• 沒有找到結果。

Propagation of smallness and size estimate in the second order elliptic equation with discontinuous complex Lipschitz conductivity

N/A
N/A
Protected

Academic year: 2022

Share "Propagation of smallness and size estimate in the second order elliptic equation with discontinuous complex Lipschitz conductivity"

Copied!
30
0
0

加載中.... (立即查看全文)

全文

(1)

Propagation of smallness and size estimate in the second order elliptic equation with discontinuous

complex Lipschitz conductivity

Elisa Francini

Sergio Vessella

Jenn-Nan Wang

Abstract

In this paper, we would like to derive three-ball inequalities and propagation of smallness for the complex second order elliptic equation with discontinuous Lipschitz coefficients. As an application of such estimates, we study the size estimate problem by one pair of Cauchy data on the boundary. The main ingre- dient in the derivation of three-ball inequalities and propagation of smallness is a local Carleman proved in our recent paper [FVW].

Keywords: Complex second order elliptic operators, discontinuous Lipschitz co- efficients, propagation of smallness, size estimate.

MSC2020: 35B60, 35R30.

1 Introduction

In our recent paper [FVW], we derived a Carleman estimate for the second order elliptic equation with piecewise complex-valued Lipschitz coefficients. The theme of this paper is to prove some interesting results based on the Carleman estimate obtained in [FVW]. The ultimate goal is to give upper and lower bounds of the size of the inclusion embedded inside of a conductive body with discontinuous complex conductivity by only one pair of boundary measurements. A typical application of this study is to estimate the size of a cancerous tumor inside an organ by the electric impedance tomography (EIT).

For the conductivity equation with piecewise real Lipschitz coefficients, the same size estimate problem was considered in [FLVW]. We want to point out that, in many real world problems, the case of complex-valued coefficients arises naturally.

The modeling of the current flows in biological tissues or the propagation of the

Universit`a di Firenze, Italy. Email: elisa.francini@unifi.it

Universit`a di Firenze, Italy. Email: sergio.vessella@unifi.it

National Taiwan University, Taiwan. Email: jnwang@math.ntu.edu.tw

(2)

electromagnetic waves in conductive media are typical examples. In these cases, the conductivities are complex-valued functions. On the other hand, in some situations, the conductivities are not continuous functions. In the human body, different organs have different conductivities. For instance, the conductivities of heart, liver, intestines are 0.70 (S/m), 0.10 (S/m), 0.03 (S/m), respectively. Therefore, to model the current flow in the human body, it is more reasonable to consider an anisotropic complex- valued conductivity with jump-type discontinuities [MPH].

In the size estimate problem studied in [FLVW], the essential tool is a three-region inequality which is obtained by applying the Carleman estimate for the second order elliptic equation with piecewise real Lipschitz coefficients derived in [DFLVW]. Since we have the similar Carleman estimate available for the case of piecewise complex- valued Lipschitz coefficients, we can proceed the method used in [FLVW] to prove the three-region inequality. In treating the size estimate problem, the three-region inequality is enough since one only needs to cross the interface once in propagating the information in the interior to the boundary. However, the three-region inequality is inconvenient in deriving the general propagation of smallness. Therefore, in this paper, we want to derive the usual three-ball inequality for the complex second order elliptic operator even when the coefficients are piecewise Lipschitz. We will follow the ideas outlined in [CW] where the three-ball inequality was proved for the real second order elliptic operator with piecewise Lipschitz coefficients. We then apply the three-ball inequality to derive a general propagation of smallness for the second order elliptic equation with piecewise complex Lipschitz coefficients.

We would like to raise an issue in the investigation of the size estimate problem when the background medium is complex valued. Following the method used in [ARS], an important step is to derive certain energy inequalities controlling the power gap. It was noted in [BFV] that when the current flows inside and outside of the inclusion obey the usual Ohm’s law (the relation between current and voltage is linear) and the imaginary part of the conductivity outside of the inclusion is a nonzero variable function, energy inequalities (5.9), (5.10) are not likely to hold. Precisely, in this case, an example with δW = 0, but |D| 6= 0, is constructed in [BFV]. On the other hand, if both conductivities inside and outside of the inclusion are complex constants or the conductivity outside of the inclusion is real valued, then energy inequalities (5.9), (5.10) were obtained in [BFV].

A key observation found in [CNW] is that if the current flow inside of the inclusion obeys certain nonlinear Ohm’s law, we can restore the energy inequalities (5.9), (5.10).

In particular, our size estimate result applies to the case of a non-chiral medium with a chiral inclusion having real valued chirality.

