• Aucun résultat trouvé

Nash–Moser iteration and singular perturbations

N/A
N/A
Protected

Academic year: 2022

Partager "Nash–Moser iteration and singular perturbations"

Copied!
29
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Nash–Moser iteration and singular perturbations

Benjamin Texier

a,,1

, Kevin Zumbrun

b,2

aUniversité Paris Diderot (Paris 7), Institut de Mathématiques de Jussieu, UMR CNRS 7586, Paris, France bIndiana University, Bloomington, IN 47405, United States

Received 27 July 2009; received in revised form 2 November 2010; accepted 3 March 2011 Available online 19 May 2011

Abstract

We present a simple and easy-to-use Nash–Moser iteration theorem tailored for singular perturbation problems admitting a formal asymptotic expansion or other family of approximate solutions depending on a parameterε→0. The novel feature is to allow loss of powers ofεas well as the usual loss of derivatives in the solution operator for the associated linearized problem. We indicate the utility of this theorem by describing sample applications to (i) systems of quasilinear Schrödinger equations, and (ii) existence of small-amplitude profiles of quasilinear relaxation systems in the degenerate case that the velocity of the profile is a characteristic mode of the hyperbolic operator.

©2011 Elsevier Masson SAS. All rights reserved.

1. Introduction

Because the expansions themselves furnish arbitrarily accurate approximate solutions, and because the associated linearized estimates are often stiff in terms of amplitude and or smoothness, Nash–Moser iteration appears particularly well-adapted to the verification of asymptotic expansions such as arise in various singular perturbation problems depending on a small parameterε→0. However, standard Nash–Moser theorems allow only for loss of derivatives and not loss of powers ofεin the estimates on the linearized solution operator, so that to apply Nash–Moser iteration to problems that do lose powers ofεwould appear to require a careful accounting of constants throughout the entire Nash–Moser iteration to check that the argument closes.

The purpose of this article therefore is to present a simple and general-purpose theorem carrying out this account- ing, which can be applied as an easy-to-use black box to this type of problem. We conclude by presenting two sample applications for which both loss of derivatives and of powers ofεnaturally occur for the linearized problem, one for systems of quasilinear Schrödinger equations and one in existence of small-amplitude profiles of quasilinear relaxation systems. The latter, due to Métivier, Texier and Zumbrun, was treated in [17] by the approach presented here; though special cases may be treated by other methods [23,9,10,3], we do not know of any other solution in the generality considered there.

* Corresponding author.

E-mail addresses:texier@math.jussieu.fr (B. Texier), kzumbrun@indiana.edu (K. Zumbrun).

1 Research of B.T. was partially supported under NSF grant number DMS-0505780. B.T. acknowledges support from the Agence Nationale de la Recheche through grant ANR-08-BLAN-0301-01.

2 Research of K.Z. was partially supported under NSF grants No. DMS-0300487 and DMS-0801745.

0294-1449/$ – see front matter ©2011 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2011.05.001

(2)

Our approach follows a very simple proof given by Xavier Saint-Raymond [22] of a (parameter-independent) Nash–Moser implicit function theorem [8,18] in a Sobolev space setting. A novel aspect is our treatment of unique- ness, which we have not seen elsewhere – in particular the incorporation of a phase condition in the case that the linearized operator has a kernel. (See Theorem 2.20.)

We note that a parameter-dependent Nash–Moser scheme was recently used by Alvarez-Samaniego and Lannes [2]

to prove local-in-time well-posedness of model equations in oceanography. Alvarez-Samaniego and Lannes do not allow losses inεin the linearized solution operator, which is the main point here.

In [5], Iooss, Plotnikov and Toland use a parameter-dependent Nash–Moser to prove the existence of periodic standing water-waves in the case of an infinite depth. They deal with a situation in which the linearized operator might not be invertible on small sets on ε,and proceed by taking out corresponding bad regions inε. Here we consider everywhere invertible linearized operators, with losses (materialized by inverse powers ofεin the estimates) for all values of the parameterε,and proceed by keeping track of these losses.

An important reference on Nash–Moser-type theorems is Hamilton [4]. Another good reference is Alinhac and Gérard’s book [1].

