layer analysis of nonlinear reaction-diffusion equations in a smooth domain

  • Chang-Yeol Jung EMAIL logo , Eunhee Park und Roger Temam
Veröffentlicht/Copyright: 14. Januar 2016

Abstract

In this article, we consider a singularly perturbed nonlinear reaction-diffusion equation whose solutions display thin boundary layers near the boundary of the domain. We fully analyse the singular behaviours of the solutions at any given order with respect to the small parameter ε, with suitable asymptotic expansions consisting of the outer solutions and of the boundary layer correctors. The systematic treatment of the nonlinear reaction terms at any given order is novel along the singular perturbation analysis. We believe that the analysis can be suitably extended to other nonlinear problems.

1 Introduction

Nonlinear reaction-diffusion equations arise in many areas in systems consisting of interacting components. The equations describe, e.g., chemical reactions, pattern-formation, population dynamics, predator-prey equations, and competition dynamics in biological systems (see, e.g., [7, 11, 12, 31, 32, 33, 39]). One can consider a typical form of systems of reaction-diffusion equations in the form

(1.1)𝐮t=DΔ𝐮+𝐠(𝐮),

where 𝐠=𝐠(𝐮) describes a change or a local reaction of the state 𝐮 and D represents a diffusion coefficient matrix. It is also possible that the reaction 𝐠 may depend on the spatial domain variable 𝐱 and of a derivative of 𝐮, i.e., 𝐠=𝐠(𝐱,𝐮,𝐮).

In real applications like a fast reaction system, the magnitude of some coefficients in the diffusion matrix D is relatively small and hence the system can be singularly perturbed.

In this article, for the singular perturbation and boundary layer analysis aimed here, we consider the steady state system of (1.1) and study the following scalar nonlinear singularly perturbed problem which can serve as a guide for more general systems:

(1.2){-εΔuε+g(uε)=fin Ω,uε=0at Ω.

Here, 0<ε1, Ω is a general smooth domain, f=f(x,y) and g=g(u) are given smooth functions with

(1.3)g(0)=0,g(u)λ>0for all u.

For example, g(u)=u3+λu.

For small ε>0, the solutions to (1.2) display thin sharp transition layers called boundary layers which are formed due to the discrepancies between the limit solutions when ε=0 (see (2.3) below) and the boundary conditions in (1.2). The discrepancies are inevitable because the limit problem (see (2.2) below) loses high order derivatives and hence in general its solutions cannot meet the boundary conditions. Then, the small diffusion term -εΔuε smoothes out the discrepancies, which leads to sharp transition boundary layers.

Another motivation of studying boundary layers is the vanishing viscosity problem in fluid dynamics, see, e.g., [2, 5, 6, 24, 25, 8, 13, 28, 30, 23, 27, 35, 36, 37]. The typical question is on the behaviour of the Navier–Stokes flows at small viscosity, i.e., the limit behaviour or convergence to Euler flows as the viscosity tends to zero. The boundary layers play a crucial role for connecting the Navier–Stokes and Euler flows and they also do so for the singular perturbation analysis in the nonlinear reaction-diffusion equations considered here.

An additional motivation comes from the computational aspects in numerical simulations. Due to the thin boundary layers, the computational meshes are classically refined near the boundary Ω and this causes high cost in the simulations. Rather than refining meshes we propose to enrich with suitable boundary layer correctors the Galerkin or finite element space (or finite volume space). Then, we are able to use a coarse mesh and this reduces substantially the computational cost. See, e.g., [18, 17, 20, 21, 22, 38, 41] for the method of spaces enriched with boundary layer correctors. For singular perturbations analysis, see [15, 26, 42, 44] and also the recent review article [14]. See other perspectives in singular perturbations and boundary layers in [9, 16, 3, 19, 4, 10, 29, 43].

In what follows, we discuss the problems posed on a channel domain in Section 2 which is relatively easier to handle thanks to the simple geometry of the boundary. In Section 3, we cast the nonlinear reaction-diffusion equations in a general domain. We need to take into account the geometrical properties, like curvature, using the boundary fitted coordinates. Throughout this paper, we systematically handle the nonlinear term g along the singular perturbation analysis at any orders. This nonlinear treatment can apply to other nonlinear problems.

For the analysis below, we shall consider the Sobolev spaces Hs(Ω) and we define the weighted energy norm,

uε=(εuL2(Ω)2+uL2(Ω)2)12.

An exponentially small term, denoted e.s.t., is a function whose norm in all Sobolev spaces Hs(Ω) is exponentially small with, for each s, a bound of the form c1e-c2/εγ, c1,c2,γ>0, with ci, γ depending possibly on s.

2 Channel domains

For general domains, which will be studied in Section 3, we consider the domains with smooth boundaries. Since the boundary layer correctors act locally in the inward direction normal to the boundaries, transforming the Cartesian coordinate into the so-called boundary fitted one, the boundary layers can be described in channel domains, which are relatively easy to analyse. We thus consider first the simpler case of channel domains, which possess boundary layers only on one side at a flat boundary.

Let us consider the problem in a channel domain as follows:

(2.1){-εΔuε+g(uε)=fin Ω=(0,L1)×(0,L2),uε=0at x=0,L1,uε(x,y)=uε(x,y+L2)in Ω=(0,L1)×,

where f=f(x,y) is smooth and L2- periodic in y. Then, the limit problem reads

(2.2)g(u0)=fin Ω.

Since g is invertible, we write

(2.3)u0=g-1(f).

To give an idea on how to construct the boundary layers, for now we assume

(2.4)f=0at x=L1,

which, as we will see, reduces the boundary layer at x=L1, so that only the boundary layer at x=0 persists.

Thanks to (1.3) and (2.3), 0λ(u0)2(g(u0)-g(0))u0=fu0, and hence

(2.5)u0=0at x=L1andu0(x,y)=u0(x,y+L2).

2.1 Boundary layer analysis at order ε0

We now construct a zeroth order corrector to account for the discrepancy between uε and u0 at x=0. Formally, substituting uεu0+θ0 in (2.1) and subtracting (2.2) from (2.1), we find that

-εΔ(u0+θ0)+g(u0+θ0)-g(u0)=0.

Using the stretched variable x¯=x/ε and dropping non-stiff small terms, we find the zeroth order corrector equation for θ0:

-εθxx0+g(u0+θ0)-g(u0)=0.

However, in general uε-u0 does not satisfy the boundary condition in (2.1), and hence at the boundary x=0, so we propose a boundary layer corrector θ0 satisfying

(2.6){-εθxx0+g(u0+θ0)-g(u0)=0in Ω,θ0=-u0(0,y)at x=0,θ0=0at x=L1.

Although θ0 is not known explicitly, unlike in many linear problems, we can derive pointwise estimates for θ0.

Lemma 2.1.

The corrector θ0 satisfies

(2.7)|θ0(x,y)||u0(0,y)|exp(-λεx).

Proof.

Setting θ¯θ0=|u0(0,y)|exp(-λx/ε), writing θ~θ0=θ0-θ¯θ0 and then substituting in (2.6), we obtain -εθ~θxx0+g(u0+θ0)-g(u0)-λθ¯θ0=0. Since g(η)-λ0 for all η and thanks to the mean value theorem, we find, for some η1 with |η1-u0|<|θ0|, that

(2.8)-εθ~θxx0+g(η1)θ~θ0=(-g(η1)+λ)θ¯θ00.

Multiplying (2.8) by θ~θ+0=max{θ~θ0,0}, integrating over (0,L1) and noting that θ~θ+0=0 at x=0,L1, we obtain

ε0L1((θ~θ+0)x)2+λ0L1(θ~θ+0)20.

This implies θ~θ+0=0 and thus θ0-θ¯θ0=θ~θ00. On the other hand, considering this time θ~θ0=-θ0-θ¯θ0 we find that -εθ~θxx0+g(u0)-g(u0+θ0)-λθ¯θ0=0. We then similarly obtain (2.8) for this θ~θ0, and hence we deduce the same conclusion, i.e., -θ0-θ¯θ0=θ~θ00. This proves the lemma. ∎

We can also deduce some norm estimates.

Lemma 2.2.

There exists a constant c>0, independent of ε, such that

(2.9)θ0(,y)Hx1(0,L1)cε-14,θ0(,y)Lx2(0,L1)cε14.

Proof.

The second estimate of (2.9) directly follows from (2.7). To obtain the first estimate, we introduce θ¯θ=-u0(0,y)e-x/εδ(x), where δ(x) is a smooth cut-off function with δ(x)=1 for x[0,L1/4] and δ(x)=0 for x[3L1/4,). We observe that for a.e. y,

ε0L1|(θ0-θ¯θ)x|2dx=-ε0L1(θ0-θ¯θ)xx(θ0-θ¯θ)dx=0L1((-g(u0+θ0)+g(u0))+εθ¯θxx)(θ0-θ¯θ)dxcε12.

where the last inequality follows from the mean value theorem and the L2-estimate of θ¯θxx, θ¯θ, and θ0. This implies the lemma. ∎

Theorem 2.3.

Assume that (2.4) holds. Then, there exists a constant c>0 such that

(2.10)uε-u0-θ0εcε.

Proof.

Let w=uε-u0-θ0, then, thanks to (2.5), w=0 on Ω. Subtracting (2.2) and (3.14) from (2.1) we find that

(2.11){-εΔw+g(uε)-g(u0+θ0)=εΔu0+εθyy0in Ω,w=0on Ω.