The paper is organized as follows. In Section 2, we introduce some notations and state the Carleman estimate proved in [FVW]. In Section 3, we plan to prove a three-region inequality across the interface based on the Carleman estimate given in Section 2. We then combine the classical three-ball inequality and the three-region inequality to derive a three-ball inequality in Section 4. There, we also prove the propagation of smallness. Finally, we study the size estimate problem in Section 5.

(3)

2 Notations and Carleman estimate

In this section, we will state the Carleman estimate proved in [FVW] where the interface is assumed to be flat. Since our Carleman estimate is local near any point at the interface, for a general C1,1 interface, it can be flatten by a suitable change of coordinates. Moreover, the transformed coefficients away from the interface remain Lipschitz. Define H± = χRn± where Rn± = {(x0, xn) ∈ Rn−1× R|xn ≷ 0} and χRn± is the characteristic function of Rn±. In places we will use equivalently the symbols ∂,

∇ and D = −i∇ to denote the gradient of a function and we will add the index x0 or xn to denote gradient in Rn−1 and the derivative with respect to xn respectively.

We further denote ∂` = ∂/∂x`, D` = −i∂`, and ∂ξ` = ∂/∂ξ`. Let u±∈ C(Rn). We define

u = H+u++ Hu=X

±

H±u±,

hereafter, we denote P

±a±= a++ a, and L(x, D)u := X

±

H±div(A±(x)∇u±),

where

A±(x) = {a±`j(x)}n`,j=1 = {a±`j(x0, xn)}n`,j=1, x0 ∈ Rn−1, xn ∈ R (2.1) is a Lipschitz symmetric matrix-valued function. Assume that

a±`j(x) = a±j`(x), ∀ `, j = 1, · · · , n, (2.2) and furthermore

a±`j(x) = M`j±(x) + iγN`j±(x), (2.3) where (M`j±) and (N`j±) are real-valued matrices and γ > 0. We further assume that there exists λ0 > 0 such that for all ξ ∈ Rn and x ∈ Rn we have

λ0|ξ|2 ≤ M±(x)ξ · ξ ≤ λ−10 |ξ|2 (2.4) and

λ0|ξ|2 ≤ N±(x)ξ · ξ ≤ λ−10 |ξ|2. (2.5) In the paper, we consider Lipschitz coefficients A±, i.e., there exists a constant M0 > 0 such that

|A±(x) − A±(y)| ≤ M0|x − y|. (2.6) To treat the transmission conditions, we write

h0(x0) := u+(x0, 0) − u(x0, 0), ∀ x0 ∈ Rn−1, (2.7) h1(x0) := A+(x0, 0)∇u+(x0, 0) · ν − A(x0, 0)∇u(x0, 0) · ν, ∀ x0 ∈ Rn−1, (2.8)

(4)

where ν = en.

Let us now introduce the weight function. Let ϕ be

ϕ(xn) =

( ϕ+(xn) := α+xn+ βx2n/2, xn ≥ 0,

ϕ(xn) := αxn+ βx2n/2, xn < 0, (2.9) where α+, α and β are positive numbers which will be determined later. In what follows we denote by ϕ+ and ϕ the restriction of the weight function ϕ to [0, +∞) and to (−∞, 0) respectively. We use similar notation for any other weight functions.

For any ε > 0 let

ψε(x) := ϕ(xn) −ε

2|x0|2, (2.10)

and let

φδ(x) := ψδ−1x), δ > 0. (2.11) For a function h ∈ L2(Rn), we define

ˆh(ξ0, xn) = Z

Rn−1

h(x0, xn)e−ix0·ξdx0, ξ0 ∈ Rn−1.

As usual we denote by H1/2(Rn−1) the space of the functions f ∈ L2(Rn−1) satisfying Z

Rn−1

0|| ˆf (ξ0)|20 < ∞, with the norm

kf k2H1/2(Rn−1)= Z

Rn−1

(1 + |ξ0|2)1/2| ˆf (ξ0)|20. (2.12) Moreover we define

[f ]1/2,Rn−1 =

Z

Rn−1

Z

Rn−1

|f (x) − f (y)|2

|x − y|n dydx

1/2

,

and recall that there is a positive constant C, depending only on n, such that C−1

Z

Rn−1

0|| ˆf (ξ0)|20 ≤ [f ]21/2,Rn−1 ≤ C Z

Rn−1

0|| ˆf (ξ0)|20,

so that the norm (2.12) is equivalent to the norm kf kL2(Rn−1)+ [f ]1/2,Rn−1. We use the letters C, C0, C1, · · · to denote constants. The value of the constants may change from line to line, but it is always greater than 1.

We will denote by Br0(x0) the (n − 1)-ball centered at x0 ∈ Rn−1 with radius r > 0.

Whenever x0 = 0 we denote Br0 = Br0(0). Likewise, we denote Br(x) be the n-ball centered at x ∈ Rn with radius r > 0 and Br = Br(0).