Plan of the paper and scheme of the main proof.In Section 2, we state the main theorem onε-dependent Nash–

Moser iteration, giving the proof afterward in Section 3.

The proof classically combines an iteration step (Newton’s method) with a regularization procedure (in Sobolev spaces, high-frequency truncation). It is largely based on Xavier Saint-Raymond’s elegant and short proof [22]. The key in our context is to show that the bounds can be made uniform in the small parameter. Our main observation is that this can be achieved under the condition that the regularizing sequence is diverging to+∞fast enough, depending on the small parameter (see (3.19)).

In Section 4, we describe applications to systems of quasilinear Schrödinger equations, and in Section 5 to existence of small-amplitude profiles of quasilinear relaxation systems in the degenerate case that the velocity of the relaxation profile is a characteristic mode of the hyperbolic operator.

2. A simple Nash–Moser theorem 2.1. Setting

Consider two families of Banach spaces{Es}s∈R,{Fs}s∈R,and a family of equations Φε

uε

=0, uεEs, (2.1)

indexed byε(0,1),where for somem0, s0, s1∈R,withs0+2ms1,for allε,

ΦεC2(Es, Fsm), for alls0+mss1. (2.2)

Let| · |s denote the norm inEs and · sdenote the norm inFs.The norms| · |sand · s and spacesEs andFs are possiblyε-dependent, as in our applications in Sections 4 and 5. We assume that the embeddings

EsEs, FsFs, ss, (2.3)

hold, and have norms less than one:

| · |s| · |s, · s · s, ss. (2.4) We assume the interpolation property:

| · |s+σ| · |sσ−σσ | · |sσσ+σ, 0< σ < σ, (2.5) where|u|s|v|s stands for|u|s C|v|s,for someC >0 possibly depending ons ands but not onε,nor onu andv.We assume in addition the existence of a family of regularizing operators

Sθ:EsEs, θ >0, such that for allss,

(3)

|Sθuu|sθss|u|s, (2.6)

|Sθu|s

1+θss

|u|s. (2.7)

Example 2.1.Let Es=Hs

Rd

, Fs=Hs Rd

×Hs+1 Rd

, withε-dependent norms defined by

|v|s:= vHεs :=1+ |εξ|2s/2

(Fv)(ξ )

L2(Rdξ), (f, g)

s= |f|s+ |g|s+1, (2.8) whereFv denotes the Fourier transform ofv.Then (2.3), (2.4) and (2.5) hold. A family of regularizing operators EsEs is given by

Sθ:uSθ(u):=F1 χ

θ1ξ ˆ u

,

whereχ:Rd→ [0,1]is a smooth truncation function, identically equal to 1 for|ξ|1,and identically equal to 0 for

|ξ|2.The family{Sθ}θ >0satisfies (2.6) and (2.7).

Remark 2.2.The family of norms · Hεs satisfies Moser’s inequality uvHεs |u|LvHεs+ uHεs|v|L, s0, u, v∈HsL,

and, ifF is smooth and satisfiesF (0)=0,for some non-decreasingC:R+→R+, F (u)

Hεs C

|u|L

uHεs, s > d/2, uHs.

Example 2.3.Consider the family of maps Φε(u)=

1jd

Aj(u)ε∂xju, u

,

whereAj:u∈RnAj(u)∈Rn×nis smooth. By Moser’s inequality (Remark 2.2), fors >1+d/2,the application Φεmaps smoothlyHs toHs1×Hs,that is, with the notation of Example 2.1,EstoFs1.

2.2. First assumption: tame direct bounds

We assume bounds forΦεand its first two derivatives.