Multiplying by w and integrating over Ω we find that

εΩ|w|2dxdy+Ω(g(uε)-g(u0+θ0))wdxdyλ2Ω|w|2dxdy+cε2.

Here, the L2-norm of θyy0 is derived in Lemma 2.8 below. Thanks to the mean value theorem again and by observing that (g(uε)-g(u0+θ0))wλ|w|2, the theorem is proved. ∎

We can also obtain the lower bound of |θ0(x,y)|.

Lemma 2.4.

The corrector θ0 satisfies

(2.12)|θ0(x,y)||u0(0,y)|exp(-λ0εx)+e.s.t.,

where λ0=max|η-u0||θ0|g(η).

Proof.

From Lemma 2.1 and (2.3), we note that u0,θ0 are bounded and hence λ0>0 is too. We write θ~θ0=θ0-θ¯θ0, where θ¯θ0=|u0(0,y)|(exp(-λ0x/ε)-L1-1xexp(-λ0L1/ε))0.

Fixing y, we first prove (2.12) for the case u0(0,y)0. We note that θ~θ0=0 at x=0,L1. Following the proof of Lemma 2.1, we similarly find that for some η1 with |η1-u0|<|θ0|,

-εθ~θxx0+g(η1)θ~θ0(-g(η1)+λ0)θ¯θ00.

Multiplying by -θ~θ-0=-max{-θ~θ0,0} and integrating over (0,L1), we obtain

ε0L1((θ~θ-0)x)2+λ0L1(θ~θ-0)20.

This implies θ~θ-0=0 and hence |θ0|θ0θ¯θ0, which proves (2.12) for the case u0(0,y)0. For the case u0(0,y)>0, we write θ~θ0=-θ0-θ¯θ0. Then, we similarly deduce that θ~θ-0=0 and hence |θ0|-θ0θ¯θ0. This proves the lemma. ∎

Remark 2.5.

Thanks to the estimate for θ0 in L2, established in Lemma 2.2 with (2.4), Theorem 2.3 implies that

(2.13)uε-u0L2(Ω)cε14.

Furthermore, for u0(0,y)0 at some y(0,L2) the L2-norm in (2.13) has a lower bound, i.e., for some c0>0,

uε-u0L2(Ω)c0ε14.

Indeed, from Lemma 2.4 and Theorem 2.3, we find that

uε-u0L2θ0L2-uε-u0-θ0L2c2ε14-cεc0ε14.

2.2 Boundary layer analysis at arbitrary order εn, n0

Outer expansion. We now consider the higher order outer expansions uεj=0εjuj. Substituting in (2.1) and using (2.2), we formally write

(2.14)-εΔ(j=0εjuj)+g(j=0εjuj)=f.

Dropping 𝒪(εn+1) terms, we have

-εΔ(j=0n-1εjuj)+g(j=0nεjuj)f.

We identify at the order 𝒪(εj), j=0,1,,n, and find

(2.15){g(u0)=f,-εjΔuj-1+g(k=0jεkuk)-g(k=0j-1εkuk)=0,j1.

We then obtain, e.g.,

u0=g-1(f),
u1=ε-1g-1(g(u0)+εΔu0)-ε-1u0.

More generally, we recursively obtain

(2.16)uj=ε-jg-1(g(k=0j-1εkuk)+εjΔuj-1)-ε-jk=0j-1εkukfor j1.

To construct the higher order correctors, we assume, for simplicity, that f is infinitely flat at x=L1, i.e.,

(2.17)Dαf=|α|fxα1yα2=0at x=L1, for all α0,

using the multi-index notation

(2.18)α=(α1,α2)with |α|=α1+α2.

This implies that the uj, j0, are infinitely flat at x=L1, that is

Dαuj=0at x=L1, for all j,α0.

Thus, we only have boundary layers at x=0 corresponding to uj. Correctors. We now proceed with the determination of the correctors. Substituting uεj=0εj(uj+θj) in (2.1), we have formally

(2.19)-εΔ(j=0εj(uj+θj))+g(j=0εj(uj+θj))=f.

We subtract (2.14) from (2.19) to obtain

(2.20)-εΔ(j=0εjθj)+g(j=0εj(uj+θj))-g(j=0εjuj)=0.

We first need to handle the nonlinear term to identify the quantities of order εj and this is discussed below.

2.3 Treatment of the nonlinear term g(u)

In this section, we formally write the nonlinear term g(j=0εj(uj+θj))-g(j=0εjuj) at each order εj. Thanks to the Taylor expansion of g about u0, we have

g(j=0εjuj)=g(u0+j=1εjuj)=k=0g(k)(u0)k!(j=1εjuj)k.

Here, we formally consider j=1εjuj=𝒪(ε). Similarly, expanding at u0+θ0, we write

g(j=0εj(uj+θj))=k=0gk(u0+θ0)k!(j=1εj(uj+θj))k.

We first observe that

(j=1εjuj)k=|α|=k(ka)((ε1u1)α1(εlul)αl)
=|α|=k(ka)((u1)α1(ul)αl)ε(α1+2α2++lαl+)
=|α|=k(ka)uαε(α1+2α2++lαl+),

where

uα=(u1)α1(ul)αl,

using the multi-index notation

(2.21)α=(α1,,αl,)with |α|=α1++αl+,

and

(kα)=k!α1!αl!.

We similarly find that

(j=1εj(uj+θj))k=|α|=k(kα)(u+θ)αε(α1+2α2++lαl+),

where

(u+θ)α=(u1+θ1)α1(ul+θl)αl.

Hence, we note that

g(j=0εj(uj+θj))-g(j=0εjuj)=k=01k![g(k)(u0+θ0)(j=1εj(uj+θj))k-g(k)(u0)(j=1εjuj)k]
(2.22)=k=0|α|=k(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]εα1+2α2++lαl+,

using the notation (2.21). To arrange the terms at each order εj, we set α1+2α2++lαl+=j. Since the multi-index α satisfies |α|=k, we easily note that k=|α|α1+2α2++lαl+=j. If one of the αl with lj+1 is greater than or equal to 1, then α1+2α2++lαl+j+1, and hence αj+1=αj+2==0. Thus, we may write the multi-index notations as

(2.23){α=(α1,,αj),|α|=α1++αj,(u+θ)α=(u1+θ1)α1(uj+θj)αj,uα=(u1)α1(uj)αj.

Hence, using the multi-index notations in (2.23), we formally write

g(j=0εj(uj+θj))-g(j=0εjuj)
(2.24)=j=0{k=0j|α|=k,α1+2α2++jαj=j(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj.

For the analysis below, we estimate the truncation error corresponding to the expansion (2.24).

Lemma 2.6.

There exists a constant C>0, independent of ε, such that

|g(j=0nεj(uj+θj))-g(j=0nεjuj)-Gn|Cεn+1,

where

Gn=j=0n{k=0j|α|=k,α1+2α2++jαj=j(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj,

and the multi-index notations are given in (2.23).

Proof.

We first note that the Gn given above can be written as

Gn=k=0n1k!j=0n{|α|=k,α1+2α2++jαj=j(kα)[g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj.

Thanks to the multinomial theorem, we observe that

(2.25)Hn,k:=g(k)(u0+θ0)(j=1nεj(uj+θj))k-g(k)(u0)(j=1nεjuj)k
(2.26)=j=0n{|α|=k,α1+2α2++jαj=j(kα)[g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj+Jn,k,

where

(2.27)|Jn,k|Cεn+1.

On the other hand, we find from Taylor’s theorem that

g(j=0nεj(uj+θj))-g(j=0nεjuj)=k=0n1k!Hn,k+R0=Gn+k=0n1k!Jn,k+R0,

where

|R0||g(n+1)(n+1)!(ξ1)(j=1nεj(uj+θj))n+1|+|g(n+1)(n+1)!(ξ2)(j=1nεjuj)n+1|.

Here, ξ1 is between (u0+θ0) and j=0nεj(uj+θj) and ξ2 is between u0 and j=0nεjuj. The lemma follows by observing that |R0|Cεn+1. ∎

We now define the boundary layer correctors θj at order 𝒪(εj). From (2.20) and (2.24), using the stretched variable x¯=x/ε at each order 𝒪(εj), j=0,1,, we identify

(2.28){-εθxx0+g(u0+θ0)-g(u0)=0,-εj+1θxxj+{k=1j|α|=k,α1+2α2++jαj=j(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj=εjθyyj-1,j1.

Dividing by εj, rearranging terms in the latter equation at each order 𝒪(εj), and using the fact that

|α|=k=1and(α1+2α2++jαj=j)α1==αj-1=0,αj=1,

we rewrite (2.28) as

-εθxx0+g(u0+θ0)-g(u0)=0,
-εθxx1+g(u0+θ0)θ1=-(g(u0+θ0)-g(u0))u1+θyy0,

and for j2,

-εθxxj+g(u0+θ0)θj=-(g(u0+θ0)-g(u0))uj
(2.29)-k=2j|α|=k,α1+2α2++jαj=j(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]+θyyj-1.

We supplement the boundary condition on θj, for each j=0,1,, by

(2.30){θj=-uj(0,y)at x=0,θj=0at x=L1.

Remark 2.7.

We note that the corrector equations for θj, j1, are all linear and this allows us to directly apply the maximum principle. Differentiating the equations in y, the maximum principle also holds for mθjym(x,y) for m1, j0.

