• 沒有找到結果。

Carleman estimate for complex second order elliptic operators with discontinuous Lipschitz coefficients

N/A
N/A
Protected

Academic year: 2022

Share "Carleman estimate for complex second order elliptic operators with discontinuous Lipschitz coefficients"

Copied!
32
0
0

加載中.... (立即查看全文)

全文

(1)

Carleman estimate for complex second order elliptic operators with discontinuous Lipschitz coefficients

E. Francini

S. Vessella

J-N. Wang

February 16, 2020

Abstract

In this paper, we derive a local Carleman estimate for the complex second order elliptic operator with Lipschitz coefficients having jump discontinuities.

Combing the result in [BL] and the arguments in [DcFLVW], we present an ele- mentary method to derive the Carleman estimate under the optimal regularity assumption on the coefficients.

1 Introduction

Carleman estimates are important tools for proving the unique continuation property for partial differential equations. Additionally, Carleman estimates have been suc- cessfully applied to study inverse problems and controllability of partial differential equations. Most of Carleman estimates are proved under the assumption that the leading coefficients possess certain regularity. For example, for general second or- der elliptic operators, Carleman estimates were proved when the leading coefficients are at least Lipschitz [Ho3]. In general, the Lipschitz regularity assumption is the optimal condition for the unique continuation property to hold in Rn with n ≥ 3 (see counterexamples constructed by Pli´s [P] and Miller [M]). Therefore, Carleman estimates for second order elliptic operators with general discontinuous coefficients are most likely not valid. Nonetheless, recently, in the case of coefficients having jump discontinuities at an interface with homogeneous or non-homogeneous trans- mission conditions, one can still prove useful Carleman estimates, see, for example, Le Rousseau-Robbiano [LR1], [LR2], Le Rousseau-Lerner [LL], and [DcFLVW].

Above mentioned results are proved for real coefficients. 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 electromagnetic

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@ntu.edu.tw

(2)

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. For instance, in the human body, different organs have different conductivities. Therefore, to model the current flow in the human body, it is more reasonable to consider an anisotrotopic complex-valued conductivity with jump-type discontinuities [MPH].

With potential applications in mind, our goal in this paper is to derive a Car- leman estimate for the second order elliptic equations with complex-valued leading coefficients having jump-type discontinuities. Although such a Carleman estimate has been derived in [BL], we want to remark that the method used in [BL], also in [LR1], [LR2], and [LL], are based on the technique of pseudodifferential operators and hence requires C coefficients and interface; while the method in [DcFLVW]

(and its parabolic counterpart, [FV]) relies on the Fourier transform and a version of partition of unity which requires only Lipschitz coefficients and C1,1 interface. Hence, the main purpose of the paper is to extend the method in [DcFLVW], [FV] to second order elliptic operators with complex-valued coefficients. It is important to point out that even though second order elliptic operators with complex-valued coefficients can be written as a coupled second order elliptic system with real coefficients, neither the method in [LR1], [LR2], [LL] nor that in [DcFLVW] can be applied to coupled elliptic systems. Therefore, we need to work on operators with complex-valued coefficients directly.

Our strategy to derive the Carleman estimate consists of two major steps. In the first step, we treat second order elliptic operators with constant complex coeffi- cients. 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 second order elliptic operators with constant complex coefficients from [BL, Theorem 1.6]. We would like to mention that the result in [BL] is stated for quite general complex coefficients, but here we can only verify the transmission condition with our choice of weight functions for complex coefficients having small imaginary parts. So in this paper we will consider this case. In the second step, we extend the Carleman estimate to the operator with non-constant complex coefficients with small imaginary parts. This method in this step is taken from the argument in [DcFLVW, Section 4]. The key tool is a version of partition of unity.

