• Keine Ergebnisse gefunden

5.2 Solvability of the Microscopic Equations

5.2.4 Microscopic Coupling

5.2 Microscopic Solvability 83

84 5 SOLVABILITY OF THE TWO SCALE MODEL Consider in the following a possibly reduced time interval Iτ = [0, τ] with 0 < τ ≤ T and set

Mτ,α :=n

φ∈Cα(Iτ, Cper2 (Y))

φ(t, y)>0 ∀t ∈Iτ, ∀y ∈Yo . Define

BR,τ = n

(φ, cs)∈

Cα(Iτ, Cper2 (Y))2

kφkCα(Iτ,C2(Y))+kcskCα(Iτ,C2(Y)) ≤R, (φ, cs)(0,·) = (φi ni, cs,i ni)o , for R > max{kφi nikC2(Y),kcs,i nikC2(Y)}. BR,τ is a nonempty and closed (with re-spect to the Cα(Iτ, C2(Y))-norm) set. Note, that BR,τ is in general not a subset of Mτ,α×Cα(Iτ, Cper2 (Y)), sinceφ(t, y)>0 is not necessarily fulfilled.BR,τ complies with that only for certain choices of R and τ:

Proposition 5.22 (Well-definedness of S on BR,τ). Suppose φi ni(y)>0 for all y ∈Y. For any R >max{kφi nikC2(Y),kcs,i nikC2(Y)} there exist a time τ1 >0, depending on R, such that

BR,τ1 ⊂Mτ1×Cα(Iτ1, Cper2 (Y)).

Proof. It is to show that (φ, cs) ∈ BR,τ1 implies φ(t, y) > 0 for all y ∈ Y and t ≤ τ1, with a suitable τ1>0.

Consider an arbitrary but fixed R > max{kφi nikC2(Y),kcs,i nikC2(Y)} and suppose (φ, cs)∈BR,τ. Then, φ is α–Hölder continuous in time with Hölder constant ≤ R.

If φi ni(y) > 0 for all y ∈ Y, then there exists d := miny∈Y φi ni(y) > 0, because Y is compact. So,

|φ(t, y)−φi ni(y)| ≤Rtα,

and thus φ(t, y)>0 for all y ∈Y and t ≤τ1:= 2Rd 1/α . Proposition 5.22 ensures that the operator

S: BR,τ1 →[Cper1,2+2α(I×Y)]2

is well defined. Furthermore, there is a configuration of R and τ, such that S maps BR,τ into itself:

Proposition 5.23 (Self-mapping). Suppose φi ni(y) > 0 for all y ∈ Y. There exist postive numbers R0>0 and τ2 >0 such that

S: BR02 →BR02.

R0 and τ2 depend on the macroscopic coupling data, the initial data and the boundary data for the Stokes system and the elasticity equation.

5.2 Microscopic Solvability 85 Proof. Suppose 0 < τ ≤ τ1, with τ1 from Proposition 5.22, and ( ˜φ,c˜s)∈BR,τ. Set (φ, cs) =S( ˜φ,c˜s). By construction of S it is (φ, cs)(0,·) = (φi ni, cs,i ni).

It remains to show thatkφkCα(Iτ,C2(Y))+kcskCα(Iτ,C2(Y)) ≤R: The a priori estimates for the single parts of the problem, see Theorems 5.8, 5.14 and 5.19, imply that

kφkC1,2+2α(Iτ×Y)+kcskC1,2+2α(Iτ×Y)

≤c(κ)

1 +k˜csk2C(Iτ,C2(Y))+kCVk2C(Iτ)+k∇xVk2C(Iτ) +kPk2C(Iτ)+kbk2

C(Iτ,Wr2−1/r(Y×{0}))+kφi nikC2+2α(Y)+kcs,i nikC2+2α(Y)

, (5.76) where κ is an upper bound for kφk˜ C(Iτ,C2(Y)). For any τ ≤ τ1, κ can be choosen in-dependently of φ˜ and R: Due to φ˜ ∈ Cα(Iτ, Cper2 (Y)) and τ ≤ τ1 = 2Rd 1/α

, with d := miny∈Y φi ni(y), it is for 0≤t ≤τ

kφ(t, y)˜ −φi ni(y)kC2(Y) ≤Rtα ≤ d 2, and thus

sup

t∈Iτ

kφ(t)k˜ C2(Y) ≤ kφi nikC2(Y)+ d 2 =:κ.