Lemma 2.8.

The correctors θj, j0, satisfy

(2.31)|mθjym(x,y)|Cexp(-34λεx),m0.

Proof.

We use the maximum principle to prove the lemma. Let be the linear operator given by

u:=-εuxx+g(u0+θ0)u.

For j=0, we have

-εθxx0+g(u0+θ0)-g(u0)=0.

We introduce a barrier function Ψ=C01exp(-34λεx), where C01 will be chosen later. We use the mathematical induction on m starting from the case j=0. By (2.7), we already have |θ0(x,y)|cexp(-34λεx). We then see that

|θy0|=|-εθyxx0+g(u0+θ0)θy0|=|-g(u0+θ0)uy0+g(u0)uy0||g′′(η)||θ0||uy0|,

by the mean value theorem, for some η between (u0+θ0) and θ0. We also find

Ψ=(g(u0+θ0)-916λ)C01exp(-34λεx).

Since g(u0+θ0)λ and |g′′(η)||θ0||uy0| is bounded on Ω¯, we can find a positive constant C01 such that

|θy0|C01exp(-34λεx)in Ω.

By the boundary conditions of θ0, we obtain

|θy0||u0(0,y)|C01exp(-34λεx)on Ω,

where C01=max(|u0(0,y)|,C01). The maximum principle implies that

|θy0(x,y)|Ψin Ω¯.

We suppose by induction that for k(m-1), m1, there exists a positive constant C0k satisfying

|kθ0yk(x,y)|C0kexp(-34λεx)in Ω¯.

We then find that

(mymθ0)=m-1ym-1(g(u0)uy0-g(u0+θ0)uy0)
=k=0m-1[h0k(g(u0),,g(m)(u0),g(u0+θ0),,g(m)(u0+θ0),uy0,,mu0ym,
(2.32)g′′(η1),,g(m+1)(ηm))Pk(θ0)],

where P0(θ0)=θ0 and Pk(θ0)=α1++(m-1)αm-1=ki=1m-1(yiθ0)αi for k1 with some multivariate polynomials h0k and ηk between (u0+θ0) and u0. Since g and u0 are smooth, there exists C0m such that

(mymθ0)C0mexp(-34λεx)in Ω.

We infer from the boundary conditions for θ0 that

|mymθ0||mymu0(0,y)|on Ω,

and using the maximum principle we obtain that

(2.33)|mymθ0|C0mexp(-34λεx)in Ω¯,

where C0m=max(|mymu0(0,y)|,C0m). Similarly, for the case when j=1, we see that for m0,

(mymθ1)=mym(g(u0)u1-g(u0+θ0)u1+θyy0)
=k=0m+2[h1k(g(u0),,g(m+1)(u0),g(u0+θ0),,g(m+1)(u0+θ0),uy0,,mu0ym,
(2.34)u1,,mu1ym,g′′(η1),,gm+2(ηm+1))Pk(θ0)],

where P0(θ0)=θ0 and Pk(θ0)=α1++(m+2)αm+2=ki=1m+2(yiθ0)αi for k1 with for some multivariate polynomial h1k and ηk between (u0+θ0) and u0. Since g and uj are smooth and h1k are polynomials, we can find a positive constant C1m, by the result for θ0, such that

(2.35)|(mymθ1)|C1mexp(-34λεx)in Ω.

By the boundary condition on θ1, we also have

(2.36)|mymθ1||mymu1(0,y)|on Ω.

We find from (2.35) and (2.36) that for m0,

|mymθ1|C1mexp(-34λεx)in Ω¯,

by the maximum principle where C1m=max(C1m,|mymu1(0,y)|). We now suppose by induction that for k(j-1), j1, there exists a positive constant Cjm such that

(2.37)|mθkym(x,y)|Cjmexp(-34λεx),m=0,1,.

To prove (2.31) at order k=j, differentiating (2.29) in y we note that the first and third terms of the right-hand side of (2.29) are similarly estimated as for the case θ1 by (2.37). We thus estimate the second term there. Observing that for k2,

(2.38)(|α|=k,α1+2α2++jαj=j)(|α|=k,α1+2α2++(j-1)α(j-1)=j),

it suffices to show that for any m0,

(2.39)|mym(k=2j|α|=k,α1+2α2++(j-1)α(j-1)=j[g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα])|Cjmexp(-34λεx).

To prove this, we note that

(2.40)g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα=(g(k)(u0+θ0)-g(k)(u0))(u+θ)α+g(k)(u0)((u+θ)α-uα),

and using the factorization an-bn=(a-b)i=1nan-ibi-1,

(u+θ)α-uα=l=1j-1[θl(i=1αl(ul)αl-i(θl)i-1)n=1l-1(un)αnn=l+1j-1(un+θn)αn],

where α=(α1,,αj-1). Differentiating (2.40) in y, thanks to the mean value theorem, the left-hand side of (2.39) can be written as the sum of the products of θk and their derivatives in y for k(j-1). We then conclude, by assumption (2.37), that (2.39) holds true. ∎

We now estimate, for each n=0,1,, the norm of wn, where wn=uε-j=0nεj(uj+θj). Summing (2.15) for j=0 to j=n, we find

(2.41)-εΔ(j=0n-1εjuj)+g(j=0nεjuj)=f.

Summing (2.28) for j=0 to j=n, we find

-εj=0nεjθxxj+j=0n{k=0j|α|=k,α1+2α2++jαj=j(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εj=j=0nεjθj-1yy.

Thanks to Lemma 2.6, this can be written in the form

(2.42)-εΔ(j=0nεjθj)+g(j=0nεj(uj+θj))-g(j=0nεjuj)+R1=-εn+1θyyn

with

(2.43)|R1|cεn+1.

Adding the two above equations (2.41) and (2.42), we find

(2.44)-εΔ(j=0nεj(uj+θj))+g(j=0nεj(uj+θj))=f-εn+1Δun-εn+1θyyn-R1.

Subtracting (2.44) and from the first equation in (2.1), we find

(2.45)-εΔwn+g(uε)-g(j=0nεj(uj+θj))=εn+1Δun+εn+1θyyn+R1.

We multiply (2.45) by wn and since g(u)-g(v)=g(ξ)(u-v) and g(ξ)λ>0, we obtain, by a priori estimate, that

εwnH1+wnL2cεn+1.

We hence proved the following convergence theorem.

Theorem 2.9.

Assume that (2.17) holds. Let uε be the solution of (2.1) and uj and θj be given as in (2.16) and (2.29)–(2.30), respectively. Then, there exists a positive constant c>0, independent of ε, such that

uε-j=0nεj(uj+θj)εcεn+1.

2.4 Without the assumption (2.17)

If we consider a general smooth function f, i.e., if we remove the assumptions (2.17), we also expect similar boundary layers at x=L1. Let us denote similar boundary layers θj at x=0 by θlj, j0, given in (2.29). Similarly, we define the boundary layers at x=L1, denoted by θrj, which satisfy equations (2.29) but with different boundary conditions, i.e.,

{θrj=0at x=0,θrj=-uj(L1,y)at x=L1.

Then, we define the boundary layers θj, j0,

(2.46)θj=θlj+θrj.

Since the corrector equations for θlj,θrj, j1, are linear, we infer from equation (2.42) that

(2.47)-εΔ(j=0nεjθj)+g(j=0nεj(uj+θj))-g(j=0nεjuj)+R~1=-εn+1θyyn.

Here,

(2.48)R~1=R1+R2,

where R1 is given in (2.42)–(2.43) and

R2=g(u0+θl0)+g(u0+θr0)-g(u0+θ0)-g(u0).

We now note that R2 is exponentially small. Indeed, we find that, for all α0,

DαR2L2(Ω)DαR2L2((0,L1/2)×(0,L2))+DαR2L2((L1/2,L1)×(0,L2))
Dα(g(u0+θl0)+g(u0+e.s.t)-g(u0+θl0+e.s.t)-g(u0))L2((0,L1/2)×(0,L2))
+Dα(g(u0+e.s.t.)+g(u0+θr0)-g(u0+e.s.t+θr0)-g(u0))L2((L1/2,L1)×(0,L2)),

which is an exponentially small term. Then, the convergence analysis similarly follows as in the above, from which we infer the following theorem.

Theorem 2.10.

Let f be a general smooth function periodic in y with period L2, let uε be the solution of (2.1), and let uj and θj be given as in (2.16) and (2.46), respectively. Then, there exists a positive constant c>0, independent of ε, such that

uε-j=0nεj(uj+θj)εcεn+1.

3 General domains

We now return to the case of a general smooth domain, where equation (1.2) is posed:

(3.1){-εΔuε+g(uε)=fin Ω,uε=0at Ω.

Here, Ω is a general smooth domain, f=f(x,y) and g(u) are given smooth functions with

(3.2)g(0)=0,g(u)λ>0for all u.

The outer solutions uj, j0, are the same as in (2.16). That is, in Ω, we have

(3.3){u0=g-1(f),u1=ε-1g-1(g(u0)+εΔu0)-ε-1u0,uj=ε-jg-1(g(k=0j-1εkuk)+εjΔuj-1)-ε-jk=0j-1εkukfor j1.