Furthermore, in the second step, we need an interior Carleman estimate for sec- ond order elliptic operators having Lipschitz leading coefficients and with the weight function ψε. An interior Carleman estimate was proved in [Ho1, Theorem 8.3.1], but for operators with C1 leading coefficients. Another interior estimate was established in [Ho3, Proposition 17.2.3] for operators with Lipschitz leading coefficients, but with a different weight function. H¨ormander remarked in [Ho4] (page 703, line 7-8) that

”Inspection of proof of Theorem 8.3.1 in [Ho1] shows that only Lipschitz continu- ity was actually used in the proof.” But, as far as we can check, there is no formal proof of this statement in literature. To make the paper self contained, we would like give a detailed proof of interior Carleman estimate for second order elliptic opera-

(3)

tor with Lipschitz leading coefficients and with a rather general weight function, see Proposition4.1. This interior Carleman estimate may be useful on other occasions.

In this paper, we present a detailed and elementary derivation of the Carleman es- timate for the second order elliptic equations with complex-valued coefficients having jump-type discontinuities following our method in [DcFLVW]. Having established the Carleman estimate, we then can apply the ideas in [FLVW] to prove a three-region inequality and those in [CW] to prove a three-ball inequality across the interface.

With the help of the three-ball inequality, we can study the size estimate problem for the complex conductivity equation following the ideas in [CNW]. We will present these quantitative uniqueness results and the application to the size estimate in the forthcoming paper.

The paper is organized as follows. In Section 2, we introduce notations that will be used in the paper and the statement of the theorem. In Section 3, we derive a Carleman estimate for the operator having discontinuous piecewise constant coeffi- cients. This Carleman estimate is a special case of [BL, Theorem 1.6]. Therefore, the main task of Section 3 is to check the transmission condition and the strong pseu- doconvexity condition. Finally, the main Carleman estimate is proved in Section 4.

The key ingredient is a partition of unity introduced in [DcFLVW].

2 Notations and statement of the main theorem

We will state and prove the Carleman estimate for the case where the interface is flat. Since our Carleman estimate is local near any point at the interface, for 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±), (2.1) where

A±(x) = {a±`j(x)}n`,j=1 = {a±`j(x0, xn)}n`,j=1, x0 ∈ Rn−1, xn ∈ R (2.2) is a Lipschitz symmetric matrix-valued function. Assume that

a±`j(x) = a±j`(x), ∀ `, j = 1, · · · , n, (2.3)

(4)

and furthermore

a±`j(x) = M`j±(x) + iγN`j±(x), (2.4) where (M`j±) and (N`j±) are real-valued matrices and γ > 0. We further assume that there exist λ0, Λ0 > 0 such that for all ξ ∈ Rn and x ∈ Rn we have

λ0|ξ|2 ≤ M±(x)ξ · ξ ≤ Λ0|ξ|2 (2.5) and

λ0|ξ|2 ≤ N±(x)ξ · ξ ≤ Λ0|ξ|2. (2.6) 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|. (2.7) To treat the transmission conditions, we write

h0(x0) := u+(x0, 0) − u(x0, 0), ∀ x0 ∈ Rn−1, (2.8) h1(x0) := A+(x0, 0)∇u+(x0, 0) · ν − A(x0, 0)∇u(x0, 0) · ν, ∀ x0 ∈ Rn−1, (2.9) 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, (2.10) 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, (2.11)

and let

φδ(x) := ψδ−1x), δ > 0. (2.12) 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. (2.13)

(5)

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 (2.13) 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).

Theorem 2.1 Let u and A±(x) satisfy (2.1)-(2.9). There exist α+, α, β, δ0, r0, γ0 and C depending on λ0, Λ0, M0 such that if γ < γ0, δ ≤ δ0 and τ ≥ C, 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

 . (2.14) where u = H+u++ Hu, u± ∈ C(Rn) and supp u ⊂ Bδr0 0 × [−δr0, δr0], and φδ is given by (2.12).