So, the constantc in (5.76) can be choosen independently of φ˜and R, and (5.76) can be written as

kφkC1,2+2α(Iτ×Y)+kcskC1,2+2α(Iτ×Y)

≤c1(CV, V, P, b, φi ni, cs,i ni) +c2k˜csk2C(Iτ,C2(Y)), (5.77) where

c1(CV, V, P, b, φi ni, cs,i ni) =c

1 +kCVk2C(I)+k∇xVk2C(I)+kPk2C(I)+kφi nikC2+2α(Y)

+kcs,i nikC2+2α(Y)+kbk2

C(I,Wr2−1/r(Y×{0}))

. Next, note that

kc˜skC(Iτ,C2(Y)) = max

t∈Iτ kc˜s(t)−cs,i ni+cs,i nikC2(Y) ≤Rτα+kcs,i nikC2(Y). Consequently, (5.77) becomes

kφkC1,2+2α(Iτ×Y)+kcskC1,2+2α(Iτ×Y) ≤c˜1(CV, V, P, b, φi ni, cs,i ni) + ˜c2R2τ, (5.78) with

˜

c1(CV, V, P, b, φi ni, cs,i ni) =c1(CV, V, P, b, φi ni, cs,i ni) + ˜c2kcs,i nik2C2(Y).

86 5 SOLVABILITY OF THE TWO SCALE MODEL The continuous embedding

C1,2+2α(Iτ ×Y),→Cα(Iτ, C2(Y)),

see Lemma 2.12, implies together with (5.78) that

kφkCα(Iτ,C2(Y))+kcskCα(Iτ,C2(Y)) ≤c3 kφkC1,2+2α(Iτ×Y)+kcskC1,2+2α(Iτ×Y)

≤c31(CV, V, P, b, φi ni, cs,i ni) +c32R2τ. (5.79) Choose now R0 := 2˜c1c3 and τ2 >0 such thatc˜2c3R0τ212, i.e.τ2≤(2˜c2c3R0)1 . It follows from (5.79) that

kφkCα(Iτ2,C2(Y))+kcskCα(Iτ2,C2(Y)) ≤R0.

Finally, there is a choice of R and τ such that S: BR,τ →BR,τ

is a strict contraction:

Proposition 5.24 (Contraction). Consider R0 from Proposition 5.23. There exists a number τ3>0 such that the operator

S: BR03 →BR03 is a strict contraction.

Proof. Suppose 0 < τ ≤ τ2, with τ2 from Proposition 5.23, and ( ˜φi,c˜s,i)∈BR0, i = 1,2. Set(φi, cs,i) = S( ˜φi,c˜s,i). The continuity estimates of Lemmata 5.9, 5.15 and 5.21 (with C1V =C2V) imply that

1−φ2kC1,2+2α(Iτ×Y)+kcs,1−cs,2kC1,2+2α(Iτ×Y)

≤c(κ)kuˆ1+ ˆu2kC(I,W2

r( ˆQs)) kφ˜1−φ˜2kC(Iτ,C2(Y))+k˜cs,1−c˜s,2kC(Iτ,C2(Y))

, where uˆi = Sel asti c ◦ SStok es( ˜φi,c˜s,i) and κ ≥ max{kφ˜1kC(Iτ,C2(Y)),kφ˜2kC(Iτ,C2(Y))}. As seen in the proof of Proposition 5.23, κ can be choosen independently of φ˜i ∈BR0, if τ ≤τ1, with τ1 from Proposition 5.22. This is satisfied here. Furthermore, the a priori estimates of Theorems 5.8 and 5.14 yield together with ( ˜φi,c˜s,i)∈BR0

kuˆ1+ ˆu2kC(Iτ,W2

r( ˆQs)) ≤c1(κ, CV,∇V, P, b)

+c2(κ) k˜cs,1kC(Iτ,C2(Y))+k˜cs,2kC(Iτ,C2(Y))

≤c1(κ, CV,∇V, P, b) + 2c2(κ)R0

≤c(φi ni, cs,i ni, CV,∇V, P, b, R0).

5.2 Microscopic Solvability 87 This leads to

1−φ2kC1,2+2α(Iτ×Y)+kcs,1−cs,2kC1,2+2α(Iτ×Y)

≤c kφ˜1−φ˜2kC(Iτ,C2(Y))+k˜cs,1−c˜s,2kC(Iτ,C2(Y)) ,

with a constant c only depending on initial, boundary and macroscopic coupling data.

