• 沒有找到結果。

2 Statement of the result

N/A
N/A
Protected

Academic year: 2022

Share "2 Statement of the result"

Copied!
17
0
0

加載中.... (立即查看全文)

全文

(1)

Uniqueness for the two dimensional Calder´ on’s problem with unbounded conductivites

C˘ at˘ alin I. Cˆ arstea

Jenn-Nan Wang

Abstract

In this work we consider the Calder´on problem in two dimensions with conductivity γ ∈ W1,2(Ω). This condition allows for the conductivity to be unbounded. We prove a uniqueness result when ||∇ log γ||L2

is bounded by a fixed constant depending on the domain Ω.

1 Introduction

The inverse conductivity problem, first discussed by Calder´on in [8], consists of determining the conductivity of the interior of an object from measurements of electrical potential and current taken on the boundary.

One aspect of this problem is weather or not two different conductivity functions might give rise to the same set of boundary measurements. This question has been considered for different spatial dimensions and under various assumptions on the regularity of the conductivity function. An early result in dimension greater than two and for smooth conductivities was obtained by Sylvester and Uhlmann in their seminal paper [18].

Recently, Haberman and Tataru have shown in [12], also in dimensions higher than two, that uniqueness holds for Lipschitz conductivities γ such that ||∇ log γ||L is small. In [11], Haberman has further improved on this result in dimensions 3 ≤ n ≤ 6. In dimensions n = 3, 4 he proves uniqueness for conductivities in W1,n. We want to point out that in [11] the conductivity γ satisfies c ≤ γ ≤ c−1, in particular, γ ∈ L. In dimension two, Nachman proved in [14] uniqueness for conductivities in W2,p for some p > 1. Brown and Uhlmann in [7] established the uniqueness for conductivities in W1,pwith p > 2. In [2], Astala and P¨aiv¨arinta proved the uniqueness for conductivities in L. In the recent paper [3], Astala, Lassas, and P¨aiv¨arinta have proved a uniqueness result (see their Theorem 1.9) that also applies to some unbounded conductivities.

In this paper we will study the two-dimensional Calder´on’s problem in a different class of singular con- ductivities. Precisely, we prove that uniqueness holds for the strictly positive conductivity γ ∈ W1,2, with and additional constraint that ||∇ log γ||L2(Ω) < C with C = C(Ω). It is important to emphasize that we do not assume γ ∈ L. For example, γ(x) = log |log |ax|| or γ(x) = (log |ax|)s, 0 < s < 12, have finite

||∇ log γ||L2(B 1 2|a|

), thus can be used to construct examples. The second of these two examples also falls outside the conditions required by Theorem 1.9 of [3]. Therefore, our result is not a consequence of previous work. In view of the unique continuation property for the linear convection equation with L2 coefficients in R2[13], the assumption of γ ∈ W1,2 is most likely optimal for the two dimensional Calder´on’s problem. It is not yet certain that the restriction on the L2 norm of ∇ log γ is necessary.

This work is inspired by the paper of Cheng and Yamamoto in [9] where they prove a uniqueness result for the equation

4u + ~b · ∇u = 0

with ~b ∈ Lp, p > 2. Here we push their result to the optimal case p = 2 (with an added requirement that the L2 norm of ~b be bounded by a constant) and obtain the result for Calder´on’s problem as a corollary.

The proof relies on reducing the second order equation to a first order one in the complex plane (see [19], [1]

for background). It also employs an inverse scattering method developed by Beals and Coifman in [4] and by Sung in [15], [16], and [17]. Along the way we also prove a version of Brown’s result in [6] regarding the recovery of the boundary values of the conductivity from the knowledge of the Dirichlet–Neumann map.

National Center for Theoretical Sciences Mathematics Division, Taipei 106, Taiwan. Email: cicarstea@ncts.ntu.edu.tw

Institute of Applied Mathematical Sciences, NCTS, National Taiwan University, Taipei 106, Taiwan. Email: jn- wang@math.ntu.edu.tw

(2)

2 Statement of the result

We consider a domain Ω ⊂ R2 which is open, bounded, simply connected, and with sufficiently smooth boundary. For a conductivity γ : Ω → R positive and bounded away from zero, we will consider the boundary value problem

( ∇(γ∇u)(x) = 0, x ∈ Ω,

u|∂Ω= ω. (1)

The Dirichlet–Neumann map for the classical Calderon problem is then Λγ(ω) := γ|∂Ω

∂u

∂ν,

where u is the solution with boundary value ω. Equation (1) can be put in non-divergence form as ( 4u(x) + ~b(x) · ∇u(x) = 0, x ∈ Ω,

u|∂Ω= ω,

(2)

where

~b := ∇ log γ.

For this equation, we consider the Dirichlet–Neumann map Λ~b(ω) := ∂u

∂ν.

Throughout, we assume γ ∈ W1,2(Ω) or ~b ∈ L2R(Ω), where L2R(Ω) denotes the space of L2 real-valued vectors in Ω. For both forms of the equation, we consider strong solutions u ∈ W2,p(Ω), for some 1 < p < 2.

The Dirichlet boundary data then belong to W2−1p,p(∂Ω). In the appendix we give a proof of existence and uniqueness of solutions for the case when ||~b||L2(Ω)< C, with C = C(p, Ω).

Our main result is the following

Theorem 2.1. There exists a constant C = C(Ω) such that if γ1, γ2∈ W1,2(Ω), ||∇ log γ1,2||L2(Ω)< C and Λγ1 = Λγ2, then γ1= γ2.