(5)

Theorem 2.1 Let A±(x) satisfy (2.1)-(2.6). There exist α+, α, β, δ0, r0, γ0, τ0, C depending on λ0, M0 such that if γ ≤ γ0, δ ≤ δ0 and τ ≥ τ0, then

X

± 2

X

k=0

τ3−2k Z

Rn±

|Dku±|2e2τ φδ,±(x0,xn)dx0dxn+X

± 1

X

k=0

τ3−2k Z

Rn−1

|Dku±(x0, 0)|2eδ(x0,0)dx0

+X

±

τ2[eτ φδ(·,0)u±(·, 0)]21/2,Rn−1+X

±

[D(eτ φδ,±u±)(·, 0)]21/2,Rn−1

≤C X

±

Z

Rn±

|L(x, D)(u±)|2e2τ φδ,±(x0,xn)dx0dxn+ [eτ φδ(·,0)h1]21/2,Rn−1

+[Dx0(eτ φδh0)(·, 0)]21/2,Rn−1 + τ3 Z

Rn−1

|h0|2e2τ φδ(x0,0)dx0+ τ Z

Rn−1

|h1|2e2τ φδ(x0,0)dx0

 . (2.13) where u = H+u++ Hu, u± ∈ C(Rn) and supp u ⊂ Bδr0

0 × [−δr0, δr0], and φδ is given by (2.11).

Remark 2.1 In view of the proof of Theorem 2.1 in [FVW], the coefficients α+, α

are required to satisfy

α+ α

≥ κ0 > 1, (2.14)

where κ0 is general constant depending on the values of A±(0) at the interface.

Remark 2.2 It is clear that (2.13) remains valid if can add lower order terms P

±H±(W · ∇u±+ V u±), where W, V are bounded functions, to the operator L.

That is, one can substitute L(x, D)u =X

±

H±div(A±(x)∇u±) +X

±

H±(W · ∇u±+ V u±) (2.15)

in (2.13).

3 Three-region inequalities

Based on the Carleman estimate given in Theorem 2.1, we will derive three-region inequalities across the interface xn = 0. Here we consider u = H+u++ Hu satis- fying

L(x, D)u = 0 in Rn, (3.1)

where L is given in (2.15) and

kW kL(Rn)+ kV kL(Rn)≤ λ−10 .

Now all coefficients α±, β, δ0, γ0, r0, τ0 have been determined in Theorem 2.1.

(6)

Theorem 3.1 Let u be a solution of (3.1) and A±(x) satisfy (2.1)-(2.6) with h0 = h1 = 0. Moreover, the constant γ in (2.3) satisfies γ ≤ γ0 with γ0 given in The- orem 2.1. Then there exist C and R, depending only on λ0, M0, n, such that if 0 < R1, R2 ≤ R, then

Z

U2

|u|2dx ≤ (eτ0R2 + CR−41 )

Z

U1

|u|2dx

2R1+3R2R2 Z

U3

|u|2dx

2R1+2R22R1+3R2

, (3.2) where

U1 =



z ≥ −4R2, R1

8a < xn< R1 a

 , U2 =



−R2 ≤ z ≤ R1

2a, xn < R1 8a

 , U3 =



z ≥ −4R2, xn < R1 a

 , a = α+/δ,

z(x) = αxn

δ +βx2n

2 − |x0|2

2δ , (3.3)

and any δ ≤ δ0.

xn

x0

z=−4R2

z=−R2

z=R12a

xn=R1a

xn=R18a

U1

U2

Figure 1: U1 and U2 are shown in pink and yellow, respectively. U3 is the region enclosed by black boundaries. Note that since z is hyperbolic, there are parts similar to U2 and U3 lying below xn< −αδ/β. Here we are only interested in the solution near xn = 0. Thus we consider the cut-off function relative to U2 and U3 as in the figure.

(7)

Proof. Here we adopt the proof given in [FLVW]. To apply the estimate (2.13), we need to ensure that u satisfies the support condition. Let r > 0 be chosen satisfying

r ≤ min



r20,13α

8β , 2δr0 19α+ 8β



. (3.4)

We then set

R = αr 16 . It follows from (3.4) that

R ≤ 13α2

128β. (3.5)

Given 0 < R1 < R2 ≤ R. Let ϑ1(t) ∈ C0(R) satisfy 0 ≤ ϑ1(t) ≤ 1 and

ϑ1(t) =

(1, t > −2R2, 0, t ≤ −3R2. Also, define ϑ2(t) ∈ C0(R) satisfying 0 ≤ ϑ2(t) ≤ 1 and

ϑ2(t) =





0, t ≥ R1 2a, 1, t < R1

4a.