However, the boundary layers appear in the direction normal to the curved boundary. Thus, the boundary fitted coordinates, i.e., the normal and tangential components along the boundary, are necessary to devise the boundary layer correctors. Here, we consider smooth boundaries Ω which are parametrised by an arclength η and we assume that (X(η),Y(η))Ω is a regular curve, i.e., the tangent vector T=(X,Y)0 for the arclength 0η<L0 is measured counterclockwise, where L0 is the length of the boundary Ω. Hence, we may assume that it has unit speed, i.e., (X)2+(Y)2=1 (see [34]).

We then define the boundary fitted coordinates:

(3.4)ΩBL={(x,y):x=X(η)-ξY(η),y=Y(η)+ξX(η), 0η<L0, 0ξ<ξ0},

where ξ0 is the minimum radius of curvature of Ω, i.e., ξ0=1/maxκ(η). Here we note that (-Y(η),X(η)) is the inward unit normal vector to the boundary Ω.

3.1 Boundary fitted coordinates

We introduce the local orthogonal coordinate basis 𝐠k, i=1,2, on the subdomain ΩBL by setting

𝐠1=xξ𝐞1+yξ𝐞2=-Y𝐞1+X𝐞2,
𝐠2=xη𝐞1+yη𝐞2=(X-ξY′′)𝐞1+(Y+ξX′′)𝐞2,

where 𝐞1,𝐞2 is the standard basis in 2 (see, e.g., [1, 40]). We can easily compute that

(3.5)𝐠1𝐠1=(X)2+(Y)2=1,
(3.6)𝐠1𝐠2=ξXX′′+ξYY′′=0,
𝐠2𝐠2=1+2ξ(YX′′-XY′′)+ξ2((X′′)2+(Y′′)2).

Here, we note that differentiating (3.5) implies (3.6). The curvature κ(η) of Ω is derived

κ(η)=XY′′-YX′′((X)2+(Y)2)32=XY′′-YX′′

and, by the Frenet–Serret relation, T=(X′′,Y′′)=κ(η)N(η) where N(η) is the unit normal outward vector. Hence, we obtain that

𝐠2𝐠2=(1-κ(η)ξ)2.

For the variables (ξ,η) in

Ωξ,η={(ξ,η):0η<L0, 0ξ<ξ0},

since 0ξ<ξ0=1/maxκ(η), we find that 1-κ(η)ξ>0, and thus we have

h1:=𝐠1𝐠1=1,h2:=𝐠2𝐠2=1-κ(η)ξ.

The gradient and Laplacian operators are then defined as follows:

=i=12𝐠ihi2ζi,Δ=1h1h2i=12ζi(h1h2hi2ζi),

where (ζ1,ζ2)=(ξ,η). Hence, we find that

=(-Yξ+(X-ξY′′)σ2(ξ,η)η,Xξ+(Y+ξX′′)σ2(ξ,η)η),
(3.7)andΔ=2ξ2-κ(η)σ(ξ,η)ξ+σ2(ξ,η)2η2+ξκ(η)σ3(ξ,η)η,

where

(3.8)σ(ξ,η)=11-κ(η)ξ.

Then, the model equation in (3.1), can be written as

(3.9)-ε2uεξ2+εκ(η)σ(ξ,η)uεξ-εσ2(ξ,η)2uεη2-εξκ(η)σ3(ξ,η)uεη+g(uε)=f.

Remark 3.1.

On the unit circle domain, using polar coordinates, we find X(η)=cosη, Y(η)=sinη and σ(ξ)=(1-ξ)-1 (note κ(η)=1). The operators , Δ are simplified to

=(-cosηξ-sinη1-ξη,-sinηξ+cosη1-ξη),
andΔ=1(1-ξ)22η2-11-ξξ+2ξ2.

3.2 Boundary layer analysis at orders ε0 and ε12

Unlike the channel domain (2.1), we will need to introduce a corrector θ12 to settle the error from the curvature κ(η).

In general, to find appropriate asymptotic expansions for the boundary layers, we preform the following expansions near the boundary ξ=0. Subtracting (2.41) from (3.9) and observing that ε is the thickness of the boundary layers, as indicated in previous sections, which suggests to use the stretched variable ξ¯ξ=ξ/ε, we find that

(-2ξ¯ξ2+εκ(η)σ(ξ,η)ξ¯ξ-εσ2(ξ,η)2η2-εξκ(η)σ3(ξ,η)η)(uε-j=0nεjuj)
(3.10)+g(uε)-g(j=0nεjuj)=εn+1Δun.

To address the terms at all the orders of ε in the boundary layers, we have to resolve the effect of curvature κ(η). Here, κ(η)=0 for the channel domain. We first observe that σ(ξ,η)=(1-κ(η))-1=l=0(κ(η)ξ¯ξ)lεl2 and the powers of σ(ξ,η) can be similarly expressed (see also (3.38) below). We then take into account the mean value theorem for the first two terms in the second line of (3.10), i.e., g(uε)-g(j=0nεjuj)=g(u)(uε-j=0nεjuj) for some u. In this way, we can balance the difference between uε and the outer expansion at order n by using the inner expansion near Ω in the form

(3.11)uε-j=0nεjujj=0n(εθj+εj+12θj+12)near Ω.

By comparing the terms of the same order εj on Ω, we deduce from (3.11) the following boundary conditions for j0:

(3.12){θj=-ujon Ω,θj+12=0on Ω.

We now find two leading order correctors θ0 and θ12 satisfying

(3.13)-ε(θξξ0+εθξξ12)+εκ(η)(θξ0+εθξ12)+g(u0+θ0+εθ12)-g(u0)0.

By the Taylor expansion, dropping smaller terms and using the stretched variable ξ¯ξ=ξ/ε, we obtain θ¯θ0 at order 𝒪(ε0), which is a solution of

(3.14)-θ¯θξ¯ξξ¯ξ0+g(u0+θ¯θ0)-g(u0)=0,

and from (3.13) at order 𝒪(ε12), we again find θ¯θ12 such that

(3.15)-θ¯θξ¯ξξ¯ξ12+g(u0+θ¯θ0)θ¯θ12=-κ(η)θ¯θξ¯ξ0.

These equations are supplemented with the respective boundary conditions, i.e.,

(3.16){θ¯θ0=-u0|ξ=0=-u0(X(η),Y(η))at ξ=0,θ¯θ0=0at ξ=ξ0,  {θ¯θ12=0at ξ=0,θ¯θ12=0at ξ=ξ0.

Then, we extend θ¯θ0 and θ¯θ12 by zero for ξ>ξ0 and these extensions are still denoted by the same notations. However, these extensions are not smooth. Thus, for the analysis below, let us define

θ0=θ¯θ0δ(ξ),θ12=θ¯θ12δ(ξ),

where δ(ξ) is a smooth cut-off function given by

(3.17)δ(ξ)={1,ξ[0,ξ0/4],0,ξ[3ξ0/4,).

Remark 3.2.

We obtain θ0 by multiplying θ¯θ0 by the cut-off function. This allows us to use the same estimates for θ0 as θ¯θ0. We proceed similarly for θ12.

Lemma 3.3.

The following pointwise estimate holds for θ¯θ0:

(3.18)|θ¯θ0||u0|ξ=0exp(-λεξ).

Furthermore, the derivatives of θ¯θ0 satisfy pointwise, for l,m,n0, the following:

(3.19)|ξnl+mθ¯θ0ξlηm|cεn-l2exp(-12λεξ).

Proof.

We first set ψ=|u0|ξ=0exp(-λξ¯ξ). Writing θ~θ=θ¯θ0-ψ, we deduce from (3.14) that

-θ~θξ¯ξξ¯ξ+g(u0+θ¯θ0)-g(u0)-λψ=0,

which implies

(3.20)-θ~θξ¯ξξ¯ξ+g(η)θ~θ0.

Thanks to (3.16), multiplying (3.20) by θ~θ+(ξ¯ξ,)H01(0,) and integrating over (0,), we find that

0|(θ~θ+)ξ¯ξ|2dξ¯ξ+λ0|θ~θ+|2dξ¯ξ0.

This implies θ~θ+=0 for all ξ¯ξ0 and thus θ¯θ0ψ. The other inequality -θ¯θ0ψ similarly follows.

We now find the estimates for the derivatives as in (3.19). For l=0 and n=0, it immediately follows, from Lemma 2.8, that

|mθ¯θ0ηm|cexp(-34λεξ).

Multiplying by ξn, the estimate (3.19) follows for l=0 and n0.

Then, at higher orders l, we use the multi-index α=(l-2,m) with l2 and Dα=ξ¯ξl-2ηm. Applying the operator Dα to (3.14), we have

ξ¯ξlηmθ¯θ0=(Dαθ¯θ0)ξ¯ξξ¯ξ=Dα(g(u0+θ¯θ0)-g(u0)).

Thanks to the mean value theorem, we observe that

(3.21)Dα(g(u0+θ¯θ0)-g(u0))=finite sum of products of Dβθ¯θ0 with β(l-2,m),

and we can thus inductively prove (3.19) for any l0 as long as the case for l=1 is proved.

For l=1, we let μ>0, which will be determined. We infer from (3.14) that

(θ¯θξ¯ξ0eμξ¯ξ)ξ¯ξ=(g(u0+θ¯θ0)-g(u0))eμξ¯ξ+μθ¯θξ¯ξ0eμξ¯ξ=(g(u0+θ¯θ0)-g(u0))eμξ¯ξ+(μθ¯θ0eμξ¯ξ)ξ¯ξ-μ2θ¯θ0eμξ¯ξ.