By construction, it is ( ˜φ1−φ˜2)(0, y) = (˜cs,1−c˜s,2)(0, y) = 0, and since φ˜i and c˜s,i belong toCα(Iτ, C2(Y)), it follows

kφ˜1−φ˜2kC(Iτ,C2(Y)) ≤ταkφ˜1−φ˜2kCα(Iτ,C2(Y)), kc˜s,1−c˜s,2kC(Iτ,C2(Y)) ≤ταk˜cs,1−c˜s,2kCα(Iτ,C2(Y)), and consequently

1−φ2kC1,2+2α(Iτ×Y)+kcs,1−cs,2kC1,2+2α(Iτ×Y)

≤c τα kφ˜1−φ˜2kCα(Iτ,C2(Y))+k˜cs,1−c˜s,2kCα(Iτ,C2(Y))

. The continuous embedding

C1,2+2α(Iτ ×Y),→Cα(Iτ, C2(Y)), see Lemma 2.12, implies

1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y))

≤c τ˜ α kφ˜1−φ˜2kCα(Iτ,C2(Y))+k˜cs,1−c˜s,2kCα(Iτ,C2(Y))

. Choose now τ3 such that c τ˜ 3α <1. This finishes the proof.

So at last, everything is prepared to prove the solvability of the coupled microscopic problem as the most important result in section 5.2. It is formulated in the following theorem:

Theorem 5.25(Existence and uniqueness of a solution of the coupled microscopic prob-lem). SupposeCV(x ,·),∇xV(x ,·), P(x ,·)∈C(Iτ3)forx ∈S0, withτ3from Proposition 5.24. Assume furthermore that

b(x ,·,·)∈C(Iτ3, Wr,per2−1/r(Y × {0})), φi ni(x ,·), cs,i ni(x ,·)∈Cper2+2α(Y),

withφi ni(x , y)>0 for all y ∈Y. Then, there exists a unique solution(φ, cs, v , p, u)(x) of (3.33) – (3.42) with

φ(x), cs(x)∈Cper1,2+2α(Iτ3 ×Y), ˆ

v(x)∈C(Iτ3,[Wr,per2 ( ˆQl K)]3), ˆ

p(x)∈C(Iτ3, Wr,per1 ( ˆQl K)), ˆ

u(x)∈C(Iτ3,[Wr,per2 ( ˆQs)]3),

88 5 SOLVABILITY OF THE TWO SCALE MODEL for some 0< α < 12, r > 1−2α6 . This solution satisfies the a priori estimate

kφ(x)kC1,2+2α(Iτ

3×Y)+kcs(x)kC1,2+2α(Iτ

3×Y)+kˆv(x)kC(I

τ3,Wr,per2 ( ˆQl K))

+kˆp(x)kC(Iτ

3,Wr,per1 ( ˆQl K))+ku(xˆ )kC(Iτ

3,Wr,per2 ( ˆQs))

≤c

1 +kCV(x)k2C(Iτ

3)+k∇xV(x)k2C(Iτ

3)+kP(x)k2C(Iτ

3)

+kb(x)k2

C(Iτ3,Wr2−1/r(Y×{0}))+kφi ni(x)kC2+2α(Y)+kcs,i ni(x)k2C2+2α(Y)

.

(5.80)

Proof. The assumptions for Banach’s fixed point theorem, see 2.18, on the operator S: BR03 →BR03

are fulfilled and so, there exists a unique fixed point in BR03. Any fixed point(φ, cs) of S, together with

(ˆv ,p) =ˆ SStokes(φ, cs), uˆ=Selastic(ˆv ,p, φ)ˆ

solves (3.33) – (3.42). Due to Theorems 5.8, 5.14 and 5.19, it is φ(x), cs(x) ∈Cper1,2+2α(Iτ3×Y).

As seen in Proposition 5.23, (φ, cs) satisfy kφkC1,2+2α(Iτ3×Y)+kcskC1,2+2α(Iτ3×Y) ≤c R0,

and by the definition of R0 in the proof of Proposition 5.23 and the estimates of Lemmata 5.8 and 5.14 it follows (5.80).

It remains to show that the just found solution is unique not only in BR03, but also in M×C(Iτ3, Cper2 (Y)). Suppose therefore that(φi, cs,i)∈M×C(Iτ3, Cper2 (Y)),i = 1,2, are fixed points of S. Note, that(φi, cs,i) also belong toCper1,2+2α(Iτ3×Y), due to Theorems 5.8, 5.14 and 5.19, and thus to Cα(Iτ3, Cper2 (Y)). Since (φi, cs,i) satisfy the same initial condition, it isk(φ1−φ2)(0)kC2(Y) =k(cs,1−cs,2)(0)kC2(Y) = 0. As seen in Proposition 5.24, the estimate

1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y))

≤c(κ)ταkuˆ1+ ˆu2kC(Iτ,W2

r( ˆQs))1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y))

, is satisfied for any τ ∈ Iτ3, where κ ≥ max{kφ1kC(Iτ

3,C2(Y)),kφ2kC(Iτ

3,C2(Y))}. It follows that k(φ1−φ2)(t)kC2(Y) =k(cs,1−cs,2)(t)kC2(Y) = 0 for t ∈[0, τ], if

c(κ)ταkuˆ1+ ˆu2kC(I

τ,Wr2( ˆQs)) <1.