Finally, we define ϑ(x) = ϑ(x0, xn) = ϑ1(z(x))ϑ2(xn), where z is defined by (3.3).

We now check the support condition for ϑ. From its definition, we can see that supp ϑ is contained in





z(x) = αxn

δ + βx2n

2 − |x0|2

2δ > −3R2, xn < R1

2a.

(3.6)

In view of the relation

α+ > α (see (2.14)) and a = α+ δ , we have that

R1 2a < δ

· R1 < δ α

· αr 16 < δr, i.e., xn< δr ≤ δr02 ≤ δr0. Next, we observe that

−3R2 > −3R = −3αr 16 > α

δ (−δr) + β

2(−δr)2,

(8)

which gives −δr0 < −δr < xn due to (3.4). Consequently, we verify that |xn| < δr <

δr0. One the other hand, from the first condition of (3.6) and (3.4), we see that

|x0|2

2δ < 3R2xn

δ +βx2n

2 ≤ 3αr 16 + α

δ · δr + β

2 · δ2r2

≤ 19α+ 8β

16 r ≤ 19α+ 8β

16 r02 ≤ δ 8r02, which gives |x0| < δr0/2.

Since h0 = 0, we have that

ϑ(x0, 0)u+(x0, 0) − ϑ(x0, 0)u(x0, 0) = 0, ∀ x0 ∈ Rn−1. (3.7) Applying (2.13) to ϑu and using (3.7) yields

X

± 2

X

|k|=0

τ3−2|k|

Z

Rn±

|Dk(ϑu±)|2e2τ φδ,±(x0,xn)dx0dxn

≤CX

±

Z

Rn±

|L(x, D)(ϑu±)|2e2τ φδ,±(x0,xn)dx0dxn

+ Cτ Z

Rn−1

|A+(x0, 0)∇(ϑu+(x0, 0)) · ν − A(x0, 0)∇(ϑu)(x0, 0) · ν|2e2τ φδ(x0,0)dx0 + C[eτ φδ(x0,0) A+(x0, 0)∇(ϑu+)(x0, 0) · ν − A(x0, 0)∇(ϑu)(x0, 0) · ν]21/2,Rn−1.

(3.8) We now observe that ∇ϑ1(z) = ϑ01(z)∇z = ϑ01(z)(−xδ0,αδ + βxδ2n) and it is nonzero only when

−3R2 < z < −2R2. Therefore, when xn = 0, we have

2R2 < |x0|2

2δ < 3R2. Thus, we can see that

|∇ϑ(x0, 0)|2 ≤ CR−22  6R2 δ +α2

δ2



≤ CR2−2. (3.9)

By h0(x0) = h1(x0) = 0, (3.9), and the easy estimate of [DFLVW, Proposition 4.2],

(9)

we can estimate τ

Z

Rn−1

|A+(x0, 0)∇(ϑu+(x0, 0)) · ν − A(x0, 0)∇(ϑu)(x0, 0) · ν|2e2τ φδ(x0,0)dx0 + [eτ φδ(x0,0) A+(x0, 0)∇(ϑu+)(x0, 0) · ν − A(x0, 0)∇(ϑu)(x0, 0) · ν]21/2,Rn−1

≤ CR−22 e−4τ R2

 τ

Z

{

4δR2≤|x0|≤ 6δR2}

|u+(x0, 0)|2dx0+ [u+(x0, 0)]21/2,{4δR

2≤|x0|≤ 6δR2}



+ Cτ2R2−3e−4τ R2 Z

{

4δR2≤|x0|≤ 6δR2}

|u+(x0, 0)|2dx0

≤ Cτ2R−32 e−4τ R2E,

(3.10) where

E = Z

{

4δR2≤|x0|≤ 6δR2}

|u+(x0, 0)|2dx0+ [u+(x0, 0)]21/2,{4δR

2≤|x0|≤ 6δR2}.

Writing out L(x, D)(ϑu±) and considering the set where ∇ϑ 6= 0, it is not hard to estimate

X

± 1

X

|k|=0

τ3−2|k|

Z

{−2R2≤z≤R12a, xn<R14a}

|Dku±|2e2τ φδ,±(x0,xn)dx0dxn

≤ CX

± 1

X

|k|=0

R2(|k|−2)2 Z

{−3R2≤z≤−2R2, xn<R12a}

|Dku±|2e2τ φδ,±(x0,xn)dx0dxn

+ C

1

X

|k|=0

R2(|k|−2)1 Z

{−3R2≤z,R14a<xn<R12a}

|Dku+|2e2τ φδ,+(x0,xn)dx0dxn

+ Cτ2R−32 e−4τ R2E

≤ CX

± 1

X

|k|=0

R2(|k|−2)2 e−4τ R2e(α+−α−)δ R14a Z

{−3R2≤z≤−2R2, xn<R14a}

|Dku±|2dx0dxn

+

1

X

|k|=0

R2(|k|−2)1 eα+δ R12ae2δ2β (R12a)2 Z

{z≥−3R2,R14a<xn<R12a}

|Dku+|2dx0dxn + Cτ2R−32 e−4τ R2E.