Remark 2.2 Estimate (2.14) is a local Carleman estimate near xn = 0. As men- tioned above, by flattening the interface, we can derive a local Carleman estimate near a C1,1 interface from (2.14). Nonetheless, an estimate like (2.14) is sufficient for some applications such as the inverse problem of estimating the size of an inclu- sion by one pair of boundary measurement (see, for example, [FLVW]).

3 Carleman estimate for operators with constant coefficients

The purpose of this section is to derive (2.14) for L(x, D) with discontinuous piecewise constant coefficients. More precisely, we derive (2.14) for L0(D), where L0(D) is

(6)

obtained from L(x, D) by freezing the variable x at (x00, 0). Without loss of generality, we take (x00, 0) = (0, 0) = 0 and thus

L0(D)u = L(0, D)u =X

±

H±div(A±(0)∇u±).

Since L0 has piecewise constant coefficients, to prove (2.14), we will apply [BL, The- orem 1.6]. So the task here is to verify the strong pseudoconvexity and transmission conditions for operator L0 with the weight function given in (2.11).

To streamline the presentation, we define Ω1 := {xn < 0}, Ω2 := {xn > 0}. On each side of the interface, we have complex second order elliptic operators. We denote

Pk= X

1≤j,`≤n

a(k)`j D`Dj, k = 1, 2,

where a(1)`j = a`j and a(2)`j = a+`j. Here we denote a(k)`j = a(k)`j (0). Corresponding to (2.3)-(2.6), we have

a(k)`j = a(k)j` , (3.1)

a(k)`j = M`j(k)+ iγN`j(k), (3.2) λ0|ξ|2 ≤ M(k)ξ · ξ ≤ Λ0|ξ|2, (3.3) λ0|ξ|2 ≤ N(k)ξ · ξ ≤ Λ0|ξ|2. (3.4) Since some computations in the verification of the transmission conditions are useful in proving the strong pseudoconvexity condition, we will begin with the discussion of the transmission conditions at the interface {xn= 0}.

3.1 Transmission conditions

We consider the natural transmission conditions that use the interface operators Tk1 = (−1)k, Tk2 = (−1)k X

1≤j≤n

a(k)njDj

that correspond to the continuity of the solution and of the normal flux, respectively.

We now write the weight function

ψε(x) = ϕ(xn) − ε

2|x0|2, (3.5)

where

ϕ(xn) =

1(xn), xn < 0 ϕ2(xn), xn ≥ 0, and

ϕk(xn) = αkxn+ 1 2βx2n

(7)

with α1, α2 > 0 (corresponding to α and α+ in (2.10), respectively) and β > 0.

Notice that ϕ is smooth in Ω1, Ω2 and is continuous across the interface. Then we have

∇ψε(0) =

((0, · · · , 0, α1), xn< 0 (0, · · · , 0, α2), xn≥ 0.

Following the notations and the calculations in [BL, Section 1.7.1], we have for ω := (0, ξ0, ν, τ ) with ξ0 = (ξ1, · · · , ξn−1) 6= 0, ν = en and λ ∈ C,

˜t1k,ψε(ω, λ) = (−1)k and

˜t2k,ψε(ω, λ) = (−1)ka(k)nn((−1)kλ + iτ ∂xnψε(0)) + (−1)k X

1≤j≤n−1

a(k)njj+ iτ ∂xjψε(0))

= (−1)ka(k)nn((−1)kλ + iτ αk) + (−1)k X

1≤j≤n−1

a(k)njξj.

The principal symbols of Pk, k = 1, 2, can be written as pk(ξ) = a(k)nn((ξn+ X

1≤j≤n−1

a(k)nj a(k)nn

ξj)2+ bk0)), (3.6)

where

bk0) = (a(k)nn)−2 X

1≤`,j≤n−1

(a(k)`j a(k)nn − a(k)n`a(k)nj`ξj. (3.7) We also need to introduce the principal symbol of the conjugate operators

˜

pk,ψε(ω, λ) = a(k)nn h

(−1)kλ + iτ ∂xnψε(0) + X

1≤j≤n−1

a(k)nj a(k)nn

j+ iτ ∂xjψε(0))2

+ bk0+ iτ ∂x0ψε(0))i

= a(k)nnh

(−1)kλ + iτ αk+ X

1≤j≤n−1

a(k)nj a(k)nn

ξj2

+ bk0)i .