Assumption 2.4.For someγ00, γ10,for alls such thats0+mss1m,for allu, v, wEs+m,there holds

Φε(u)

sC0

1+ |u|s+m

, (2.9)

Φε (u)·v

sC1

εγ1|v|s0+m|u|s+m+ |v|s+m

, (2.10)

Φε

(u)·(v, w)

sC2

ε1|v|s0+m|w|s0+m|u|s+m+εγ1|w|s0+m|v|s+m+εγ1|v|s0+m|w|s+m

, (2.11) where the functionsCj=Cj(ε, (|u|s0+ ˜m)0m˜m)satisfy

sup

Cj, j=0,1,2, ε∈(0,1),|u|s0+mεγ0 <+∞. The simplest example is given by a product map:

Example 2.5. Consider the mapΦ0:uHs(Rd;R)(u2, u)Hs×Hs(Rd;R),for d/2< s. There holds, by Remark 2.2,

Φ0(u)

Hεs×Hεs |u|LuHεs+ uHεs,

(4)

where the weighted norm · Hεs is defined in (2.8). For smallε,the classical Sobolev embeddingHs L,for d/2< s0s,has a large norm ifHs is equipped with the weighted norm · Hεs:

|u|Lεd/2uHs0

ε , so that

Φ0(u)

Hεs×Hεs

1+εd/2uHs0

ε

uHεs,

and Φ0(u)·v

Hεs×Hεs

1+εd/2uHεs0

vHεs +εd/2vHεs0uHεs. In particular, Assumption 2.4 holds withm=0, γ0=γ1=d/2.

Another simple example is given by the map defined in Example 2.3:

Example 2.6.Consider the mapΦε:EsFs1,introduced in Example 2.3, where the families of functional spaces Es andFs and theirε-dependent norms were introduced in Example 2.1. Letd/2< s0<1+d/2 be given. Just like in the above remark, for smallε,the classical Sobolev embedding

HN+s0 Rd

WN, Rd

, N∈N, has a large norm

|v|WN,εNd/2|v|N+s0, (2.12)

where| · |s= · Hεs is defined in (2.8). By Remark 2.2, forss0, Aj(u)ε∂xju

sCAj

|u|L

|u|s+1+ |ε∂xju|L|u|s

,

whereCAj:R+→R+is non-decreasing and depends only onAj, sandd.By (2.12), forss0+1, Aj(u)ε∂xju

sCAj

εd/2|u|s0

1+εd/2|u|s0+1

|u|s+1,

and (2.9) holds with C0=1+

1jd

CAj

εd/2|u|s0

1+εd/2|u|s0+1

. (2.13)

Besides, Φε

(u)v

s

1jd

Aj(u)·v ε∂xju

s+Aj(u)ε∂xjv

s+ |v|s+1

1jd

CAj,A j

|u|L

εd/2|v|s0+1|u|s+1+

1+εd/2|u|s0+1

|v|s+1

and (2.10) holds withγ0=γ1=d/2 and C1=1+

1jd

CAj,A j

εd/2|u|s0

1+εd/2|u|s0+1

.

The bound forε)is similar. We conclude that Assumption 2.4 holds withγ0=γ1=d/2.

Remark 2.7. In connection with (2.12), where the embedding constant blows up to+∞ as εdecreases to 0, the

“constants”Cj in Assumption 2.4 arenotassumed to be non-decreasing in their arguments; in Example 2.6 this is reflected in (2.13).

Example 2.8.GivenT >0,consider the functional spaces Es=C1

[0, T], Hs1 Rd

C0

[0, T], Hs Rd

, Fs1=C0

[0, T], Hs1 Rd

×Hs Rd

,

(5)

with norms

|u|s:= sup

0tT

ε∂tu(t )

Hεs1+u(t )

Hεs

, (f1, f2)

s1:= sup

0tT

f1(t )

Hεs1+ f2Hεs, where the weighted Sobolev · Hεs norms are defined in (2.8).

Lets0ands1be given such thatd/2< s0<1+d/2< s1,withs0+2s1.Let a bounded family(fε)ε(0,1)Fs11;meaningfεFs11for allε,with supε(0,1)fεs11<.

Consider the family of maps defined by Φε(u):=

ε∂tu+

1jd

Aj(u)ε∂xjuf1ε, u|t=0f2ε

, (2.14)

where the mapsAj:u∈RnAj(u)∈Rn×n are smooth. Then,Φε mapsEs toFs1 fors0+1ss1,and we can check exactly as in Example 2.6 that Assumption 2.4 holds withγ0=γ1=d/2.