This follows immediately from Theorem 2.2 and Proposition 2.1 stated below.

Theorem 2.2. There exists a constant C = C(Ω) such that if ~b1,~b2∈ L2R(Ω), ||~b1,2||L2(Ω)≤ C and Λ~b

1 = Λ~b

2, then ~b1= ~b2.

Note that this second theorem is more general than the first. It is this result that we will prove below.

We also prove the following boundary determination result.

Proposition 2.1. If γ ∈ W1,2(Ω), the trace γ|∂Ω can be determined from the Dirichlet-to-Neumann map Λγ.

Since the proof of Proposition 2.1 is of a different nature from that of Theorem 2.2, the principal result of this paper, we have provided it as an appendix.

3 Complex variable notation and the first order form of the equa- tion

We will identify R2with C and use the usual notation dz = dx1+i dx2, d¯z = dx1−i dx2, ∂z =12(∂x1−i∂x2),

z¯ =12(∂x1+i∂x2). If u ∈ W2,p(Ω) is a solution of (2), then w := ∂zu ∈ W1,p(Ω) is a solution of the equation

z¯w + Bw + ¯B ¯w = 0, (3)

where B(z) := 14(b1+ ib2). We can also go the other way:

(3)

Lemma 3.1. (e.g. see [19]) Given w ∈ W1,p(Ω) a solution to (3), then there exists a solution u ∈ W2,p(Ω) of (2) such that w = ∂zu.

For Dirichlet boundary conditions, as we have in (2), w must satisfy 2Re ( ˙zw)|∂Ω= ∂ω

∂t (the tangential derivative along ∂Ω). (4) Here ˙z is the derivative of a parametrization z(t) of the boundary ∂Ω. In the case of Neumann boundary conditions, we would instead need

2Im ( ˙zw)|∂Ω= ∂u

∂ν|∂Ω. (5)

For this see [1], section 16.5. Note that the relations (4) and (5) are valid for real solutions u of (2). Since ~b is a real vector, these relations remain true for complex-valued solutions of (2).

We define the Cauchy transform

(Cf )(z) := −1 π

Z

f (ζ) ζ − zd2ζ.

It maps C : Lq(Ω) → W1,q(Ω) continuously for all 1 < q < ∞ (again, see [1]). The operator C is an “inverse”

to the differential operator ∂z¯. For f ∈ C(Ω) ∩ C( ¯Ω)

(C∂¯zf )(z) = −1 πlim



Z

Ω−B(z,)

ζ¯

f (ζ) ζ − z

d ¯ζ ∧ dζ 2i

= 1 2πi lim



Z

∂B(z,)

f (ζ) − f (z)

ζ − z dζ + 2πif (z) − Z

∂Ω

f (ζ) ζ − zdζ

! , so

(C∂¯zf )(z) = f (z) − 1 2πi

Z

∂Ω

f (ζ)

ζ − zdζ. (6)

Also, defining hϕ, f i :=R

ϕf , for any ϕ ∈ Cc(Ω)

hϕ, ∂z¯Cf i = hC∂z¯ϕ, f i = hϕ, f i and so we have

z¯Cf = f. (7)

Given the mapping properties of the Cauchy transform, the formulas (6), (7) extend to any f ∈ W1,q(Ω), 1 < q < ∞.

The Cauchy transform may be used to turn (3) into an integral equation.

Lemma 3.2. A function w ∈ W1,p(Ω) that satisfies (3) will also satisfy the integral equation w(z) + C(Bw + ¯B ¯w)(z) = 1

2πi Z

∂Ω

w(ζ) ζ − zdζ.

Also, if Φ ∈ Hol(Ω) ∩ W1,p(Ω) and w ∈ W1,p(Ω) satisfies

w(z) + C(Bw + ¯B ¯w)(z) = Φ(z), (8)

then w satisfies (3) and

1 2πi

Z

∂Ω

w(ζ)

ζ − zdζ = Φ(z).

Next we would like to show the existence of solutions to (8), with suitable Φ. A contraction principle argument of the same form as the one in the proof of Proposition A.1 would work. Instead, we present a different type of argument which doesn’t require the norm of the coefficients to be small. Testing the equation (8) against a test function φ we get

hφ −

B + ¯Bw¯ w

C(φ), wi = hφ, Φi. (9)

(4)

Let A := B + ¯B ¯w/w ∈ L2(Ω), ||A||L2(Ω) ≤ 2||B||L2(Ω), and ψ := C(φ). Let 1 < p < 2 and p= 2p/(2 − p).

In order to obtain an a priori estimate for the Lp norm of w, we would like that

z¯ψ − Aψ = |w|p−2w.¯ (10)

It turns out that one solution of (10) is explicitly expressed by ψ = C

|w|p−2w e¯ −CA eCA.

The left hand side of (9) is just ||w||pLp∗(Ω). The right hand side of (9) consists of two terms:

h|w|p−2w, Φi + hAψ, Φi =: I + II.¯ H¨older’s inequality implies

I ≤ ||w||pLp∗−1(Ω)||Φ||Lp∗(Ω).