Repeat the argument to show thatk(φ1−φ2)(t)kC2(Y) =k(cs,1−cs,2)(t)kC2(Y)= 0 not only on [0, τ]: If k(φ1−φ2)(t0)kC2(Y) = k(cs,1−cs,2)(t0)kC2(Y) = 0 for some t0 ∈Iτ3,

5.2 Microscopic Solvability 89 then it isk(φ1−φ2)(t)kC2(Y) =k(cs,1−cs,2)(t)kC2(Y) = 0 for allt ∈[t0, t0+τ(t0)]∩Iτ3. This shows that the set

I0 :=

t ∈Iτ3

k(φ1−φ2)(t)kC2(Y) =k(cs,1−cs,2)(t)kC2(Y) = 0

is an open subset of Iτ3, and it is not empty because 0 ∈ I0. But since t 7→ k(φ, cs)(t)kC2(Y) is continuous, I0 is also closed in Iτ3 and therefore I0 = Iτ3. This proves uniqueness of the solution of (3.33) – (3.42).

With the statement of Theorem 5.25, the main goal concerning the analysis for the microscopic problem is reached. Until here, a macroscopic pointx ∈S0 was fixed. This sections ends with necessary preparations for the micro-macro-coupling: Investigation of the regularity with respect to x ∈ S0 and continuity with respect to the coupling data. The answers are formulated in the following Lemmata:

Lemma 5.26 (Regularity with respect to x ∈ S0). Suppose that CV, ∇xV, P, b, φi ni andcs,i ni depend continuously onx ∈S0. Then, the solution of (3.33) – (3.42) depends continuously on x ∈S0 and

kφkC(S0,C1,2+2α(Iτ

3×Y))+kcskC(S0,C1,2+2α(Iτ

3×Y))+kˆvkC(I

τ3×S0,Wr,per2 ( ˆQl K))

+kpkˆ C(I

τ3×S0,Wr,per1 ( ˆQl K))+kukˆ C(I

τ3×S0,Wr,per2 ( ˆQs))

≤c

1 +kCVk2C(Iτ

3×S0)+kbk2

C(Iτ3×S0,Wr2−1/r(Y×{0})) +k∇xVk2C(Iτ

3×S0)

+kPk2C(Iτ

3×S0)+kφi nikC(S0,C2+2α(Y))+kcs,i nik2C(S

0,C2+2α(Y))

.

(5.81)

Proof. Suppose for i = 1,2 points xi ∈ S0 and set CiV = CV(xi), φi = φ(xi) and cs,i =cs(xi). Analogously to the proof of Proposition 5.24 it holds for τ ∈Iτ3

1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y))

≤c τα1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y))

+kC1V−C2VkC(Iτ)

, (5.82) with a constantc >0 depending only on the initial, boundary and macroscopic coupling data. As long as c τα <1, it follows

1−φ2kCα(Iτ,C2(Y))+kcs,1−cs,2kCα(Iτ,C2(Y)) ≤ckC1V −C2VkC(Iτ).

Repeating these arguments, starting with arbitrary t0 ∈ Iτ3 as initial time, leads to an analogous estimate as (5.82) on the time interval [t0, t0+ τ] ∩Iτ3, with a constant

˜

c depending on kφi(t0)kC2+2α(Y) and kcs,i(t0)kC2+2α(Y) instead of kφi(0)kC2+2α(Y) and kcs,i(0)kC2+2α(Y). Thanks to the a priori estimate (5.80), the mentioned constant c˜can in fact be choosen independently oft0, such that

1−φ2kCα([t0,t0],C2(Y))+kcs,1−cs,2kCα([t0,t0],C2(Y)) ≤ckC1V −C2VkC([t0,t0+τ]), as long ast0+τ ≤τ3 and c τ <˜ 1. This proves

kφ(x1)−φ(x2)kCα(Iτ3,C2(Y))+kcs(x1)−cs(x2)kCα(Iτ3,C2(Y)) ≤ckCV(x1)−CV(x2)kC(Iτ3),

90 5 SOLVABILITY OF THE TWO SCALE MODEL with a constantc >0depending only on the initial, boundary and macroscopic coupling data. The right hand side tends to zero for |x1−x2| →0, sinceCV is continuous with respect to x. It follows that φ, cs and, due to Lemmata 5.9, 5.15, also vˆ, pˆand uˆare continuous with respect to x.