Remark 2.9.Assumption 2.4 is stable by shifts that are both smooth and small, in the following sense: ifΦεsatisfy Assumption 2.4, with associated parametersm, s0, s1,andγ0,and if(aε)ε(0,1)Es1,with supε(0,1)εγ0|aε|s0+m<

,thenΦ˜ε:=Φε(aε+ ·)satisfies Assumption 2.4 with the same parameters.

2.3. Second assumption: tame inverse bounds

Our second and key assumption states that ifuis small enough in some “low” norm, then ε)(u) has a right inverse, with a right-inverse bound (2.16) that is possibly stiff with respect toεand may show a loss of derivatives:

Assumption 2.10.For somer0, r0,there holdss0+ms1−max(m, r),and for someγ0, κ0,for all ssuch thats0+mss1−max(m, r),for alluEs+r such that

|u|s0+m+rεγ, (2.15)

the mapε)(u):Es+mFs has a right inverseΨε(u):

Φε

(u)Ψε(u)=Id :FsFs, satisfying, for allgFs+r,

Ψε(u)g

sκ

gs0+m+r|u|s+r+ gs+r

, (2.16)

whereC=C(ε,|u|s0+m+r)satisfies sup

C, ε(0,1),|u|s0+m+rεγ <.

Remark 2.11(On stiffness of the right-inverse bound (2.16)).The right-inverse bound (2.16) is stiff with respect toε ifκ >0.The caseκ=0 corresponds to no loss inεand is covered for instance by the result of Alvarez-Samaniego and Lannes [2].

Remark 2.12(On losses of derivatives in the right-inverse bound (2.16)).The loss of derivatives is parameterized byrandr.Estimate (2.16) states indeed that we can solve the linearized equationε)(u)v=g,with a bound for the solutionv=Ψε(u)gthat gives a control of the low norm|v|s in terms of high norms,|u|s+r andgs+r,of the background and source.

The caser=r=0 corresponds to no loss and can typically be handled by Picard iteration, as in the classical existence proof for quasilinear symmetric systems.

Estimate (2.16) in Assumption 2.10 istamewith respect torandr.This is essential: one may check that the proof of Theorem 2.19 (given in Section 3) collapses if it is not.

The distinction betweenrandris somewhat illusory, since we can always redefineFs andm,so thatr=0.We also note that the proof withr=0 is the same as withr=0.

(6)

A simple example of a family of maps satisfying Assumption 2.10 is given by systems of symmetric quasilinear equations in a semi-classical setting, as detailed in the following example. (This example is meant to test our theory in the favorable context of symmetric quasilinear systems, where a Lax iteration scheme is certainly preferable to a Nash–Moser scheme.)

Example 2.13.Consider the functional spaces and the family of maps introduced in Example 2.8. Again,T >0 and d/2< s0<1+d/2< s1,are given, withs1measuring the regularity of(f1ε, f2ε),ands0+2s1.Then, as discussed in Example 2.8, Assumption 2.4 holds withγ0=γ1=d/2 andm=1.

Givens such thats0+1ss1−1,anduEs+1, g=(g1, g2)Fs,consider the equationε)(u)v=g, corresponding to the linearized initial-value problem

⎧⎨

ε∂tv+

1jd

Aj(u)ε∂xjv=g1

1jd

Aj(u)vε∂xju, v|t=0=g2.

(2.17) Assume that the mapsAj take values in the symmetric matrices. Then, by the classical linear hyperbolic theory, there exists a uniquevEs solution of (2.17). We now show thatvsatisfies an estimate of the form (2.16).

The classical commutator estimate gives, for 0|α|sandwHs, e

(ε∂x)αAj(u)ε∂xjw, (ε∂x)αw

L2C

|u|L

|ε∂xu|Lw2Hεs+ uHεs|w|LwHεs

, (2.18)

where the weighted Sobolev norm · Hεs is defined in (2.8). Besides, by Remark 2.2, Aj(u)wε∂xju

HεsC

|u|L

|ε∂xju|L

wHεs+ uHεs|w|L

+ uHs+1 ε |w|L

. (2.19)

If we now use (2.12) withN =0,1,together with estimates (2.18) and (2.19), we find that the solutionv to (2.17) satisfies

v(t )2

Hεs g22Hεs+ t 0

ε1g1HεsvHεsdt+ t 0

C

ε1d/2uHs0+1 ε

v2Hεsdt

+ t 0

C

εd/2uHs0+1 ε

ε1d/2uHs+1

ε vHεs0vHεsdt. (2.20)