To deal with the second term we begin by noticing that |w|p−2w ∈ L¯ p/(p−1)(Ω). With the help of Trudinger’s inequality, we can see that eCA, e−CA∈ Lq(Ω) for any q < ∞ (for example, see the computation in [13]). It follows that ψ is in all Lr(Ω) with r < 2p/(p− 2). We can then bound the second term

II ≤ ||A||L2(Ω)||ψ||Lr(Ω)||Φ||Lp∗ +(Ω), r = 2(p+ ) p+  − 2 with  > 0. Since ||ψ||Lr(Ω)≤ C||w||pLp∗−1(Ω), we get

II ≤ C||w||pLp∗−1(Ω)||A||L2(Ω)||Φ||Lp∗ +(Ω). Putting these together implies that

kwkLp∗(Ω)≤ C(1 + kAkL2(Ω))kΦkLp∗ +(Ω)

and

kwkW1,p(Ω)≤ C(1 + kAkL2(Ω))kΦkLp∗ +(Ω).

If Φ ∈ W1,p+(Ω) for some  > 0 and B ∈ C0(Ω), then a solution to (8) is known to exits (see, for example, the proof of [9, Lemma 3.3]). Given B ∈ L2(Ω) let Bn ∈ C0(Ω) be such that ||Bn− B||L2 → 0, and let wn be the solutions of

wn+ C(Bnwn+ ¯Bnn) = Φ.

Then

||wn||W1,p+/2(Ω)≤ C||Φ||W1,p+(Ω)(1 + ||B||L2).

Also

wn− wm+ C(Bn(wn− wm) + Bn(wn− wm)) = −C((Bn− Bm)wm+ (Bn− Bm)wm).

According to the observation above, the right had side is in W1,p+/2(Ω) and

||RHS||W1,p+/2(Ω)≤ C||Bn− Bm||L2||Φ||W1,p+(Ω)(1 + ||B||L2).

Applying the a priori estimate to the difference wn− wm we get

||wn− wm||W1,p(Ω)≤ C||Bn− Bm||L2||Φ||W1,p+(Ω)(1 + ||B||L2)2.

This proves that the sequence of approximate solutions is Cauchy in W1,p(Ω). That the limit is a solution to the equation follows from the continuity of C and the continuity of products w.r.t. strong convergence.

We have therefore proven:

Lemma 3.3. If 1 < p < 2, Φ ∈ W1,p+(Ω),  > 0 with p +  < 2, then the equation w + C(Bw + ¯B ¯w) = Φ

has a unique solution w ∈ W1,p+/2(Ω) and the following estimate holds

kwkW1,p+/2(Ω)≤ CkΦkW1,p+(Ω), (11)

where C = C(kAkL2(Ω), p, , Ω).

(5)

4 CGO solutions (introduction of the parameter k)

In this section, we want to discuss complex geometrical optics (CGO) solutions of (3). These special solutions are useful in Calder´on’s problem. Let k = kx+ iky ∈ C. If w is a solution to (3) then αk(z) := w(z)e2ikz¯ satisfies the equation

¯zαk(z) + B(z)αk(z) + e2ikz+k ¯z)B(z)αk(z) = 0 in Ω. (12) We will use the notation

ek(z) = exp



−i

2(¯kz + k ¯z)



= ei(kxx+kyy). Clearly, |ek(z)| = 1. Analogously to the previous section, we have

Lemma 4.1. If αk∈ W1,p(Ω), 1 < p < 2, satisfies (12), then it also satisfies the integral equation αk(z) + C(Bαk+ ekB ¯¯αk)(z) = 1

2πi Z

∂Ω

αk(ζ) ζ − z dζ.

Conversely, the equation

αk(z) + C(Bαk+ ekB ¯¯αk)(z) = 1 (13) has a unique solution αk ∈ W1,p(Ω) and

||αk||W1,p(Ω)≤ C(||~b||L2(Ω), Ω, p).

This solution will also satisfy (12) and 1 2πi

Z

∂Ω

αk(ζ)

ζ − z dζ = 1.

4.1 Differentiability in k

In order to investigate the differentiability with respect to k of αk , for some fixed k = kx+ iky we introduce the notations

δκα = 1

κ(αk+κ(z) − αk(z)) and δκe = 1

κ(ek+κ(z) − ek(z)).

Note that this quantity satisfies the integral equation

δκα + C Bδκα + ek+κBδ¯ κα + δ¯ κe ¯B ¯αk = 0.

We will only consider real valued κ (the case of imaginary κ will only different by a minus sign). In this case δκα + C Bδκα + ek+κBδ¯ κα = −C (δκe) ¯B ¯αk . (14) Let  > 0 be such that p +  < 2, then it follows from (11) that

||δκα||W1,p+/2(Ω)≤ C(||~b||L2(Ω), p, )||δκe||L(Ω). In order to prove that δκα is Cauchy, we need to estimate the quantity

κ,κ0α := δκα − δκ0α.

Note that ∆κ,κ0e → 0 in C( ¯Ω). It is clear that ∆κ,κ0α satisfies the integral equation

κ,κ0α + C B∆κ,κ0α + ek+κ0B∆¯ κ,κ0α

= −C (∆κ,κ0e) ¯B ¯αk+ (ek+κ0− ek+κ) ¯Bδκα . Applying the a priori estimate (11) again we obtain

||∆κ,κ0α||W1,p(Ω)≤ C(||~b||L2(Ω), p, ) ||∆κ,κ0e||L(Ω)+ κ||δκe||L(Ω)+ κ0||δκ0e||L(Ω) .

We have thus shown that δκα is Cauchy in W1,p(Ω) as κ → 0 for real κ. Taking the limit in (14) we see that

kxα ∈ W1,p(Ω) satisfies

kxα + C B∂kxα + ekB∂¯ kxα = −C (ixek) ¯B ¯αk .

We can again apply (11) to conclude that ∂kxα is bounded in W1,p(Ω) uniformly in k. The case of imaginary κ is almost identical. We conclude that

Proposition 4.1. The partial derivatives of αk with respect to k exist and ∂kxαk, ∂kyαk ∈ L(Ck; W1,p(Ω)).