Integrating over (ξ¯ξ,), we find

C-θ¯θξ¯ξ0eμξ¯ξ=I1(ξ¯ξ)-μθ¯θ0eμξ¯ξ,

where

I1(ξ¯ξ)=ξ¯ξ(g(u0+θ¯θ0)-g(u0)-μ2θ¯θ0)(t,η)eμtdt.

Then,

(3.22)θ¯θξ¯ξ0=Ce-μξ¯ξ-I1(ξ)e-μξ¯ξ+μθ¯θ0,

and integrating over (ξ¯ξ,), we find that

(3.23)-θ¯θ0=D+μ-1Ce-μξ¯ξ-ξ¯ξI1(s)e-μsds+μξ¯ξθ¯θ0(s,η)ds.

Let μ=pλ, where p, a constant independent of ε, is to be determined. For 0<p<1, since (3.18) yields |θ¯θ0(t,η)|cexp(-λt), we find that

|I1(ξ¯ξ)|cξ¯ξe-λ(1-p)tdtce-λ(1-p)ξ¯ξ,

and

|ξ¯ξI1(s)e-μsds|cξ¯ξe-λsdsce-λξ¯ξ.

Let us choose p=34 and thus μ=34λ. Applying the boundary conditions to (3.23), i.e., θ¯θ00 as ξ¯ξ and θ¯θ0=-u0|ξ=0 at ξ¯ξ=0, we find that D=0 and that

-u0(0,η)=θ¯θ0(0,η)=-μ-1C+0I1(s)e-μsds-μ0θ¯θ0(s,η)ds.

Thus,

C=μ[u0(0,η)+0I1(s)e-μsds-μ0θ¯θ0(s,η)ds].

To estimate ηmθ¯θξ¯ξ0, we apply ηm to (3.22) and find that

|ηmθ¯θξ¯ξ0|(|ηmC|+|ηmI1(ξ¯ξ)|)exp(-34λεξ)+c|ηmθ¯θ0|
(3.24)cexp(-34λεξ).

Here, from (3.21) with α=(0,m), i.e., l=2, and (3.19) with l=0, we used the fact that

|ηmC|+|ηmI1(ξ¯ξ)|c.

We similarly derive the pointwise estimate for θ¯θ12 and the result appears in the following lemma.

Lemma 3.4.

The corrector θ¯θ12 satisfies pointwise, for l,m,n0, the following:

(3.25)|ξnl+mθ¯θ12ξlηm|cεn-l2exp(-12λεξ).

Proof.

Let be the linear operator given by

u=-uξ¯ξξ¯ξ+g(u0+θ¯θ0)u.

We repeat the same argument as in the proof of Lemma 3.3. For l=0, the estimates (3.25) hold true by the maximum principle applied to (ηmθ¯θ12) with (3.15). For l=1, we infer from (3.15) that

(θ¯θξ¯ξ12eμξ¯ξ)ξ¯ξ=g(u0+θ¯θ0)θ¯θ12eμξ¯ξ+κ(η)θ¯θξ¯ξ0eμξ¯ξ+(μθ¯θ12eμξ¯ξ)ξ¯ξ-μ2θ¯θ12eμξ¯ξ.

We integrate over (ξ¯ξ,), to find that

C-θ¯θξ¯ξ12eμξ¯ξ=I2(ξ¯ξ)-μθ¯θ12eμξ¯ξ,

where

I2(ξ¯ξ)=ξ¯ξ(g(u0+θ¯θ0)θ¯θ12+κ(η)θ¯θξ¯ξ0-μ2θ¯θ12)(t,η)eμtdt.

Let μ=pλ, where p is a constant independent of ε to be chosen later. From estimates (3.19), we note that |θ¯θξ¯ξ0(t,η)|cexp(-λt/2). Using the same argument as in the proof of Lemma 3.3, we only need to show that for 0<p<1,

(3.26)|I2(ξ¯ξ)|cξ¯ξe-λ(1-p)tdtce-λ(1-p)ξ¯ξ.

Following then the same procedure as in the proof of Lemma 3.3 with the boundary conditions (3.16), we can obtain (3.24) for θ¯θ12.

For l2, differentiating (3.15) in ξ¯ξ and using estimates (3.19), the lemma is inductively proved. ∎

We now find the norm estimate in L2 and H1 for θ¯θ0 and θ¯θ12 in the next lemma.

Lemma 3.5.

Let l,m,n0. Then, there exists c>0 such that

(3.27)ξnl+mθ¯θ0ξlηmLξ2(Ω)+ξnl+mθ¯θ12ξlηmLξ2(Ω)cεn-l2+14.

Proof.

We infer from Lemma 3.3 and Lemma 3.4 that (3.27) holds. ∎

We now introduce the analogue of Theorem 2.3.

Theorem 3.6.

Assume f is a general smooth function and Ω is a general smooth domain. Then, there exists a positive constant c>0 such that

uε-u0-θ0εcε34.

Proof.

We set w=uε-u0-θ0. To avoid the singularity of σ(ξ,η)=(1-κ(η)ξ)-1, we use the smooth cut-off function δ~δ(ξ) such that

(3.28)δ~δ(ξ)={1,ξ[0,ξ0/2],0,ξ[3ξ0/4,).

Here, we recall that ξ0=1/maxκ(η). Noting that θ0=0 for ξξ0/2 and denoting by (,) the scalar product in the space L2(Ω), we can write

(εΔθ0-g(u0+θ0)+g(u0),w)=(εΔθ0-g(u0+θ0)+g(u0),δ~δ(ξ)w)
=(G0+E0,δ~δ(ξ)w)+(εΔθ¯θ0-g(u0+θ¯θ0)+g(u0),δ~δ(ξ)w)
(3.29)=((G0+E0)δ~δ(ξ),w)+(R0δ~δ(ξ),w),

thanks to (3.14). Here,

E0=εΔ(θ0-θ¯θ0),
G0=g(u0+θ¯θ0)-g(u0+θ0),
R0=-εκ(η)σ(ξ,η)θ¯θ0ξ+εσ2(ξ,η)2θ¯θ0η2+εξκ(η)σ3(ξ,η)θ¯θ0η.

Since θ0-θ¯θ0=θ¯θ0(1-δ(ξ)), it is very easy to prove that G0 and E0 are e.s.t. We also find that

R0δ~δ(ξ)L2(Ω)cε34.

Taking the inner product of (3.3) and (3.1), respectively, with w, we write

(3.30)(-εΔuε+g(uε),w)=(f,w),
(3.31)(g(u0),w)=(f,w).

From (3.29), (3.30) and (3.31), we find that

(-εΔw+g(uε)-g(u0+θ0),w)
=(-εΔuε+g(uε),w)-(g(u0),w)+(εΔu0,w)+(εΔθ0-g(u0+θ0)+g(u0),w)
=(εΔu0,w)+(G0δ~δ(ξ),w)+(R0δ~δ(ξ),w)
cε34wL2(Ω).

Thanks to (3.2), this completes the proof of Theorem 3.6. ∎

Theorem 3.7.

Assume that f is a general smooth function and Ω is a general smooth domain. Then, there exists a positive constant c>0 such that

uε-u0-θ0-ε12θ12εcε.

Proof.

We define w12=uε-u0-θ0-εθ12, then we find

(εΔ(θ0+εθ12)-g(u0+θ0+εθ12)+g(u0),w12)
=(εΔ(θ0+εθ12)-g(u0+θ0+εθ12)+g(u0),δ~δ(ξ)w12)
=(G12+E12,δ~δ(ξ)w12)+(εΔ(θ¯θ0+εθ¯θ12)-g(u0+θ¯θ0+εθ¯θ12)+g(u0),δ~δ(ξ)w12)
(3.32)=((G12+E12)δ~δ(ξ),w12)+(R12δ~δ(ξ),w12),

by (3.14) and by (3.15), where

(3.33){E12=εΔ(θ0+εθ12-θ¯θ0-εθ¯θ12),G12=g(u0+θ¯θ0+εθ¯θ12)-g(u0+θ0+εθ12),R12=-εκ(η)σ(ξ,η)ξ(θ¯θ0+εθ¯θ12)+εκ(η)ξθ¯θ0+εσ2(ξ,η)2η2(θ¯θ0+εθ¯θ12)+εξκ(η)σ3(ξ,η)η(θ¯θ0+εθ¯θ12)+g(u0+θ¯θ0)+g(u0+θ¯θ0)εθ¯θ12-g(u0+θ¯θ0+εθ¯θ12).

We note that thanks to the Taylor expansion, we have

g(u0+θ¯θ0+εθ¯θ12)-g(u0+θ¯θ0)=g(u0+θ¯θ0)εθ¯θ12+T12,

where

(3.34)T12L2(Ω)cε(θ¯θ12)2L2(Ω)cε54.