We now restrict to a backgrounduHs+1satisfying

|u|s0+2ε1+d/2. (2.21)

Then, by Gronwall’s Lemma, estimate (2.20) used withs=s0,implies the bound max[0,T]vHs0

ε g2Hs0

ε +ε1max

[0,T]g1Hs0

ε , (2.22)

which we use back in (2.20) to obtain max[0,T]vHεs g2Hεs+ε1max

[0,T]g1Hεs

+ε1d/2

g2Hεs0 +ε1max

[0,T]g1Hεs0

max[0,T]uHs+1

ε . (2.23)

By using Eq. (2.17)(i) directly, we find the bound ε∂tvHs1

ε C

|u|L,|ε∂xu|L

vHεs + uHεs

|v|L+ |ε∂xv|L

. (2.24)

Note in (2.24) the occurrence of|ε∂xv|L,which cannot be controlled by (2.22). This forces us to go back to (2.20) withs=s0+1.At this point we make full use of (2.21), while a bound for|u|s0+1sufficed in order to estimatevHεs, and obtain

max[0,T]vHs0+1

ε g2Hs0+1

ε +ε1max

[0,T]g1Hs0+1

ε . (2.25)

(7)

Combining (2.23), (2.24) and (2.25), we obtain that, ifusatisfies (2.21), then there holds

|v|sε2d/2gs0+1|u|s+1+ε1gs.

We can conclude that if for allj, Aj is symmetric, then the family of maps and functional spaces defined in Exam- ple 2.8 satisfy Assumption 2.10 withγ=1+d/2, κ=2+d/2, r=1 andr=0.

Remark 2.14. If Φε satisfies Assumption 2.10 with parameters γ , κ, r, r, given a family (aε)ε(0,1)Es1, if supε(0,1)εγ|aε|s0+m+r<,thenΦ˜ε:=Φε(aε+ ·)satisfies Assumption 2.10, with the same parameters asΦε. 2.4. Third assumption: existence of an approximate solution

We consider a family of mapsΦε:EsFsmsatisfying Assumptions 2.4 and 2.10, with associated parameters m, s0,s1,andγ0, γ1, γ , r, r,andκ.

Assumption 2.15.Letkbe such that

max(κ+γ0,2κ+γ1, κ+γ ) < k. (2.26)

For some positive functionp¯= ¯p(m, r, r, γ0, γ1, γ , κ, k)2m+max(r, r)specified in Remark 2.23, there holds s0+m+max

r, r

+ ¯p < s1, (2.27)

and, for somessatisfying s0+m+max

r, r

s < s1− ¯p, (2.28)

there holds Φε(0)

sεk. (2.29)

We first comment on (2.27):

Remark 2.16.In Example 2.13, the indexs1measures the regularity of the sourcef1ε and the initial datumf2ε;in this view inequality (2.27) should be understood as a regularity requirement on the data. In particular, as discussed in Remark 2.23, askapproaches from above the limiting value max(κ+γ0,2κ+γ1, κ+γ ),the parameterp¯blows up to+∞,meaning that we require the data to be infinitely regular in this limit.

In our examples, inequality (2.29) reflects the existence of a family of approximate solutions toΦε=0:

Remark 2.17.Consider a family of mapsΦεC2(Es, Fsm)satisfying Assumptions 2.4 and 2.10, and an associated family of approximate solutionsuεaEs1 to the equationsΦε=0,in the sense thatΦε(uεa)sεk,withkands satisfying (2.26)–(2.28). Then, the mapsΦ˜ε:=Φε(uεa+ ·)satisfy Assumption 2.15.

For quasilinear systems, small initial data give crude examples of approximate solutions:

Example 2.18.Consider the initial-value problem

tu+

1jd

Aj(u)∂xju=0, u|t=0=εσa(x), (2.30)

and the associated family of mapsΦε defined in (2.14), with (f1ε, f2ε)(0, εσa).Assume thataHs1,withs1 satisfying (2.27). As described in Examples 2.8 and 2.13, ifAjis symmetric for allj,then Assumptions 2.4 and 2.10 hold forΦεin the functional spacesEs, Fs introduced in Example 2.8, for anyT >0,forssatisfying (2.28).