(6)

4.2 Behavior as k → ∞

Because the αkare bounded in W1,p(Ω) uniformly in k, we can extract a subsequence, also denoted αk, such that αk * α0 in Lp and αk → α0 in Lq for any 1 ≤ q < p. Of course, α0∈ W1,p(Ω).

To find the equation satisfied by α0 we are going to integrate (13) against a test function ϕ ∈ Cc(Ω), then for the first term we have

hϕ, αki → hϕ, α0i.

For the second term, since Cϕ ∈ L(Ω) and B ∈ L2(Ω),

hϕ, C(Bαk)i = −hBCϕ, αki → −hBCϕ, α0i = hϕ, C(Bα0)i.

To handle the third term, we write

hϕ, C(ekB ¯¯αk)i = −hekBCϕ, ¯¯ α0i − hekBCϕ, ¯¯ αk− ¯α0i.

Since ¯αk→ ¯α0 strongly in L2(Ω),

hekBCϕ, ¯¯ αk− ¯α0i → 0.

As ¯α0BCϕ ∈ L¯ p(Ω), the Riemann-Lebesgue lemma implies hekBCϕ, ¯¯ α0i → 0.

Putting these together we get that

hϕ, α0+ C(Bα0) − 1i = 0, for any ϕ ∈ Cc(Ω) and so

α0(z) + C(Bα0)(z) = 1, ∀z ∈ Ω.

Of course it then follows that

¯zα0+ Bα0= 0, and applying back the Cauchy transform we need to have

1 2πi

Z

∂Ω

α0(ζ)

ζ − z dζ = 1.

We have thus proved

Lemma 4.2. There exists a subsequence of solutions αk of (13) such that αk * α0 in W1,p(Ω) ⊂ Lp and αk→ α0 in Lq for any 1 ≤ q < p. The limit satisfies the integral equation

α0(z) + C(Bα0)(z) = 1, ∀z ∈ Ω,

and 1

2πi Z

∂Ω

α0(ζ)

ζ − z dζ = 1.

In fact

α0= e−C(B). (15)

Proof. We only have to show that the representation formula (15) holds. First not that extending B to the whole plane such that B = Bχand using the equation to extend α0 we have that α0∈ 1 + W1,p(C).

Define h := α0eC(B)and note that ∂z¯h = 0 in C, so it is holomorphic on C. Since B is supported within the bonded domain Ω, C(B) and C(Bα0) are both holomorphic on C − ¯Ω. Furthermore, they both decay as 1/z at infinity. It follows then that limz→∞h = 1 and by Liouville’s theorem this implies that h ≡ 1.

(7)

5 Cauchy transforms of B

1

and B

2

In this section, we would like to show that under the assumptions of Theorem 2.2, the Cauchy transforms of B1 and B2 are identical outside of the domain Ω. Let α1,k, α2,k∈ W1,p(Ω) be the solutions of the integral equations

αj,k+ C(Bjαj,k+ ekjα¯j,k) = 1, j = 1, 2.

First note that, according to Lemma 4.1, 1 2πi

Z

∂Ω

α1,k(ζ)

ζ − z dζ = 1, ∀z ∈ Ω.

Define w1(z, k) := ei2¯kzα1,k(z), w1(·, k) ∈ W1,p(Ω). It satisfies

¯zw1+ B1w1+ ¯B11= 0.

By Lemma 3.1, there exists a u1(·, k) ∈ W2,p(Ω) such that w1= ∂zu1and 4u1+ ~b1· ∇u1= 0.

Denote ω := u1|∂Ω∈ W2−1p,p(∂Ω) and let u2∈ W2,p(Ω) be the solution of ( 4u2+ ~b2· ∇u2= 0,

u2|∂Ω= ω.

The existence of u2 with boundary condition ω is guaranteed by choosing the constant C in the statement of Theorem 2.2 such that Proposition A.1 applies. Then w2:= ∂zu2will satisfy

¯zw2+ B2w2+ ¯B22= 0.

Since u1and u2share the same Dirichlet data, Re ( ˙zw1) = Re ( ˙zw2). Since we are assuming the Dirichlet–

Neumann maps produced by ~b1 and ~b2 are identical, the Neumann data of u1 must be the same as u2, so Im( ˙zw1) = Im ( ˙zw2). It follows then that

w1|∂Ω= w2|∂Ω. Define α02,k(z) := w2(z, k)ei2¯kz. It satisfies the differential equation

z¯α02,k+ B2α02,k+ ek2α¯02,k= 0 and, as α1,k|∂Ω= α02,k|∂Ω, we see that α02,ksatisfies the integral equation

α02,k+ C(B2α02,k+ ek2α¯02,k) = 1.

It follows the uniqueness of the solution that α02,k= α2,k. Consequently, we proved that Lemma 5.1. α1,k|∂Ω= α2,k|∂Ω.

We know from Lemma 4.2 that we can find a subsequence such that αj,k * αj,0 in W1,p(Ω), j = 1, 2, and

αj,0+ C(Bjαj,0) = 1.

The trace operator τ maps W1,p(Ω) to W1−1p,p(∂Ω) ⊂ L3−p2p (∂Ω) continuously and is compact from W1,p(Ω) to Ls(∂Ω) for 1 < s < 3−p2p . Therefore

αj,k|∂Ω→ αj,0|∂Ω, in Ls(∂Ω).