(3.11)

Let us recall U1 = {z ≥ −4R2, R8a1 < xn < Ra1}, U2 = {−R2 ≤ z ≤ R2a1, xn < R8a1}.

From (3.11) and interior estimates (Caccioppoli’s type inequality), we can derive that

(10)

τ3e−2τ R2 Z

U2

|u|2dx0dxn= τ3e−2τ R2 Z

{−R2≤z≤R12a, xn<R18a}

|u|2dx0dxn

≤ X

±

τ3 Z

{−2R2≤z≤R12a, xn<R14a}

|u±|2e2τ φδ,±(x0,xn)dx0dxn

≤ CX

± 1

X

|k|=0

R2(|k|−2)2 e−4τ R2e(α+−α−)δ R14a Z

{−3R2≤z≤−2R2, xn<R14a}

|Dku±|2dx0dxn

+

1

X

|k|=0

R2(|k|−2)1 eα+δ R12ae2δ2β (R12a)2 Z

{z≥−3R2,R14a<xn<R12a}

|Dku+|2dx0dxn + Cτ2R−32 e−4τ R2E

≤ CR−41 e−3τ R2 Z

{−4R2≤z≤−R2, xn<R1a }

|u|2dx0dxn+ Cτ2R−32 e−4τ R2E

+ CR−41 e(1+

βR1 4α2

)τ R1Z

{z≥−4R2,R18a<xn<R1a }

|u|2dx0dxn

≤CR−41

 e2τ R1

Z

U1

|u|2dx0dxn+ τ2e−3τ R2F

 ,

(3.12)

where

F = Z

U3

|u|2dx0dxn and we used the inequality βR21

< 1 in view of (3.5). Remark that we estimate E by F using the trace estimate and the interior estimate.

Dividing τ3e−2τ R2 on both sides of (3.12) gives Z

U2

|u|2dx ≤ CR−41



e2τ (R1+R2) Z

U1

|u|2dx0dxn+ e−τ R2F



. (3.13)

We now discuss two cases. If R

U1|u|2dx0dxn6= 0 and e0(R1+R2)

Z

U1

|u|2dx0dxn< e−τ0R2F, then we can choose a τ > τ0 so that

e2τ (R1+R2) Z

U1

|u|2dx0dxn= e−τ R2F.

With such τ , it follows from (3.13) that Z

U2

|u|2dx ≤ CR1−4e2τ (R1+R2) Z

U1

|u|2dx0dxn

= CR−41

Z

U1

|u|2dx0dxn

2R1+3R2R2

(F )2R1+2R22R1+3R2.

(3.14)

(11)

If R

U1|u|2dx0dxn = 0, then letting τ → ∞ in (3.13) we have R

U2|u|2dx0dxn = 0 as well. The three-regions inequality (3.2) obviously holds.

On the other hand, if

e−τ0R2F ≤ e0(R1+R2) Z

U1

|u|2dx0dxn, then we have

Z

U2

|u|2dx0dxn ≤ (F )2R1+3R2R2 (F )2R1+2R22R1+3R2

≤ exp (τ0R2)

Z

U1

|u|2dx0dxn

2R1+3R2R2

(F )2R1+2R22R1+3R2 .

(3.15)

Putting together (3.14), (3.15) implies Z

U2

|u|2dx0dxn≤ (exp (τ0R2) + CR1−4)

Z

U1

|u|2dx0dxn

2R1+3R2R2

(F )2R1+2R22R1+3R2 . (3.16)

2

4 Propagation of smallness

In this section, we will derive a general propagation of smallness for solutions satis- fying (3.1), L(x, D)u = 0 in Rn, using the ideas given in [CW]. For the region away from the interface, classical three-ball inequalities are shown to hold for the com- plex second order elliptic operators [CNW]. We will mainly focus on the inequalities across the interface. Let us first fix some notations. Assume that Ω ⊂ Rn is an open bounded domain with Lipschitz boundary and Σ ⊂ Ω is a C1,1 hypersurface. Fur- thermore, assume that Ω \ Σ only has two connected components, which we denote Ω±. Let A±(x) = (a`j(x)±)n`,j=1, W (x), V (x) be bounded measurable complex valued coefficients defined in Ω. We say that

ζ± := (A±, W, V ) ∈V (Ω±, λ0, M0, K1, K2) if A± satisfy (2.2)-(2.6) for x ∈ Ω± and W, V satisfy

kW kL(Ω) ≤ K1, kV kL(Ω) ≤ K2. We will use the notation Lζ to denote

Lζ(x, D)u =X

±

H±div(A±(x)∇u±) +X

±

H±(W · ∇u±+ V u±) in Ω.