On the other hand,

-εκ(η)σ(ξ,η)δ~δ(ξ)ξ(θ¯θ0+εθ¯θ12)+εκ(η)δ~δ(ξ)ξθ¯θ0L2(Ω)
=-εεκ(η)ξθ¯θ12δ~δ(ξ)-εκ(η)j=1(κ(η)εξ¯ξ)jδ~δ(ξ)ξ(θ¯θ0+εθ¯θ12)L2(Ω)
(3.35)cε54,

by Remark 3.2 and (3.27). We infer from Lemma 3.4, (3.34) and (3.35) that

R12δ~δ(ξ)L2(Ω)cε54,

while E12 and G12 are e.s.t. We now find from (3.32) that

(-εΔw12+g(uε)-g(u0+θ0+εθ12),w12)
=(-εΔuε+g(uε)-g(u0),w12)+(εΔu0,w12)+(εΔ(θ0+εθ12)-g(u0+θ0+εθ12)+g(u0),δ~δ(ξ)w12)
(3.36)=(εΔu0,w12)+((G12+E12)δ~δ(ξ),w12)+(R12δ~δ(ξ),w12).

We then find from (3.36) and the mean value theorem that

(-εΔw12+g(ζ)w12,w12)=(εΔu0,w12)+((G12+E12)δ~δ(ξ),w12)+(R12δ~δ(ξ),w12)

for some ζ between uε and (u0+θ0+εθ12). We then conclude

(-εΔw12+g(ζ)w12,w12)cεw12L2(Ω),

which implies

εw12H1(Ω)2+λw12L2(Ω)2cε2+λ2w12L2(Ω)2.

3.3 Boundary layer analysis at arbitrary orders εn and εn+12, n0

Similarly as in (2.20), we formally write

(3.37)-εΔ(j=0εj(θj+εθj+12))+g(j=0εj(uj+θj+εθj+12))-g(j=0εjuj)=0.

Because of the one-dimensional nature of these boundary layers near the boundary in the direction normal to the boundary, we now introduce the boundary fitted coordinates. We transform the Laplacian Δ as in (3.7) and (3.8).

Using the geometric series expansions

(1-r)-1=l=0rl,(1-r)-2=l=0(l+1)rland(1-r)-3=12l=0(l+1)(l+2)rl,

we formally write, e.g.,

(3.38)σ3(ξ,η)=(1-κ(η)ξ)-3=12l=0(l+1)(l+2)(κ(η)ξ¯ξ)lεl2.

Then,

εΔ(j=0(εjθ¯θj+εj+12θ¯θj+12))=2ξ¯ξ2(j=0(εjθ¯θj+εj+12θ¯θj+12))-κ(η)(l=0(κ(η)ξ¯ξ)lεl2)ξ¯ξ(j=0(εj+12θ¯θj+εj+1θ¯θj+12))
+(l=0(l+1)(κ(η)ξ¯ξ)lεl2)2η2(j=0(εj+1θ¯θj+εj+32θ¯θj+12))
+ξ¯ξ2κ(η)(l=0(l+1)(l+2)(κ(η)ξ¯ξ)lεl2)η(j=0(εj+32θ¯θj+εj+2θ¯θj+12))
(3.39)=2ξ¯ξ2(j=0(εjθ¯θj+εj+12θ¯θj+12))+I1+I2+I3.

Using

l=0(κ(η)ξ¯ξ)lεl2=m=0(κ(η)ξ¯ξ)2mεm+m=0(κ(η)ξ¯ξ)2m+1εm+12,

we rearrange the terms according to the order of ε and then we find the last expression in (3.39). The Ik are defined and expanded as follows

(3.40){I1n=-κ(η)j=0n(εjJ1j(θ¯θ)+εj+12J1j+12(θ¯θ)),I2n=j=0n(εjJ2j(θ¯θ)+εj+12J2j+12(θ¯θ)),I3n=ξ¯ξ2κ(η)j=0n(εjJ3j(θ¯θ)+εj+12J3j+12(θ¯θ)),

where

(3.41){J1j(θ¯θ)=k=0j-1(κ(η)ξ¯ξ)2(j-k)-1θ¯θkξ¯ξ+k=0j-1(κ(η)ξ¯ξ)2(j-k)-2θ¯θk+12ξ¯ξ,J1j+12(θ¯θ)=k=0j(κ(η)ξ¯ξ)2(j-k)θ¯θkξ¯ξ+k=0j-1(κ(η)ξ¯ξ)2(j-k)-1θ¯θk+12ξ¯ξ,
(3.42){J2j(θ¯θ)=k=0j-1(2(j-k)-1)(κ(η)ξ¯ξ)2(j-k)-22θ¯θkη2+k=0j-2(2(j-k)-2)(κ(η)ξ¯ξ)2(j-k)-32θ¯θk+12η2,J2j+12(θ¯θ)=k=0j-1(2(j-k))(κ(η)ξ¯ξ)2(j-k)-12θ¯θkη2+k=0j-1(2(j-k)-1)(κ(η)ξ¯ξ)2(j-k)-22θ¯θk+12η2,
(3.43){J3j(θ¯θ)=k=0j-2(2(j-k)-2)(2(j-k)-1)(κ(η)ξ¯ξ)2(j-k)-3θ¯θkη+k=0j-2(2(j-k)-3)(2(j-k)-2)(κ(η)ξ¯ξ)2(j-k)-4θ¯θk+12η,J3j+12(θ¯θ)=k=0j-1(2(j-k)-1)(2(j-k))(κ(η)ξ¯ξ)2(j-k)-2θ¯θkη+k=0j-2(2(j-k)-2)(2(j-k)-1)(κ(η)ξ¯ξ)2(j-k)-3θ¯θk+12η.

We note that when the upper limit and exponents in the expressions of Jkj and Jkj+12 are negative these terms are set to be zero. To handle the nonlinear term g, we derive the analogue of Lemma 2.6. Replacing ε by ε, n by 2n+d (d=0,1), and then renaming u2j, θ2j, u2j+1, θ2j+1, respectively, as uj, θj, uj+12, θj+12 and setting uj+12=0, we obtain the following lemma.

Lemma 3.8.

There exists a constant C>0, independent of ε, such that

(3.44){|g(j=0n(εjuj+εjθj)+j=0n-1εj+12θj+12)-g(j=0nεjuj)-Gn|Cε2n+1,|g(j=0n(εjuj+εjθj+εj+12θj+12))-g(j=0nεjuj)-Gn+12|Cε2n+2,

where, for d=0,1,

(3.45)Gn+d2=m=02n+d{k=0m|α|=k,α1+2α2++mαm=m(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εm,

and the multi-index notations are defined, for d=0,1, as follows

(3.46){α=(α1,,α2j+d),|α|=α1++α2j+d,(u+θ)α=(u12+θ12)α1(uj+d2+θj+d2)α2j+d,uα=(u12)α1(uj+d2)α2j+d,

with uj+12=(uj+12)α2j+1=0 if α2j+11 and (uj+12)0=1 for j=0,1,2,.

We now construct high order boundary layers. From the formal sum (3.37) and Lemma 3.8 with n=, we can write

-εΔ(j=0εj(θj+εθj+12))
(3.47)+m=0{k=0m|α|=k,α1+2α2++mαm=m(kα)1k![g(k)(u0+θ0)(u+θ)α-g(k)(u0)uα]}εm=0.

We then observe that at order 𝒪(ε)

-θ¯θξ¯ξξ¯ξ1+g(u0+θ¯θ0)θ¯θ1=(g(u0)-g(u0+θ¯θ0))u1-g′′(u0+θ¯θ0)2(θ¯θ12)2-κ(η)((κ(η)ξ¯ξθ¯θξ¯ξ0+θ¯θξ¯ξ)12+θ¯θηη0.

Incorporating (3.39) and (3.40), and from (3.47) at m=2j, we obtain at order 𝒪(εj) for j1 that

-θ¯θξ¯ξξ¯ξj+g(u0+θ¯θ0)θ¯θj=-(g(u0+θ¯θ0)-g(u0))uj
-{k=22j|α|=k,α1+2α2++2jα2j=2j(kα)1k![g(k)(u0+θ¯θ0)(u+θ¯θ)α-g(k)(u0)uα]}
(3.48)-κ(η)J1j(θ¯θ)+J2j(θ¯θ)+ξ¯ξ2κ(η)J3j(θ¯θ),

where the multi-index notation is described in (3.46). On the other hand, at order 𝒪(εj+12) for j1, from (3.47) at m=2j+1 we similarly find that

-θ¯θξ¯ξξ¯ξj+12+g(u0+θ¯θ0)θ¯θj+12=-(g(u0+θ¯θ0)-g(u0))uj+12
-{k=22j+1|α|=k,α1+2α2++(2j+1)α2j+1=2j+1(kα)1k![g(k)(u0+θ¯θ0)(u+θ¯θ)α-g(k)(u0)uα]}
(3.49)-κ(η)J1j+12(θ¯θ)+J2j+12(θ¯θ)+ξ¯ξ2κ(η)J3j+12(θ¯θ).

The two equations (3.48) and (3.49) are, respectively, supplemented with the boundary conditions

{θ¯θj=-uj|ξ=0=-uj(X(η),Y(η))at ξ=0,θ¯θj=0at ξ=ξ0,  {θ¯θj+12=0at ξ=0,θ¯θj+12=0at ξ=ξ0.

As before, we extend θ¯θj and θ¯θj+12 by zero for ξ>ξ0 and these extensions are still denoted by the same notations.

Multiplying, respectively, (3.14) by ε0, (3.15) by ε12, (3.48) by εj and (3.49) by εj+12, from j=1 to j=n, and adding these resulting equations we find that

-j=0nεj(θ¯θξ¯ξξ¯ξj+εθ¯θξ¯ξξ¯ξj+12)
(3.50)=-m=02n+1{k=0m|α|=k,α1+2α2++mαm=m(kα)1k![g(k)(u0+θ¯θ0)(u+θ¯θ)α-g(k)(u0)uα]}εm+I1n+I2n+I3n,

where Ikn are described as in (3.40).