(3.8)

Let us introduce A(k), B(k) ∈ R for k = 1, 2 such that bk0) = (a(k)nn)−2 X

1≤`,j≤n−1

(a(k)`j a(k)nn − a(k)n`a(k)nj`ξj

= (A(k)− iB(k))2,

(3.9)

where A(k)≥ 0. We also denote X

1≤j≤n−1

a(k)nj a(k)nn

ξj = E(k)+ iF(k), (3.10)

(8)

where E(k), F(k)∈ R. Using (3.8), (3.9), and (3.10), we can write

˜

p2,ψε = a(2)nn[(λ + iτ α2+ E(2)+ iF(2))2+ (A(2)− iB(2))2]

= a(2)nn[(λ + iτ α2+ E(2)+ iF(2)+ i(A(2)− iB(2)))

· (λ + iτ α2+ E(2)+ iF(2)− i(A(2)− iB(2)))]

= a(2)nn(λ − σ1(2))(λ − σ2(2)), where

σ1(2) = −E(2)− B(2)− i(τ α2+ F(2)+ A(2)), σ2(2) = −E(2)+ B(2)− i(τ α2+ F(2)− A(2)).

On the other hand, we can write

˜

p1,ψε = a(1)nn[(−λ + iτ α1+ E(1)+ iF(1))2+ (A(1)− iB(1))2]

= a(1)nn[(λ − iτ α1− E(1)− iF(1)+ i(A(1)− iB(1)))

· (λ − iτ α1− E(1)− iF(1)− i(A(1)− iB(1)))]

= a(1)nn(λ − σ1(1))(λ − σ2(1)), where

σ1(1) = E(1)+ B(1)+ i(τ α1+ F(1)+ A(1)), σ2(1) = E(1)− B(1)+ i(τ α1+ F(1)− A(1)).

Let us introduce the polynomial

Kk,ψε(ω, λ) := Y

Im σ(k)j ≥0

(λ − σ(k)j ).

Now we state the definition of transmission conditions given in [BL, Definition 1.4].

Definition 3.1 The pair {Pk, ψε, Tkj, k = 1, 2, j = 1, 2} satisfies the transmission condition at ω if for any polynomials q1(λ), q2(λ), there exist polynomials U1(λ), U2(λ) and constant c1, c2 such that

(q1(λ) = c1˜t11,ψε(ω, λ) + c2˜t21,ψε(ω, λ) + U1(λ)K1,ψε(ω, λ), q2(λ) = c1˜t12,ψε(ω, λ) + c2˜t22,ψε(ω, λ) + U2(λ)K2,ψε(ω, λ).

In order to check the transmission conditions, we need to study the polynomial Kk,ψε(ω, λ). For this reason, we need to determine the signs of the imaginary parts of the roots σ(k)j defined above. Note that we can write

bk0) = 1 a(k)nn

X

1≤`,j≤n−1

a(k)`j ξ`ξj − (E(k)+ iF(k))2. (3.11)

(9)

Since bkplays an essential role, we begin by working some calculations on the matrix

1 a(k)nn

A(k), where A(k) is the matrix (a(k)`j ). Let a(k)nn = |a(k)nn|e. Choosing ξ = en, we have that

a(k)nn = X

1≤`,j≤n

a(k)`j ξ`ξj = X

1≤`,j≤n

M`j(k)ξ`ξj+ iγ X

1≤`,j≤n

N`j(k)ξ`ξj.