Here, by an abuse of notation, we denote H± = χ±.

(12)

For an open set U ⊂ Rn and a number s > 0, we define Us = {x ∈ Rn : dist(x, U ) < s}, Us = {x ∈ U : dist(x, ∂U ) > s}, and

sU = {sx : x ∈ U }.

Definition 4.1 We say that Ω ∈ Ck,1, k ∈ N with constants ρ1, M1 if for any point P ∈ ∂Ω, after a rigid transformation, P = 0 and

Ω ∩ Γρ1,M1(0) = {(x0, xn) : x0 ∈ Rn−1, |x0| < ρ1, xn∈ R, xn > Φ(x0)}, where Φ is a Ck,1 function such that Φ(0) = 0, kΦkCα,1(Bρ1(0)) ≤ M1, and

Γρ1,M1(0) = {(x0, xn) : x0 ∈ Rn−1, |x0| < ρ1, |xn| ≤ M1}.

Throughout this paper, when saying that a domain is Ck,1, we will mean that it is Ck,1 with constants ρ1 and M1.

Definition 4.2 We say that Σ is C1,1 with constants ρ0, K0 if for any point P ∈ Σ, after a rigid transformation, P = 0 and

±∩ Cρ0,K0(0) = {(x0, xn) : x0 ∈ Rn−1, |x0| < ρ0, xn∈ R, xn≷ ψ(x)},

where ψ is a C1,1 function such that ψ(0) = 0, ∇x0ψ(0) = 0, kψkC1,1(Bρ0(0)) ≤ K0, and

Cρ0,K0(0) = {(x0, xn) : x0 ∈ Rn−1, |x0| < ρ0, |xn| ≤ 1

2K0ρ20}.

If Σ is as above, then we may ”flatten” the boundary around the point P (without loss of generality P = 0) via the local C1,1-diffeomeorphism

ΨP(x, y) = (x, y − ψ(x)).

With these assumptions, we will prove a propagation of smallness result as follows.

Theorem 4.1 Suppose u ∈ H1(Ω) solves

Lζu = 0 in Ω.

Then there exist γ0, depending on λ0, M0, and h0, depending on λ0, M0, K1, K2, ρ0, K0, such that if γ < γ0 and 0 < h < h0, r/2 > h, D ⊂ Ω is connected, open, and D \ Σ has two connected components, denoted by D±, such that Br(x0) ⊂ D, dist(D, ∂Ω) ≥ h, then

kukL2(D)≤ CkukδL2(Br(x0))kuk1−δL2(Ω), where

C = C1 |Ω|

hn

12

eC3h−s, δ ≥ τC2|Ω|hn ,

with s = s(λ0, K1, K2), C1, C2 > 0, τ ∈ (0, 1) depending on λ0, M0, K1, K2, ρ0, K0.

(13)

We would like to remark that the propagation of smallness in Theorem 4.1 is valid regardless the locations of D and Br0(x0), which may intersect the interface Σ.

The strategy of proving Theorem 4.1 consists two parts. When we are at one side of the interface, we can use the usual propagation of smallness for equations with Lipschitz complex coefficients based on [CNW]. When near the interface, we then use the three-region inequality derived above to propagate the smallness across the interface. The rest of this section is devoted to the proof of Theorem 4.1.

4.1 Propagation of smallness away from the interface

In this subsection, we want to derive a propagation of smallness for second order complex elliptic operators with Lipschitz leading coefficients. We consider ζ :=

(A, W, V ) ∈ V (U, λ0, M0, K1, K2), where U is an open bounded domain and the value of γ in (2.3) is irrelevant. Note that here A is Lipschitz without jumps in U . The following three-ball inequality was proved in [CNW, Theorem 3].

Proposition 4.1 Assume that ζ := (A, W, V ) ∈ V (U, λ0, M0, K1, K2) and u ∈ Hloc1 (Ω) solves Lζu = 0 in U . Then there exist positive constants R = R(n, λ0, M0) and s = s(λ0, K1, K2) such that if 0 < r0 < r1 < λ0r2/2 < √

λ0R/2 with Br2(x0) ⊂ U , then

kukL2(Br1(x0))≤ CkukτL2(Br0(x0))kuk1−τL2(Br2(x0)), (4.1) where C is explicitly given by

C = eC1(r−s0 −r2−s) with C1 > depending on λ0, M0, K1, K2, and

τ = (2r10)−s− r−s2

r0−s− r2−s = (2r1/r2λ0)−s− 1 (r0/r2)−s− 1 .