Lemma 5.1 them implies

α1,0|∂Ω= α2,0|∂Ω. We can now prove

(8)

Γ

z

R

Figure 1: Contour for the determination of e−C(Bj)(z).

Lemma 5.2. If Λ~b1= Λ~b2, then C(B1)(z) = C(B2)(z) for any z ∈ C − Ω.

Proof. By equation (15), we know that e−C(B1)|∂Ω= e−C(B2)|∂Ω.

Let z ∈ C − ¯Ω. Since e−C(Bj) is holomorphic in C − ¯Ω, using the integration contour from the figure, we can write

e−C(Bj)(z) = 1 2πi

Z

∂Ω

e−C(Bj)(ζ)

ζ − z dζ − 1 2πi

Z

|ζ|=R

e−C(Bj)(ζ) ζ − z dζ.

as ζ → ∞, e−C(Bj)= 1 + O(1ζ) so

lim

R→∞

1 2πi

Z

|ζ|=R

e−C(Bj)(ζ)

ζ − z dζ = 1.

It then follows that

e−C(B1)(z) = e−C(B2)(z),

for any z ∈ C − ¯Ω. This can happen only if C(B1)(z) = C(B2)(z) + 2πni. Since both Cauchy transforms vanish at infinity we must have n = 0.

6 A related first order system

To proceed further, we now would like to apply an inverse scattering method for a related first order system based on [4], [15], [16], and [17]. A similar idea was also used in [7] and [9]. We define

Cj:= eCBje−CBjj, j = 1, 2.

Note that, since CBj− CBj = 2iIm (CBj), Cj ∈ L2(Ω) and ||Cj||L2(Ω) = ||Bj||L2(Ω). We also define the matrix

Qj:=

 0 −Cj

−Cj 0

 .

Let µj(z) be the 2 × 2 matrix valued function that satisfies the integral equation µj(z, k) = I + C (ekQjµ¯j) = I − 1

π Z

ek(ζ)

ζ − zQj(ζ)µj(ζ, k) d2ζ. (16) Consider the orthogonal matrix

R := 1

√2

1 1 1 −1



Conjugating Qj by R we get

RQjRt=−Cj 0 0 Cj



(9)

Thus, conjugating the integral equation (16) by R, we obtain a decoupled system of four scalar integral equations. We then can apply the same method we have used to prove the existence of solutions to the equation (8) to show (16) has unique solutions µj(·, k) ∈ W1,p(Ω). These solutions satisfy the differential equation

z¯µj(z, k) − ek(z)Qj(z)µj(z, k) = 0 and the equality

I = 1 2πi

Z

∂Ω

µj(ζ, k)

ζ − z dζ, ∀z ∈ Ω, k ∈ C.

Also, just like the CGO solutions αk, µj are differentiable in k and ∂kxµj, ∂kyµj ∈ W1,p(Ω).

Define

η1(z, k) := ei2¯kzµ1(z, k).

This new matrix-valued function satisfies the differential equation

z¯η1= Q1η¯1. Let

v1:=1 1 i −i

 η1, then

z¯v1=1 2

1 1 i −i

 Q1

1 i 1 −i



¯

v1= −C11. Finally let

w1:= e−CB1v1

and we get that w1 satisfies the matrix differential equation

z¯w1(z, k) + B1(z)wj(z, k) + B1(z)w1(z, k) = 0, ∀z ∈ Ω, k ∈ C.

Applying Lemma 3.1 to the components of w1, we obtain that there is a matrix-valued function u1 ∈ W2,p(Ω) such that w1= ∂zu1 and

4u1+ ~b1· ∇u1= 0.

Let ω := u1|∂Ω∈ W2−1p,p(∂Ω) and let u2∈ W2,p(Ω) be the 2 × 2 matrix valued solution of the equation ( 4u2+ ~b2· ∇u2= 0,

u2|∂Ω= ω.

Just as before, we know that both the Dirichlet and the Neumann data of u1 and u2 coincide. Then w2:= ∂zu2 satisfies

¯zw2+ B2w2+ ¯B22= 0 and

w1|∂Ω= w2|∂Ω. Also v2:= eCB2w2 satisfies

z¯v2= −C2¯v2

and since, by Lemma 5.2, eCB1|∂Ω= eCB2|∂Ωwe have

v2|∂Ω= v1|∂Ω. Let

η2:=1 2

1 −i 1 i



v2, i.e., v2:=1 1 i −i

 η2, and finally

µ02:= e2i¯kzη2.

(10)

This last matrix-valued function satisfies the differential equation

¯zµ02= ekQ2µ¯02 and

µ1|∂Ω= µ02|∂Ω. Applying the Cauchy transform to the differential equation we get

µ02= Φ2+ C(ekQ2µ¯02), where

Φ2(z) = 1 2πi

Z

∂Ω

µ02(ζ)

ζ − z dζ = 1 2πi

Z

∂Ω

µ1(ζ)

ζ − z dζ = I.

So µ02≡ µ2and we have Lemma 6.1. µ1|∂Ω= µ2|∂Ω.

7 The ∂

¯k

equation

We now extend µj(z, k), j = 1, 2, to z ∈ C \ Ω by setting µj(z, k) = I + C (ekQjµ¯j) = I − 1

π Z

ek(ζ)

ζ − zQj(ζ)µj(ζ, k) d2ζ, ∀ z ∈ C \ Ω.

Abusing notation a little, if we write Qj= Qjχ, then µj(z, k) satisfies µj(z, k) = I + C (ekQjµ¯j) = I − 1

π Z

C

ek(ζ)

ζ − zQj(ζ)µj(ζ, k) d2ζ, ∀ z ∈ C, ∀ k ∈ C.