If the maps uAj(u) satisfy the bounds |Aj(u)||u|, for0,in particular if they are -homogeneous, then there holds Φεσa)s11εσ (1+)+1.Using Remark 2.17 above, and the specific values ofγ0, γ1, γ and κ given in Examples 2.8 and 2.13, this implies that ifσ (1+) >3(1+d/2),then Assumption 2.15 is satisfied by Φ˜ε:=Φεσa+ ·).

(8)

2.5. Results

Our main result gives existence inEs+m,wheressatisfies (2.28), of a solution to Eq. (2.1).

Theorem 2.19(Existence). Under Assumptions2.4,2.10and2.15, forεsmall enough, there exists a real sequenceθjε, satisfyingθjε→ +∞asj→ +∞andεis held fixed, such that the sequence

uε0:=0, uεj+1:=uεj+Sθε

jvjε, vjε:= −Ψε uεj

Φε uεj

, (2.31)

is well defined and converges, as j → ∞and ε is held fixed, to a solution uε of (2.1)ins+mnorm, withs as in(2.28), which satisfies the bound

uε

sεkκ. (2.32)

In applications (Sections 4 and 5), we apply Theorem 2.19 to a mapΦε(uεa+ ·),with the notation of Remark 2.17, so that in practice Theorem 2.19 is not a result about small solutions: the smallness condition (2.32) bears on the perturbation variable u, the full solution to Φε=0 being uεa+u. The smallness condition (2.29) is an accuracy condition bearing on the approximate solutionuεa;we show in Remark 2.22 that this condition is sharp in the usual implicit function theorem setting without losses in derivatives.

We supplement the above existence result by the following local uniqueness theorem. Contrary to Theorem 2.19, Theorem 2.20 does not rely on a Nash–Moser iterative scheme.

Theorem 2.20(Local uniqueness). Under Assumptions2.4, 2.10and2.15, forεsmall enough, if(Φε)is invertible, i.e.,Ψεis also a left inverse, then the solution described in Theorem2.19is unique in a ball of radiuso(εmax(κ+γ10,γ )) ins0+2m+r norm. More generally, if uˆε is a second solution within this ball, then(uˆεuε)is approximately tangent toKer(Φε)(uε), in the sense that its distance ins0norm fromKer(Φε)(uε)iso(| ˆuεuε|s0). In particular, if Ker(Φε)(uε)is finite-dimensional, thenuis the unique solution in the ball satisfying the additional “phase condition”

Πuε

uˆεuε

=0, (2.33)

whereΠuε is any uniformly bounded projection ontoKer(Φε)(uε).(In a Hilbert space, any orthogonal projection ontoKer(Φε)(uε).)

In a non-Hilbertian context, the existence of such a projectionΠuεis discussed in Remark 2.25 below.

2.6. Remarks

We first remark that the proofs use only part of the information contained in (2.16) and (2.11):

Remark 2.21.An examination of the proofs of Theorems 2.19 and 2.20 reveals that for existence we require esti- mate (2.16) only forf in the image ofΦεorε), sinceΨεis estimated only in composition with one or the other of these operators, while for uniqueness we need only the estimate forΨε(u)(Φε)(u)that would result by composing estimates (2.16) and (2.11).

Next we remark that the approximation rate is sharp by comparison with the standard Newton scheme:

Remark 2.22.Forγ0, γκ+γ1, corresponding to a critical valuekc=2κ+γ1in (2.26), Theorem 2.19 states that a loss ofεκ in the linear estimates means that, with the notation of Remark 2.17,Φε(uεa)sε+γ1+η,anyη >0, is the accuracy needed on the approximate solution.