The proof of Proposition4.1 relies on the Carleman estimate derived in [CGT].

Having established the three-ball inequality (4.1), we can prove the following prop- agation of smallness based on the chain of balls argument in [ARRV, Theorem 5.1].

We will not repeat the argument here.

Proposition 4.2 Assume that ζ := (A, W, V ) ∈ V (U, λ0, M0, K1, K2) and u ∈ Hloc1 (Ω) solves Lζu = 0 in U . Let 0 < h < r/2 with r ≤ √

λ0R/2, D ⊂ U con- nected, open, and such that Br(x0) ⊂ D, dist(D, ∂U ) ≥ h. Then

kukL2(D) ≤ CkukδL2(Br(x0))kuk1−δL2(U ), where

C = C2 |U | hn

12

eC3h−s, δ ≥ τC4|U|hn , with C2, C3, C4 > 0 depending on λ0, M0, K1, K2.

(14)

4.2 Propagation of smallness – an intermediate result

Here we would like to prove an intermediate propagation of smallness result in which the small ball lies entirely on one side of the interface. Assume that D ⊂⊂ Ω is open and connected. Recall that we have assumed that Ω \ Σ and D \ Σ both have two connected components, denoted by Ω± and D±, respectively. Let ωΣ be the surface measure induced on Σ by the Lebesgue measure on Rn. We will consider coefficients

ζ = (A±, W, V ) ∈V (Ω±, λ0, M0, K1, K2).

We can now prove the following propagation of smallness result.

Theorem 4.2 Suppose u ∈ H1(Ω) solves Lζu = 0 in Ω. Then there exist γ0, de- pending on λ0, M0, and h0, depending on λ0, M0, K1, K2, ρ0, K0, such that if γ < γ0 and 0 < h ≤ h0, h < r/2, Br(x0) ⊂ D+, and dist(D, ∂Ω) ≥ h, then

kukL2(D)≤ CkukδL2(Br(x0))kuk1−δL2(Ω), where

C = C1 |Ω|

hn

"

1 + ωΣ(Σ ∩ Ω) hn−1

12#

eC3h−s, δ ≥ τC2|Ω|hn , with C1, C2, C3 > 0, τ ∈ (0, 1) depending on λ0, M0, K1, K2, ρ0, K0.

The difficult part of proving Theorem4.2 is to obtain L2 estimates of the solution in a neighborhood of Σ. We will use Theorem3.1to overcome this difficulty. However, we cannot apply Theorem3.1directly. The family of regions given in Theorem3.1has one serious drawback. If we choose the parameters R1 = θ ¯R1, R2 = θ ¯R2, θ ∈ (0, 1), the vertical sizes of the regions would scale like θ, while their horizontal sizes would scale like θ12. Using just these two parameters in the proof would then lead to constants in the propagation of smallness inequality (i.e. the constants C and δ in Theorem 4.1) that depend on the geometry of Ω, D, and Br(x0) in a way that is not invariant under a rescaling of these sets. Therefore, we will study how the three-region inequality (3.2) behaves under scaling.

Let us first introduce the scaled coefficients ζ =˜  ˜A±, ˜W , ˜V

∈V (Rn±, λ0, M0, K1, K2).

For 0 < θ ≤ 1, let Lθζ˜(·, D)v =X

±

H±div( ˜A±(θ·)∇v±) +X

±

H±



θ ˜W (θ·) · ∇v±+ θ2V (θ·)v˜ ±

 .

Note that if ˜ζ = ˜A±, ˜W , ˜V



∈V (Rn±, λ0, M0, K1, K2), then

 ˜A±(θ·), θ ˜W (θ·), θ2V (θ·)˜ 

∈V (Rn±, λ0, θM0, θK1, θ2K2).

(15)

It is also clear that if Lζ˜u = 0 in Ω, then uθ(x) = θ−2u(θx) solves Lθζ˜uθ = 0 in θ−1Ω.

Moreover, if U ⊂ θ−1Ω, we have Z

θU

|u(x)|2dx = θn+4 Z

U

|uθ(y)|2dy.

We therefore obtain, by scaling, the following corollary to Theorem 3.1.

Proposition 4.3 Assume that the assumptions in Theorem3.1hold. Let 0 < R1, R2 ≤ R, θ ∈ (0, 1], and

L˜γu = 0 in θU3, then

Z

θU2

|u|2 ≤ (eτ0R2 + CR−41 )

Z

θU1

|u|2

2R1+3R2R2 Z

θU3

|u|2

2R1+2R22R1+3R2 .