We can easily see that

lim

|z|→∞µj(z, k) = I, that

µj(·, k) ∈ W1,p(Ω), and µj(z, k) − I decays like 1z as z → ∞.

To simplify the notations, we suppress the index j. In other words, the computations below are carried out for µj, j = 1, 2, respectively. Define the matrix

ν(z, k) := µ11(z, k) e−k(z)µ12(z, k) e−kµ21(z, k) µ22(z, k)

! .

Note that

ν(z, k) = I + 1 π

Z

C(ζ)

ζ−¯¯ zν21(ζ, k) ek(ζ−z)C(ζ)ζ−z ν22(ζ, k)

ek(ζ−z)C(ζ)

ζ−z ν11(ζ, k) C(ζ)ζ−¯¯ zν12(ζ, k)

 d2ζ

and ν is the unique solution of this integral equation. Repeating the proof of Proposition 4.1, we can show that ∂k¯ν(z, k) exists in W1,p(Ω) and it satisfies

¯kν = e−k(z) 2πi

Z

0 ek(ζ)C(ζ)µ22(ζ, k) ek(ζ)C(ζ)µ11(ζ, k) 0

! d2ζ

+e−k(z) π

Z

C(ζ)

ζ−¯¯ zek(z)∂¯kν21(ζ, k) ek(ζ)C(ζ)ζ−zk¯ν22(ζ, k)

ek(ζ)C(ζ)

ζ−z¯kν11(ζ, k) C(ζ)ζ−¯¯ zek(z)∂k¯ν12(ζ, k)

 d2ζ.

(11)

This can be rewritten as ek(z)∂¯kν = 0 T12(k)

T21(k) 0

!

+1 π

Z

ek(ζ−z)C(ζ)

ζ−¯¯ z ek(ζ)∂¯kν21(ζ, k) C(ζ)ζ−zek(ζ)∂k¯ν22(ζ, k)

C(ζ)

ζ−zek(ζ)∂k¯ν11(ζ, k) ek(ζ−z)C(ζ)ζ−¯¯ z ek(ζ)∂¯kν12(ζ, k)

 d2ζ, (17) where we define

T12(k) = 1 2πi

Z

ek(ζ)C(ζ)µ22(ζ, k) d2ζ, T21(k) = 1

2πi Z

ek(ζ)C(ζ)µ11(ζ, k) d2ζ.

(18)

Comparing the integral equation (17) with

ν(z, k) 0 T12(k) T21(k) 0

!

=

= T21(k)ν12(z, k) T12(k)ν11(z, k) T21(k)ν22(z, k) T12(k)ν21(z, k)

!

= 0 T12(k) T21(k) 0

!

+ 1 π

Z

C

ek(ζ−z)C(ζ)

ζ−¯¯ z T21(k)ν22(ζ, k) C(ζ)ζ−zT12(k)ν21(ζ, k)

C(ζ)

ζ−zT21(k)ν12(ζ, k) ek(ζ−z)C(ζ)ζ−¯¯ z T12(k)ν11(ζ, k)

 d2ζ, we deduce from the uniqueness of integral equation (17) that

k¯ν(z, k) = e−k(z)ν(z, k) 0 T12(k) T21(k) 0

! . Since

µ11= 1 − C(ekC) + C(ekC ¯C(e−kC ¯¯µ11)), µ22= 1 − C(ekC) + C(ekC ¯C(e−kC ¯¯µ22)), it follows that µ11= µ22, so also T12= T21. Now let us denote

T (k) =

 0 T12(k) T12(k) 0



and specify Tj(k) be as T above defined for Cj, j = 1, 2. At this point we can also note that

k¯ν(z, k) = e−k(z)ν(z, k)T (k) = e−k(z)T (k)ν(z, k).

We have

Lemma 7.1. T1(k) = T2(k) for all k.

The proof of this is identical to the argument given for the corresponding statement of [9, (4.25)] as part of the proof of their Lemma 4.4. The result of our Lemma 6.1 is needed in the proof.

We can now use a known result of Brown on the mapping properties of the scattering map to conclude that

Lemma 7.2. C1= C2.

Proof. Note that the off-diagonal entries of Qjare zero. We use Theorem 2 of [5]. There it is shown that the mapping Q → T is invertible, hence injective, provided kQkL2 <√

2. This can be arranged by choosing the constant C in the statement of Theorem 2.2 appropriately. Since we have already determined that T1= T2, it follows that Q1= Q2.

(12)

8 Conclusion of the argument

By Lemma 7.2 we have that

B1eCB1−CB1 = B2eCB2−CB2, ∀ z ∈ Ω, and hence

B1eCB1−CB1 = B2eCB2−CB2, ∀ z ∈ C

since B1 = B2 = 0 in C \ Ω. In view of Lemma 5.2, if we let Z := C(B2− B1), then Z ∈ W1,2(C) and supp(Z) ⊆ Ω. Denote

κ := eZ−Z¯ . Then we obtain

B1= κB2

and

z¯Z = (1 − κ)B2, z ∈ C.

By an easy computation (see [9, P. 1389]), we can deduce that

|κ − 1| ≤ 2|Z|

and thus

|∂z¯Z| ≤ 2|B2| |Z|

holds. Applying the generalized Liouville theorem of [7, Corollary 3.11] (or [1, Theorem 5.8.3]), it follows then that Z ≡ 0 and so B1≡ B2. This ends the proof of Theorem 2.2.