Hence, from (3.3), (3.4), we have that

λ0 ≤ Re (a(k)nn) ≤ Λ0 and λ0 ≤ Im (a(k)nn) γ ≤ Λ0 and so that θ ∈ [0, π/2). Let us evaluate

(a(k)nn)−1A(k) = |a(k)nn|−1(M(k)+ iγN(k))(cos θ − i sin θ)

= |a(k)nn|−1[cos θM(k)+ γ sin θN(k)+ i(− sin θM(k)+ γ cos θN(k))].

(3.12) Using (3.3), (3.4) again, we see that for ξ ∈ Rn

Re ((a(k)nn)−1A(k)ξ · ξ) = |a(k)nn|−1[cos θM(k)ξ · ξ + γ sin θN(k)ξ · ξ]

≥ |a(k)nn|−1λ0(cos θ + γ sin θ)|ξ|2. (3.13) In fact, since cos θ = Mnn(k)|a(k)nn|−1 and sin θ = γNnn(k)|a(k)nn|−1, while |a(k)nn|2 = (Mnn(k))2+ γ2(Nnn(k))2, we have

|a(k)nn|−1(cos θ + γ sin θ) = Mnn(k)+ γ2Nnn(k)

(Mnn(k))2+ γ2(Nnn(k))2 ≥ λ0(1 + γ2) Λ20(1 + γ2) = λ0

Λ20. (3.14) Combining (3.13) and (3.14) implies

Re ((a(k)nn)−1A(k)ξ · ξ) ≥ λ20

Λ20|ξ|2 := ˜λ1|ξ|2. (3.15) Now let us write

λ˜1|ξ|2 ≤ Re ((a(k)nn)−1A(k)ξ · ξ)

=Re [ X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj+ 2 X

1≤j≤n−1

a(k)nj a(k)nn

ξnξj+ ξn2]

n2+ 2b(k)00n+ b(k)10),

(3.16)

where

b(k)00) = Re ( X

1≤j≤n−1

a(k)nj a(k)nn

ξj) = Re (E(k)+ iF(k)) = E(k)

(10)

and

b(k)10) = Re ( X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj).

Substituting ˜ξn = ξn= −b(k)00) into (3.16) gives

λ˜1(|ξ0|2+ | ˜ξn|2) ≤ ˜ξn2− 2b(k)00) ˜ξn+ b(k)10) = −(b(k)00)2+ b(k)10), which implies

˜λ10|2 ≤ Re ( X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj) − Ek2. (3.17) Putting (3.11) and (3.17) together gives

Re (bk(x0, ξ0)) = Re ( X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj) − (E(k))2+ (F(k))2

≥ ˜λ10|2+ (F(k))2 > 0.

(3.18)

The following lemma guarantees the positivity of A(k). Lemma 3.1 Assume that (3.3) and (3.4) hold. Then

A(k)

qλ˜10|2 + |F(k)|2 > |F(k)|. (3.19) ProofFrom (3.9), it is easy to see that

A(k) = Rep bk =

s a +√

a2 + b2

2 ,

where a = Re bk and b = Im bk. We have from (3.18) that a > 0 and thus A(k) ≥√

a ≥

q˜λ10|2+ (F(k))2 > |F(k)|.

2

Lemma 3.1 implies

Im σ(2)1 = −(τ α2+ F(2)+ A(2)) = −τ α2− F(2)− A(2)

≤ −τ α2− |F(2)| − F(2) ≤ −τ α2 < 0 (3.20) and

Im σ1(1) = τ α1+ F(1)+ A(1) > τ α1+ F(1)+ |F(1)| ≥ τ α1 > 0. (3.21)

(11)

We are now ready to check the transmission condition defined in Definition 3.1.

Being able to satisfy this condition depends on the degree of K1,ψε and K2,ψε, that is, on the number of roots with negative imaginary parts.

Case 1. ˜p2,ψε has two roots in {Im z < 0}, i.e., −τ α2 − F(2) + A(2) < 0 in view of (3.20). In this case, we have that

K2,ψε = 1, while K1,ψε has degree 1 or 2 (note (3.21)).