Lemma 3.9.

For l,m,n0 and j=0,1,2,, there exist c>0 such that, pointwise,

(3.51)|ξnl+mθ¯θjξlηm|+|ξnl+mθ¯θj+12ξlηm|cεn-l2exp(-12λεξ).

Moreover, θ¯θj and θ¯θj+12 satisfy, for j=0,1,2,,

(3.52)ξnl+mθ¯θjξlηmLξ2(Ω)+ξnl+mθ¯θj+12ξlηmLξ2(Ω)cεn-l2+14.

Proof.

Using Lemma 3.3 and Lemma 3.4 and the induction on j, from (3.48) and (3.49), we obtain the estimates (3.19) and (3.25) for θ¯θj and θ¯θj+12, j=0,1,2,. Here, as indicated in (2.38), for k2 we used the fact that the right-hand side of (3.48) and (3.49) involve only the preceding boundary layer correctors. From the pointwise estimates (3.51), we readily obtain the L2- norm estimates (3.52). ∎

Lemma 3.10.

There exists c>0 such that

(εΔ(j=0n(εjθ¯θj+εj+12θ¯θj+12))-In)δ~δ(ξ)L2(Ω)cεn+54,

where δ~δ(ξ) is given in (3.28), and

In=j=0nεj(θ¯θj+εθ¯θj+12)ξ¯ξξ¯ξ+I1n+I2n+I3n

with Iin given by (3.40).

Proof.

Using the Laplacian (3.7) in terms of ξ,η, we write

(3.53)εΔ(j=0n(εjθ¯θj+εj+12θ¯θj+12))-εj=0n(εjθ¯θj+εj+12θ¯θj+12)ξξ=K1n+K2n+K3n,

where

K1n=-εκ(η)σ(ξ,η)ξ(j=0n(εjθ¯θj+εj+12θ¯θj+12)),K2n=εσ2(ξ,η)2η2(j=0n(εjθ¯θj+εj+12θ¯θj+12)),K3n=εξκ(η)σ3(ξ,η)η(j=0n(εjθ¯θj+εj+12θ¯θj+12)).

We now only have to prove that

(3.54)(Kin-Iin)δ~δ(ξ)L2κεn+54,i=1,2,3.

Noting that j=0nk=0j=k=0nj=kn and j=0nk=0j-1=k=0nj=k+1n, from (3.40) and (3.41) we write

I1n=-κ(η)k=0n(j=k+1nεj(κ(η)ξ¯ξ)2(j-k)-1+j=knεj+12(κ(η)ξ¯ξ)2(j-k))θ¯θkξ¯ξ
-κ(η)k=0n(j=k+1nεj(κ(η)ξ¯ξ)2(j-k)-2+j=k+1nεj+12(κ(η)ξ¯ξ)2(j-k)-1)θ¯θk+12ξ¯ξ
=-κ(η)k=0n(εk+1m=02(n-k)(κ(η)ξ)m)θ¯θkξ-κ(η)k=0n(εk+32m=02(n-k)-1(κ(η)ξ)m)θ¯θk+12ξ.

Then, we find

(K1n-I1n)δ~δ(ξ)L2εκ(η)(j=0nεj(σ(ξ,η)-m=02(n-j)(κ(η)ξ)m)θ¯θjξ)δ~δ(ξ)L2
+εκ(η)(j=0nεj+12(σ(ξ,η)-m=02(n-j)-1(κ(η)ξ)m)θ¯θj+12ξ)δ~δ(ξ)L2
cεj=0n(εj(κ(η)ξ)2(n-j)+1δ~δ(ξ)θ¯θjξL2+εj+12(κ(η)ξ)2(n-j)δ~δ(ξ)θ¯θj+12ξL2)
cεn+54,

where in the last inequality we used Lemma 3.9. Note that in (3.42) and (3.43) the sum k=0j-2 can be replaced by k=0j-1 because the terms there for k=j-1 do not contribute. Then, permuting summations as above, we can also prove (3.54) for i=2,3. The lemma thus follows. ∎

We use Lemmas 3.8 and 3.10, and equation (3.50), to find that

(3.55)-εΔ(j=0n(εjθ¯θj+εj+12θ¯θj+12))+g(j=0n(εjuj+εjθ¯θj+εj+12θ¯θj+12))-g(j=0nεjuj)=Rn+R,

where

Rnδ~δ(ξ)L2cεn+54,|R|cε2n+2.

As before, we define now θj=θ¯θjδ(ξ) and θj+12=θ¯θj+12δ(ξ) for each j0, where δ is defined as in (3.17). Note that θj,θj+12 satisfy the boundary conditions as in (3.12).

Theorem 3.11.

Assume that f is a smooth function, Ω is a general smooth domain and uε is solution of (2.1). Let uj and θj satisfy (3.3) and (3.48), respectively. Then, for every n0, there exists a constant c>0, independent of ε, such that

(3.56)uε-j=0nεj(uj+θj+εθj+12)εcεn+1,
(3.57)uε-j=0nεj(uj+θj)-j=0n-1εj+12θj+12εcεn+34.

Proof.

We set

wn+12=uε-j=0nεj(uj+θj+εθj+12).

We use the smooth cut-off function δ~δ(ξ) to eliminate the singularity of σ(ξ,η) where δ~δ(ξ) is given by (3.28). Then, by a similar argument as before, we obtain that

(εΔj=0nεj(θj+εθj+12)-g(j=0nεj(uj+θj+εθj+12))+g(j=0nεjuj),wn+12)
=((Gn+12+En+12)δ~δ(ξ),wn+12)+(Rn+12δ~δ(ξ),wn+12),

where

En+12=εΔ(j=0nεj(θj+εθj+12)-j=0nεj(θ¯θj+εθ¯θj+12)),
Gn+12=g(j=0nεj(uj+θ¯θj+εθ¯θj+12))-g(j=0nεj(uj+θj+εθj+12)),
Rn+12=Rn+R,

where Rn and R are given in (3.55). We note that En+12 and Gn+12 are e.s.t., and Rn+12δ~δ(ξ)L2(Ω)εn+1. We now find

(-εΔwn+12+g(uε)-g(j=0nεj(uj+θj+εθj+12)),wn+12)
=(εn+1Δun,wn+12)+((Gn+12+En+12)δ~δ(ξ),wn+12)+(Rn+12δ~δ(ξ),wn+12).

Here, we used (2.41) which is obtained from summing (3.3).

We finally obtain from the mean value theorem that, for some ζ between uε and j=0nεj(uj+θj+εθj+12),

(-εΔwn+12+g(ζ)wn+12,wn+12)cεn+1wn+12L2(Ω).

This proves (3.56). From Lemma 3.9, we note that εn+12θn+12εcεn+34, and the estimate (3.57) follows from (3.56). ∎

Award Identifier / Grant number: DMS 1206438

Award Identifier / Grant number: DMS 1510249

Award Identifier / Grant number: 2015K2A1A2070543

Award Identifier / Grant number: 2015R1D1A1A01059837

Funding statement: This work was supported in part by NSF grants DMS 1206438 and 1510249, by a research fund of Indiana University. Chang-Yeol Jung was supported under the framework of international cooperation program managed by the National Research Foundation of Korea (grant no. 2015K2A1A2070543) and supported by the National Research Foundation of Korea grant funded by the Ministry of Education (grant no. 2015R1D1A1A01059837).

References

[1] G. K. Batchelor, An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge, 1988. Suche in Google Scholar

[2] R. E. Caflisch and M. Sammartino, Existence and singularities for the Prandtl boundary layer equations, ZAMM Z. Angew. Math. Mech. 80 (2000), no. 11–12, 733–744. 10.1002/1521-4001(200011)80:11/12<733::AID-ZAMM733>3.0.CO;2-LSuche in Google Scholar

[3] M. Cannone, M. C. Lombardo and M. Sammartino, Well-posedness of Prandtl equations with non-compatible data, Nonlinearity 26 (2013), no. 12, 3077–3100. 10.1088/0951-7715/26/12/3077Suche in Google Scholar

[4] M. Cannone, M. C. Lombardo and M. Sammartino, On the Prandtl boundary layer equations in presence of corner singularities, Acta Appl. Math. 132 (2014), 139–149. 10.1007/s10440-014-9912-1Suche in Google Scholar

[5] K. W. Chang and F. A. Howes, Nonlinear Singular Perturbation Phenomena: Theory and Applications, Appl. Math. Sci. 56, Springer, New York, 1984. 10.1007/978-1-4612-1114-3Suche in Google Scholar

[6] B. Desjardins, E. Grenier, P.-L. Lions and N. Masmoudi, Incompressible limit for solutions of the isentropic Navier–Stokes equations with Dirichlet boundary conditions, J. Math. Pures Appl. (9) 78 (1999), no. 5, 461–471. 10.1016/S0021-7824(99)00032-XSuche in Google Scholar

[7] A. Ducrot and M. Langlais, A singular reaction-diffusion system modelling prey-predator interactions: Invasion and co-extinction waves, J. Differential Equations 253 (2012), no. 2, 502–532. 10.1016/j.jde.2012.04.005Suche in Google Scholar