Take the maximum with respect to x ∈S0 on both sides of the a priori estimate (5.80) to prove (5.81).

Lemma 5.27 (Continuity with respect to the coupling data). Suppose C1V, C2V ∈C(Iτ3×S0) and denote by φi, cs,i,uˆi,vˆi and pˆi, i = 1,2, the corre-sponding solutions of the microscopic problem (3.33) – (3.42). These solutions depend locally Lipschitz continuous on C1V and C2V, i.e. if kCiVkC(Iτ3×S0) ≤ R, for some R > 0, then

1−φ2kC(S0,C1,2+2α(Iτ

3×Y))+kcs,1−cs,2kC(S0,C1,2+2α(Iτ

3×Y))

+kˆv1−vˆ2kC(I

τ3×S0,Wr,per2 ( ˆQl K))+kˆp1−pˆ2kC(I

τ3×S0,Wr,per1 ( ˆQl K))

+kuˆ1−uˆ2kC(I

τ3×S0,Wr,per2 ( ˆQs))

≤ckC1V −C2VkC(Iτ3×S0),

(5.83)

with a constant c >0 depending on R.

Proof. Consider first fixedx ∈S0. Analogously to the proof of Lemma 5.26 it holds kφ1(x)−φ2(x)kCα(Iτ3,C2(Y))+kcs,1(x)−cs,2(x)kCα(Iτ3,C2(Y)) ≤ckC1V(x)−C2V(x)kC(Iτ3), with a constantc >0depending only on the initial, boundary and macroscopic coupling data. The continuity estimates of Lemmata 5.9 and 5.15 then imply

1(x)−φ2(x)kC1,2+2α(Iτ

3×Y)+kcs,1(x)−cs,2(x)kC1,2+2α(Iτ

3×Y)

+kˆv1(x)−vˆ2(x)kC(I

τ3,Wr,per2 ( ˆQl K))+kpˆ1(x)−pˆ2(x)kC(I

τ3,Wr,per1 ( ˆQl K))

+kuˆ1(x)−uˆ2(x)kC(I

τ3,Wr,per2 ( ˆQs))

≤ckC1V(x)−C2V(x)kC(Iτ

3).

Taking the maximum with respect to x ∈S0 proves (5.83).

Remark 5.28. The proofs of the last two Lemmata work also with less reg-ularity assumptions on CV as for example CV ∈ L2(S0, C(Iτ3)) and then lead to φ∈L2(S0, C1,2+2α(Iτ3 ×Y)) etc. But note, that the spaces L2(S0, C(Iτ3)) and C(Iτ3, L2(S0))do not coincide, and it is not clear, how to prove CV ∈L2(S0, C(Iτ3))as a solution of the macroscopic problem.

The only microscopic quantity, which occurs in the macroscopic problem as coupling datum, is the microscopic mean value c¯s. Therefore, the following Lemma, which is an obvious consequence of the Lemmata 5.26 and 5.27, is stated explicity:

5.2 Microscopic Solvability 91 Lemma 5.29 (On the microscopic mean value c¯s). It holds

kc¯skC1(Iτ3,C(S0)) ≤c

1 +kCVk2C(Iτ

3×S0)+k∇xVk2C(Iτ

3×S0)+kPk2C(Iτ

3×S0)

+kbk2

C(Iτ3×S0,Wr2−1/r(Y×{0}))+kφi nikC(S0,C2+2α(Y))

+kcs,i nik2C(S

0,C2+2α(Y))

, and

kc¯s,1−c¯s,2kC1(Iτ3,C(S0)) ≤ckC1V −C2VkC(Iτ3×S0), with the same constantc >0 as in Lemma 5.27.

Proof. Note that Y does not depend on t and thus

ts(x , t) =∂t

Z

Y

cs(x , t, y) dy = Z

Y

tcs(x , t, y) dy =∂tcs(x , t).

The statements follow from

|f¯(x , t)|= Z

Y

f(x , t, y) dy

≤ kf(x , t)kC(Y)

Z

Y

1 dy , f ∈ {cs, ∂tcs}, and |Y|= 1 and estimates (5.81) and (5.83) respectively.

92 5 SOLVABILITY OF THE TWO SCALE MODEL