Since ˜t12,ψ

ε(ω, λ) = 1 and

˜t22,ψε(ω, λ) = a(2)nn(λ + iτ α2+ X

1≤j≤n−1

a(2)nj a(2)nn

ξj),

for any q2(λ), we simply choose

U2(λ) = q2(λ) − c1˜t12,ψε − c222,ψε. On the other hand, we have ˜t11,ψε(ω, λ) = −1 and

˜t21,ψ

ε(ω, λ) = a(1)nn(λ − iτ α1− X

1≤j≤n−1

a(1)nj a(1)nn

ξj).

Then for any polynomial q1(λ), we choose U1(λ) to be the quotient of the division between q1 and K1,ψε. The remainder term is equal to c1˜t1,ψε + c2˜t2,ψε with suitable c1, c2.

Case 2. Assume that Im σ2(2) ≥ 0 and Im σ2(1) ≥ 0, i.e.,

−τ α2− F(2)+ A(2) ≥ 0, τ α1+ F(1)− A(1) ≥ 0.

Then K1,ψε has degree 2 and K2,ψε has degree 1. In order to avoid this case, we need to be sure that if −τ α2− F(2)+ A(2) ≥ 0, then τ α1+ F(1)− A(1) < 0, that is,

τ α2+ F(2)− A(2) ≤ 0 ⇒ τ α1+ F(1)− A(1) < 0.

This can be achieved by assuming that α2

α1 > A(2)− F(2)

A(1)− F(1), ∀ ξ0 6= 0. (3.22) Recall that A(k)− F(k)> 0, k = 1, 2. We remark that all A(k) and F(k) are homoge- neous of degree 1 in ξ0. Hence (3.22) holds provided

α2 α1

= max

0|=1

 A(2)− F(2) A(1)− F(1)



+ 1. (3.23)

(12)

Hence, if we assume (3.23), then the transmission condition is satisfied.

Case 3. Each symbol has exactly one root in {Im z < 0}, i.e., τ α1+ F(1)− A(1) < 0, −τ α2− F(2)+ A(2) > 0.

In this case, we have

K1,ψε = (λ − σ1(1)), K2,ψε = (λ − σ2(2)).

Given polynomials q1(λ), q2(λ), there exist U1(λ), U2(λ) such that q1(λ) = U1(λ)K1,ψε + ˜q1,

q2(λ) = U2(λ)K2,ψε + ˜q2,

where ˜q1, ˜q2 are constants in λ. The transmission condition is satisfied if there exists constants µ1, µ2, c1, c2 so that

(q˜1 = µ1K1,ψε + c1˜t11,ψε + c221,ψε,

˜

q2 = µ2K2,ψε + c1˜t12,ψε + c222,ψε, namely,

(q˜1 = µ1(λ − σ1(1)) − c1+ c2a(1)nn(λ − iτ α1− E(1)− iF(1))

˜

q2 = µ2(λ − σ2(2)) + c1+ c2a(2)nn(λ + iτ α2+ E(2)+ iF(2)). (3.24) System (3.24) is equivalent to









µ1+ c2a(1)nn = 0 µ2+ c2a(2)nn = 0

µ1σ1(1)+ c1+ c2a(1)nn(iτ α1 + E(1)+ iF(1)) = −˜q1

− µ2σ2(2)+ c1+ c2a(2)nn(iτ α2+ E(2)+ iF(2)) = ˜q2.