This condition is sharp even for convergence of a standard Newton iteration scheme uεn+1:=uεnΨε

uεn Φε

uεn

, uε0:=uεa, uεa

s+mεγ1,

for problems with no loss of derivatives (r=r=0), corresponding by the computation

(9)

Φε uε1

s= 1 0

(1t ) Φε

uε0+t

uε1uε0

·

uε1uε0, uε1uε0

s

ε1uε0

s+uε1uε0

s

+εγ1uε1uε02

s

εγ1uε1uε02

s

εγ1Ψε uε0

Φε uε02

s

ε(2κ+γ1)Φε uε02

s

εηΦε uε0

s

in the casem=0 to the condition that errorε(uεn)s)n∈Ndecreases at the first step.

We make precise the parameterp¯that appears in Assumption 2.15:

Remark 2.23.From (3.23), we find thatp¯is

¯

p=m+ inf

N >N0

(N+1)(m+max(r, r)+M)

1−κkM , (2.34)

with

N0:= κ+km

k(2κ+γ1), M:=max γ0

k k,1

2

1+γ1 k +m

N + κ kN

, m:=max

m+r, r .

We observe the following asymptotic behavior as k approaches from above the critical value kc :=

max(κ+γ0,2κ+γ1, κ+γ )given in (2.26):

• Ifkc=2κ+γ1,thenp¯blows up like(kkc)2askkc.

• Ifkc=κ+γ0,orkc=κ+γ ,thenp¯blows up like(kkc)1askkc. The phase condition (2.33) can in some situations be made explicit:

Remark 2.24.LetΠube a bounded projection onto ker(Φε)(u),as in (2.33). If the map(u, v)Πuvis continuous inEs0×Es0,uniformly inε,then the implicit phase condition (2.33) can be replaced by the explicit

Π0

uˆεuε

=0.

See [21, Section 2], for related discussions of uniqueness up to phase conditions.

We discuss the existence of the projection mentioned in Theorem 2.20:

Remark 2.25. We first remark that if Ker(Φε)(uε)is finite-dimensional, then a bounded projection exists by the Hahn–Banach Theorem; see, e.g., [19].

In the infinite-dimensional case, we note that a projectionΠ onto a subspaceS of a Banach space is bounded if and only if the distance fromsS to KerΠ is greater than or equal to|s|/C for some uniformC >0. (Indeed,

|s| = |Π (st )|C|st|for allsS,t∈KerΠis equivalent to the statement thatΠis bounded, sincest runs over the entire Banach space assandtare varied.)

This implies that if there is an isometry between spaces Fsε and a common set of spaces Fs0, and if that Ker(Φε)(uε), considered (under mapping by this isometry) as a subset of Fs0 is finite-dimensional, with a limit asε→0,then, there exist a family of projectionsΠεonto Ker(Φε)(uε)that are uniformly bounded with respect to εin eachFsε, forεsufficiently small.

Indeed, by the above note (second paragraph of the present remark), there exists a bounded projectionΠ0onto the limit asε→0 of Ker(Φε)(uε). Denote byF˜:=KerΠ0the associated complementary subspace. DefiningΠεto be the projection alongF˜onto Ker(Φε)(uε), we find by the Hahn–Banach Theorem, compactness of the intersection of the unit ball with Ker(Φε)(uε), and continuity, thatΠεis bounded forεsufficiently small.

Références

Documents relatifs

Our main result is to show existence with sharp rates of decay and dis- tance from the Chapman–Enskog approximation of small-amplitude quasi- linear relaxation shocks in the

We prove an abstract Nash–Moser implicit function theorem with parameters which covers the applications to the existence of finite dimensional, differentiable, invariant tori

We study the global Cauchy problem for nonlinear Schrödinger equations with cubic interactions of derivative type in space dimension n 3.. The global existence of small

Frehse, Bellman Systems of Stochastic Differential Games with three Players in Optimal Control and Partial Differ- ential Equations, edited by J.L. Hildebrandt

This class K L contains almost all fundamental operators (we refer the reader to [32] for the classical case L = − ∆) related to the Schr¨odinger operator

Since the non-linearity in the right-hand side is a singular ab- sorption term near the boundary, a non-trivial weak solution may not be positive everywhere in Ω and compact

Boundary Harnack inequality and a priori estimates of singular solutions of quasilinear elliptic equations.. Marie-Françoise Bidaut-Veron, Rouba Borghol,

The main tool in the proof is Foun- tain Theorem, because the weak solutions of system (S) correspond to the critical points of the C 1 functional H..