[8] W. E., Boundary layer theory and the zero-viscosity limit of the Navier–Stokes equation, Acta Math. Sin. (Engl. Ser.) 16 (2000), no. 2, 207–218. 10.1007/s101140000034Suche in Google Scholar

[9] W. Eckhaus and E. M. de Jager, Asymptotic solutions of singular perturbation problems for linear differential equations of elliptic type, Arch. Ration. Mech. Anal. 23 (1966), 26–86. 10.1007/BF00281135Suche in Google Scholar

[10] S.-I. Ei and H. Matsuzawa, The motion of a transition layer for a bistable reaction diffusion equation with heterogeneous environment, Discrete Contin. Dyn. Syst. 26 (2010), no. 3, 901–921. 10.3934/dcds.2010.26.901Suche in Google Scholar

[11] T. Funaki, Singular limit for stochastic reaction-diffusion equation and generation of random interfaces, Acta Math. Sin. (Engl. Ser.) 15 (1999), no. 3, 407–438. 10.1007/BF02650735Suche in Google Scholar

[12] S. Gaucel and M. Langlais, Some remarks on a singular reaction-diffusion system arising in predator-prey modeling, Discrete Contin. Dyn. Syst. Ser. B 8 (2007), no. 1, 61–72. 10.3934/dcdsb.2007.8.61Suche in Google Scholar

[13] G.-M. Gie and C.-Y. Jung, Vorticity layers of the 2D Navier–Stokes equations with a slip type boundary condition, Asymptot. Anal. 84 (2013), no. 1–2, 17–33. 10.3233/ASY-131164Suche in Google Scholar

[14] G.-M. Gie, C.-Y. Jung and R. Temam, Recent progresses in boundary layer theory, Discrete Contin. Dyn. Syst. 36 (2016), 2521–2583. 10.3934/dcds.2016.36.2521Suche in Google Scholar

[15] J. Grasman, On the Birth of Boundary Layers, Math. Centre Tracts 36, Mathematisch Centrum, Amsterdam, 1971. Suche in Google Scholar

[16] M. Hamouda and R. Temam, Boundary layers for the Navier–Stokes equations. The case of a characteristic boundary, Georgian Math. J. 15 (2008), no. 3, 517–530. 10.1515/GMJ.2008.517Suche in Google Scholar

[17] H. D. Han and R. B. Kellogg, The use of enriched subspaces for singular perturbation problems, Proceedings of the China–France Symposium on Finite Element Methods (Beijing 1982), Science Press, Beijing (1983), 293–305. Suche in Google Scholar

[18] H. Han and R. B. Kellogg, Differentiability properties of solutions of the equation -ε2Δu+ru=f(x,y) in a square, SIAM J. Math. Anal. 21 (1990), no. 2, 394–408. 10.1137/0521022Suche in Google Scholar

[19] D. Han, A. L. Mazzucato, D. Niu and X. Wang, Boundary layer for a class of nonlinear pipe flow, J. Differential Equations 252 (2012), no. 12, 6387–6413. 10.1016/j.jde.2012.02.012Suche in Google Scholar

[20] Y. Hong, C.-Y. Jung and R. Temam, On the numerical approximations of stiff convection-diffusion equations in a circle, Numer. Math. 127 (2014), no. 2, 291–313. 10.1007/s00211-013-0585-xSuche in Google Scholar

[21] C.-Y. Jung, Finite elements scheme in enriched subspaces for singularly perturbed reaction-diffusion problems on a square domain, Asymptot. Anal. 57 (2008), no. 1–2, 41–69. 10.3233/ASY-2008-0865Suche in Google Scholar

[22] C.-Y. Jung and T. B. Nguyen, Semi-analytical numerical methods for convection-dominated problems with turning points, Int. J. Numer. Anal. Model. 10 (2013), no. 2, 314–332. Suche in Google Scholar

[23] C.-Y. Jung, M. Petcu and R. Temam, Singular perturbation analysis on a homogeneous ocean circulation model, Anal. Appl. (Singap.) 9 (2011), no. 3, 275–313. 10.1142/S0219530511001832Suche in Google Scholar

[24] T. Kato, Remarks on the Euler and Navier–Stokes equations in 𝐑2, Nonlinear Functional Analysis and its Applications, Part 2 (Berkeley 1983), Proc. Sympos. Pure Math. 45, American Mathematical Society, Providence (1986), 1–7. 10.1090/pspum/045.2/843590Suche in Google Scholar

[25] J. P. Kelliher, Vanishing viscosity and the accumulation of vorticity on the boundary, Commun. Math. Sci. 6 (2008), no. 4, 869–880. 10.4310/CMS.2008.v6.n4.a4Suche in Google Scholar

[26] R. B. Kellogg and M. Stynes, Corner singularities and boundary layers in a simple convection-diffusion problem, J. Differential Equations 213 (2005), no. 1, 81–120. 10.1016/j.jde.2005.02.011Suche in Google Scholar

[27] J. Kevorkian and J. D. Cole, Multiple Scale and Singular Perturbation Methods, Appl. Math. Sci. 114, Springer, New York, 1996. 10.1007/978-1-4612-3968-0Suche in Google Scholar

[28] M. C. Lopes Filho, A. L. Mazzucato, H. J. Nussenzveig Lopes and M. Taylor, Vanishing viscosity limits and boundary layers for circularly symmetric 2D flows, Bull. Braz. Math. Soc. (N.S.) 39 (2008), no. 4, 471–513. 10.1007/s00574-008-0001-9Suche in Google Scholar

[29] A. Mazzucato, D. Niu and X. Wang, Boundary layer associated with a class of 3D nonlinear plane parallel channel flows, Indiana Univ. Math. J. 60 (2011), no. 4, 1113–1136. 10.1512/iumj.2011.60.4479Suche in Google Scholar

[30] A. Mazzucato and M. Taylor, Vanishing viscosity limits for a class of circular pipe flows, Comm. Partial Differential Equations 36 (2011), no. 2, 328–361. 10.1080/03605302.2010.505973Suche in Google Scholar

[31] M. Mimura, Reaction-diffusion systems arising in biological and chemical systems: application of singular limit procedures, Mathematical Aspects of Evolving Interfaces (Funchal 2000), Lecture Notes in Math. 1812, Springer, Berlin (2003), 89–121. 10.1007/978-3-540-39189-0_3Suche in Google Scholar

[32] J. Mo and K. Zhou, Singular perturbation for nonlinear species group reaction diffusion systems, J. Biomath. 21 (2006), no. 4, 481–488. Suche in Google Scholar

[33] P. Mottoni and F. Rothe, A singular perturbation analysis for a reaction-diffusion system describing pattern formation, Stud. Appl. Math. 63 (1980), no. 3, 227–247. 10.1002/sapm1980633227Suche in Google Scholar

[34] B. O’Neill, Elementary Differential Geometry, Academic Press, New York, 2006. 10.1016/B978-0-12-088735-4.50011-0Suche in Google Scholar

[35] R. E. O’Malley, Jr., Introduction to Singular Perturbations, Appl. Math. Mech. 14, Academic Press, New York, 1974. Suche in Google Scholar

[36] R. E. O’Malley, Jr., Singular Perturbation Methods for Ordinary Differential Equations, Appl. Math. Sci. 89, Springer, New York, 1991. 10.1007/978-1-4612-0977-5Suche in Google Scholar

[37] R. E. O’Malley, Jr., Singularly perturbed linear two-point boundary value problems, SIAM Rev. 50 (2008), no. 3, 459–482. 10.1137/060662058Suche in Google Scholar

[38] C. H. Ou and R. Wong, Shooting method for nonlinear singularly perturbed boundary-value problems, Stud. Appl. Math. 112 (2004), no. 2, 161–200. 10.1142/9789814656054_0046Suche in Google Scholar

[39] C. V. Pao, Singular reaction diffusion equations of porous medium type, Nonlinear Anal. 71 (2009), no. 5–6, 2033–2052. 10.1016/j.na.2009.01.122Suche in Google Scholar

[40] D. V. Redžić, The operator in orthogonal curvilinear coordinates, Eur. J. Phys. 22 (2001), 595–599. 10.1088/0143-0807/22/6/304Suche in Google Scholar

[41] H.-G. Roos, M. Stynes and L. Tobiska, Numerical Methods for Singularly Perturbed Differential Equations, Springer Ser. Comput. Math. 24, Springer, Berlin, 1996. 10.1007/978-3-662-03206-0Suche in Google Scholar

[42] S.-D. Shih and R. B. Kellogg, Asymptotic analysis of a singular perturbation problem, SIAM J. Math. Anal. 18 (1987), no. 5, 1467–1511. 10.1137/0518107Suche in Google Scholar

[43] R. Temam and X. Wang, Boundary layers associated with incompressible Navier–Stokes equations: The noncharacteristic boundary case, J. Differential Equations 179 (2002), no. 2, 647–686. 10.1006/jdeq.2001.4038Suche in Google Scholar

[44] F. Verhulst, Methods and Applications of Singular Perturbations, Texts Appl. Math. 50, Springer, New York, 2005. 10.1007/0-387-28313-7Suche in Google Scholar

Received: 2015-11-1
Revised: 2015-11-29
Accepted: 2015-12-7
Published Online: 2016-1-14
Published in Print: 2017-8-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Heruntergeladen am 30.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/anona-2015-0148/html
Button zum nach oben scrollen