A Existence of strong solutions for coefficients with small norm

In this appendix, we will show that if k~bkL2(Ω)is not too large, then (2) has a unique solution for any Dirichlet condition in W2−p1,p(∂Ω). The proof is based on a contraction mapping principle argument.

Proposition A.1. There exists a constant c(p, Ω) > 0 such that if ||~b||L2(Ω)< c(p, Ω), then (2) has a unique solution u ∈ W2,p(Ω).

Proof. Let H ∈ W2,p(Ω) be the unique solution to the Dirichlet boundary value problem:

( 4H(x) = 0, x ∈ Ω, H|∂Ω= ω.

Defining u0:= u − H, the equation becomes

4u0(x) + ~b(x) · ∇u0(x) = −~b · ∇H(x) (19) with u0∈ W01,p(Ω). Let L(f ) be the solution of the inhomogeneous problem

( 4U (x) = f, x ∈ Ω, U |∂Ω= 0.

Then L : Lp(Ω) → W2,p(Ω) ∩ W01,p(Ω) is continuous (e.g. Theorem 9.15 of [10]). If (19) can be solved, then the solution should satisfy

u0= −L~b · ∇u0+ ~b · ∇H . Seeing this, we define the operator T for functions v ∈ W2,p(Ω) by

T v := −L~b · ∇v +~b · ∇H .

(13)

Since ∇v ∈ W1,p(Ω) ⊂ Lp(Ω), p1 = 1p12, it follows that ~b · ∇v ∈ Lp(Ω). The same is true of the second term in the definition of T . Putting these together we get that there is a constant C(p, Ω) > 0 such that

||T v||W2,p(Ω)≤ C(p, Ω)



||~b||L2(Ω)||v||W2,p(Ω)+ ||~b||L2(Ω)||ω||

W2− 1p,p(∂Ω)

 . So T : W2,p(Ω) → W2,p(Ω). Moreover, for two v1, v2∈ W2,p(Ω),

||T v1− T v2||W2,p(Ω)≤ C(p, Ω)||~b||L2(Ω)||v1− v2||W2,p(Ω).

If ||~b||L2(Ω)< C(p,Ω)1 , the operator T is a contraction hence it will have a fixed point: the solution of (19).

Uniqueness follows easily from these same estimates.

B Recovering the conductivity at the boundary

Since γ ∈ W1,2(Ω) we can consider its trace γ|∂Ω ∈ W12,2(∂Ω). In this section, following the method of Brown in [6], we will prove that we can recover γ|∂Ω from the Dirichlet-to-Neumann map. We cannot quote the result of [6] directly since there the conductivity is assumed to be bounded.

First we need to straighten out the boundary. If P ∈ ∂Ω, then there is a ball B(P, R) and a diffeomorphism φ : Ω → ˜Ω such that φ(Ω ∩ B(P, R)) ⊂ {y2> 0} and 0 ∈ φ(∂Ω ∩ B(P, R)) ⊂ {y2= 0}. In the new coordinates the equation (1), denoting F := φ−1 and v := u ◦ F , becomes

k(akjjv) = 0, where

akj(y) := γ(F (y))∂iφk(F (y))∂iφj(F (y)) det(DF )(y).

Hereafter, we adopt the summation convention. We can make sure that Dφ ∈ C1∩ L, DF ∈ C1∩ L, so akj∈ Wloc1,2.

Choose η : R → [0, ∞), supp η ⊂ [−1, 1], η|[−12,12]≡ 1, smooth. Also choose α ∈ R2 such that

iφj(P )αjiφk(P )αk = ∂iφj(P )δj2iφk(P )δk2,

iφj(P )αjiφk(P )δk2= 0.

In other words, R2 3 α satisfies |Dφ(p)α| = |Dφ(P )e2| and Dφ(P )α ⊥ Dφ(P )e2. Define µ := Dφ(P )(iα − e2) ∈ C2. Then µ · µ = 0, µ · ¯µ = 2|Dφ(P )e2|2> 0. With these notations we introduce the functions

vN(y) := η(N1/2y1)η(N1/2y2)eN (iα−e2)·y=: ψ(N1/2y)E(N y), where N ∈ N.

Lemma B.1. There is a non-zero constant A such that Z

˜

aij(y)∂ivN(y)∂jN(y) dy = N12γ(F (0))A + o(N12), as N → ∞.

Proof. Let ˜γ(y) := γ(F (y)) det(DF )(y). We compute Z

˜

aij(y)∂ivN(y)∂j¯vN(y) dy

= N2 Z

˜

˜

γ(0)µ · ¯µψ(N1/2y)2e−2N y2 dy + N2

Z

˜

(˜γ(y) − ˜γ(0)) µ · ¯µψ(N1/2y)2e−2N y2 dy + N

Z

˜

aij(y)(∂iψ)(N1/2y)(∂jψ)(N1/2y)e−2N y2 dy

− N3/2 Z

˜

aij(y)δj2(∂iψ)(N1/2y)ψ(N1/2y)e−2N y2 dy

=: I1+ I2+ I3+ I4. (20)

(14)

We begin with the estimate of I4. Note that

N32|I4| ≤ sup

i,j

Z N− 12

−N− 12

Z N− 12

0

|aij(0)|e−2N y2 dy2dy1

+ sup

i,j

Z N− 12

−N− 12

Z N− 12

0

|aij(y1, 0) − aij(0)|e−2N y2 dy2dy1

+ sup

i,j

Z N− 12

−N− 12

Z N− 12

0

|aij(y1, y2) − aij(y1, 0)|

y2

y2e−2N y2 dy2dy1

=: J1+ J2+ J3. By direct computation, we have that

J1= γ(F (0))CN32 for some constant C. If 0 is a Lebesgue point for aij|∂ ˜ then

N12 Z N− 12

−N− 12

|aij(y1, 0) − aij(0)| dy1→ 0, which gives that also J2= o(N32). And finally Hardy’s inequality implies

J3≤ C sup

i,j

||∇aij||L2N74.