In order to adapt that result to the possibly curved surface Σ, we need to first consider how the three regions transform under a local boundary flattening diffeo- morphism ΨP. Pick a point P ∈ Σ and set P = 0 without loss of generality. Let (x0, xn) ∈ Cρ0,K0(0). We will try to determine when (x0, xn) ∈ Ψ−1P (θU2). To this end, we introduce the notation

y0 = x0, yn= xn− ψ(x0).

It is clear that (x0, xn) ∈ Ψ−1P (θU2) if and only if θ−1(y0, yn) ∈ U2. We denote η(x0) = ψ(x0)

|x0|2 ,

which is a bounded function due to the regularity assumption of Σ. It was proved in [CW, Lemma 3.1] that if

r < θ min δR1

6aα,2δR2 , R1

12a, θ−1ρ0, ρ1, ρ2, ρ3



, (4.2)

then ΨP(Br(P )) ⊂ θU2, where













ρ1 = αδ δ + β,

ρ2 is chosen such that 2kηkρ2+ kηk2ρ22 < 1

2, kηk = kηkL(B0

ρ0(0)), ρ3 = 2αδ

β .

In [CW, Lemma 3.2], the following relation was established.

(16)

Lemma 4.1 Ψ−1P (θU3) is contained in a ball of radius

θ

(1 + 2kηk2) 2α

a R1+ 8δR2

 + 1

a2



2 + (1 + 2kηk2)β δ



R21+ 128δ2R22 h

α+ q

α2− 8βR2

i2

1/2

centered at P .

Finally, we need to estimate the distance from Ψ−1P (θU1) to Σ ∩ Cρ0,K0. It was proved in [CW, Lemma 3.3] that

dist(Ψ−1P (θU1), Σ) > θR1

16a. (4.3)

We are now ready to prove Theorem 4.2. We will follow the arguments used in the proof of Theorem 3.1 in [CW]. Since we have slightly different constants here, we provide the proof for the sake of completeness.

Proof of Theorem 4.2. By the assumption, we may take D to be the set D = {x ∈ Ω : dist(x, ∂Ω) > h}.

We want to point out that even though the choice of α± in Theorem 2.1 depends on A±(P ) for P ∈ Σ, we can choose a pair of α± such that Carleman estimate (2.13) holds near all P ∈ Σ in view of the regularity assumptions of A± and Σ.

Consequently, we can pick R1, R2 so that we can apply Proposition4.3 at any point P ∈ Σ ∩ D. By Lemma 4.1, there is a constant d > 0, independent of P , such that Ψ−1P (θU3) ⊂ Bθd(P ). We then choose θ such that θd = h2, which implies Ψ−1P (θU3) ⊂ Ω for any P ∈ Σ ∩ D. Of course, this choice is not possible if h is too large. Therefore, we need to set h0 small enough, depending on ρ0, K0, λ0, M0, K1, K2.

With this choice of parameters, by (4.3), there is a constant 0 < µ < 1, also independent on P , so that

dist(Ψ−1P (θU1), Σ) > µh.

Note that, depending on the geometry of Σ, we again need to set h0 and R small enough so that Ψ−1P (θU1) ∩ Σµh= ∅, for any P ∈ Σ ∩ D.

It follows from (4.2) that there exists a constant ν > 0, and without loss of generality ν < µ < 1, such that B5νh(P ) ⊂ Ψ−1P (θU2). By Vitali’s covering lemma, there exist finitely many P1, . . . , PN ∈ Σ ∩ D so that

Σνh∩ D ⊂

N

[

j=1

Ψ−1P

j(θU2), (4.4)

and the balls Bνh(Pj) are pairwise disjoint. By this last property, since for small h we have ωΣνh∩ D) ∼ νhωΣ(Σ ∩ D), it follows that there is a constant C such that

N ≤ CωΣ(Σ ∩ D)

hn−1 ≤ CωΣ(Σ ∩ Ω)

hn−1 . (4.5)

參考文獻

相關文件

Based on [BL], by checking the strong pseudoconvexity and the transmission conditions in a neighborhood of a fixed point at the interface, we can derive a Car- leman estimate for

In this paper we prove a Carleman estimate for second order elliptic equa- tions with a general anisotropic Lipschitz coefficients having a jump at an interface.. Our approach does

Robinson Crusoe is an Englishman from the 1) t_______ of York in the seventeenth century, the youngest son of a merchant of German origin. This trip is financially successful,

In this paper, we have studied a neural network approach for solving general nonlinear convex programs with second-order cone constraints.. The proposed neural network is based on

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

Section 3 is devoted to developing proximal point method to solve the monotone second-order cone complementarity problem with a practical approximation criterion based on a new

Although we have obtained the global and superlinear convergence properties of Algorithm 3.1 under mild conditions, this does not mean that Algorithm 3.1 is practi- cally efficient,