Hence I4= O(1). The same type of estimates give I3= O(N12).

The first term of (20) is easily seen by direct computation to satisfy I1= γ(F (0))AN12 + o(N12).

Finally, arguing in the same way as in the case of the terms J2and J3we have N−2I2=

Z

˜

(γ(F (y1, 0)) − γ(F (0))) µ · ¯µψ(N1/2y)2e−2N y2 dy +

Z

˜

(γ(F (y1, y2)) − γ(F (y1, 0))) µ · ¯µψ(N1/2y)2e−2N y2 dy

= o(N32) + O(N74), which gives I2= o(N12).

Lemma B.2. ||∂i(aijjvN)||H−1( ˜Ω) = o(N14).

Proof. We begin by splitting the left hand side into several terms

i(aijjvN)(y) = aij(0)N3/2(∂iψ)(N1/2y)(iα − e2)jE(N y)

+ aij(0)N (∂ijψ)(N1/2y)E(N y) + ∂i

(aij(y) − aij(0))∂j(ψ(N1/2y)E(N y))

= I1+ I2+ I3. Let ϕ ∈ H01( ˜Ω) (we use the norm ||ϕ||H1

0 = ||∇ϕ||L2) and denote δ(y) the distance between y and ∂ ˜Ω.

N32 Z

˜

ϕ(y)I1(y) dy

≤ |aij(0)|

Z

˜

ϕ(y)

δ(y)δ(y)(∂iψ)(N1/2y)|(iα − e2)j|e−N y2 dy

≤ sup

ij

|aij(0)|

Z

˜

ϕ(y)2 δ(y)2 dy

12

 Z N− 12

−N− 12

Z N− 12

0

y22e−2N y2 dy

1 2

≤ C||ϕ||H1 0N74,

(15)

hence

||I1||H−1( ˜Ω)= O(N14).

In the same way we get

||I2||H−1( ˜Ω)= O(N34).

For the third term, we have

Z

˜

ϕ(y)I3(y) dy

≤ C||∇ϕ||L2

×

N1/2sup

i

 Z N− 12

−N− 12

Z N− 12

0

(aij(y) − aij(0))2(∂jψ)2(N1/2y)e−2N y2 dy

1 2

+ N sup

ij

 Z N− 12

−N− 12

Z N− 12

0

(aij(y) − aij(0))2ψ2(N1/2y)e−2N y2 dy

1 2

≤ C||∇ϕ||L2

h

N1/2J1+ N J2

i . We first look at J1,

J1≤ sup

ij

 Z N− 12

−N− 12

Z N− 12

0

(aij(y1, y2) − aij(y1, 0))2

y22 y22e−2N y2 dy

1 2

+ sup

ij

 Z N− 12

−N− 12

Z N− 12

0

(aij(y1, 0) − aij(0))2e−2N y2 dy

1 2

= K1+ K2. Using the inequality

s2e−2N s≤ N−2e−2, ∀s ∈ [0, ∞),

we get K1 = O(N−1). Using the same methods as in previous estimates, we also have K2 = o(N34). This gives J1= o(N34) and similarly also J2= o(N34). Then we have

||I3||H−1( ˜Ω)= o(N14) and the lemma is proved.

Lemma B.3. The boundary value problem

i(aijjwN) = ∂i(aijjvN), wN|∂ ˜≡ 0, has solutions that satisfy ||wN||H1

0( ˜Ω)= o(N14).

Proof. Existence of weak solutions follows from applying the standard Lax-Milgram argument in the energy space

Ha :=



v : ˜Ω → C : Z

˜

aij(y)∂iv(y)∂jv(y) dy < ∞, v|∂ ˜= 0

 .

Recall that γ is positive and bounded away from zero. So aij is a positive-definite matrix function and we can see that Ha ⊂ H01( ˜Ω). Since H−1 ⊂ (Ha)0 the right hand side of the equation the usual method can indeed be applied. Integrating the equation against wN we obtain

||∇wN||2L2 ≤ ||∇wN||L2||∂i(aijjvN)||H−1

and the result follows.

參考文獻

相關文件

6 《中論·觀因緣品》,《佛藏要籍選刊》第 9 冊,上海古籍出版社 1994 年版,第 1

Full credit if they got (a) wrong but found correct q and integrated correctly using their answer.. Algebra mistakes -1% each, integral mistakes

(12%) Among all planes that are tangent to the surface x 2 yz = 1, are there the ones that are nearest or farthest from the origin?. Find such tangent planes if

[r]

Wang, Unique continuation for the elasticity sys- tem and a counterexample for second order elliptic systems, Harmonic Analysis, Partial Differential Equations, Complex Analysis,

These results cease to hold in the general case of complex coefficients, as Alinhac [2] has constructed such operators of any order in R 2 without the strong unique

Spectral fractional elliptic operators; Caffarelli-Silvestre- Stinga type extension; Non-local operators; Weighted eigenvalue problems; Unique continuation from a measurable

39) The osmotic pressure of a solution containing 22.7 mg of an unknown protein in 50.0 mL of solution is 2.88 mmHg at 25 °C. Determine the molar mass of the protein.. Use 100°C as