(3.25)

System (3.25) has a unique solution if and only if the matrix

T =

1 0 0 a(1)nn

0 1 0 a(2)nn

σ1(1) 0 1 ζ1

0 −σ2(2) 1 ζ2

with ζ1 = a(1)nn(iτ α1+ E(1)+ iF(1)), ζ2 = a(2)nn(iτ α2+ E(2)+ iF(2)), is nonsingular. We

(13)

compute

detT =det

1 0 a(1)nn

0 1 a(2)nn

0 −σ(2)2 ζ2

− det

1 0 a(1)nn

0 1 a(2)nn

σ1(1) 0 ζ1

2+ σ2(2)a(2)nn − ζ1+ σ1(1)a(1)nn

=a(2)nn(iτ α2+ E(2)+ iF(2)− E(2)+ B(2)− iτ α2 − iF(2)+ iA(2)) + a(1)nn(−iτ α1− E(1)− iF(1)+ E(1)+ B(1)+ iτ α1+ iF(1)+ iA(1))

=a(2)nn(B(2)+ iA(2)) + a(1)nn(B(1)+ iA(1)).

Therefore, if

a(2)nn(B(2)+ iA(2)) + a(1)nn(B(1)+ iA(1)) 6= 0, (3.26) then the transmission condition holds.

We now verify (3.26). In the real case where a(2)nn, a(1)nn are positive real numbers, it is easy to see that

a(2)nnA(2)+ a(1)nnA(1) > 0 and thus (3.26) holds.

For the complex case, we want to show that there exists γ0 > 0 such that if γ < γ0, then (3.26) is satisfied. Let uk= A(k)+ iB(k) and vk = iuk = −B(k)+ iA(k). We will consider uk and vk as vectors in R2, i.e., uk = (A(k), B(k)), vk = uk = (−B(k), A(k)).

Let a(k)nn = η(k)+ iγδ(k) for η(k), δ(k) ∈ R. By the ellipticity conditions (3.3), (3.4), we have

λ0 ≤ η(k) ≤ Λ0, λ0 ≤ δ(k) ≤ Λ0. Notice that detT = 0 if and only if

(2)+ iγδ(2))(B(2)+ iA(2)) + (η(1)+ iγδ(1))(B(1)+ iA(1)) = 0, i.e.,

(2)B(2)− γδ(2)A(2)+ η(1)B(1)− γδ(1)A(1))

+ i(η(2)A(2)+ γδ(2)B(2)+ η(1)A(1)+ γδ(1)B(1)) = 0, which is equivalent to

η(2)A(2) B(2)



+ η(1)A(1) B(1)



= γδ(2)−B(2) A(2)



+ γδ(1)−B(1) A(1)



(3.27) or simply

η(2)u2+ η(1)u1 = γδ(2)v2+ γδ(1)v1. (3.28) Recall that A(k) ≥ |F(k)| > 0. Therefore, in the real case γδ(k) = 0, then (3.27) will never be satisfied. If B(1) and B(2) have the same sign, that is, either B(k) ≥ 0 or B(k) ≤ 0 for k = 1, 2, (3.28) can not hold. To see this, let us consider B(k) ≥ 0,

(14)

k = 1, 2. Then u1, u2are in the first quadrant of the plane and v1, v2 are in the second quadrant of the plane. The sets

Cu = {η(2)u2+ η(1)u1 : η(k) ≥ 0}, Cu = {γδ(2)v2+ γδ(1)v1 : γδ(k)≥ 0}

can only intersect at the original. Same thing happens if B(k) ≤ 0 for k = 1, 2.

The only case we need to investigate is when B(1) and B(2) have different signs.

For example, let us assume

B(1) > 0, B(2) < 0.

Even in this case, the intersection between Cu and Cv is non-trivial if the angle φ between u1 and u2 is less than π/2. Note that u1 is the first quadrant and u2 is in the fourth quadrant. So the angle between u1 and u2 is less than π. We would like to show that (3.28) cannot hold for φ ∈ [π/2, π) if we choose γ0 small enough.

Note that in this case cos φ ≤ 0. To do so, we estimate kη(2)u2(1)u1k from below and kδ(2)v2 + δ(1)v1k from above. We now discuss the estimate of kδ(2)v2 + δ(1)v1k from above. Compute

(2)v2+ δ(1)v1k2 = (δ(2))2[(A(2))2+ (B(2))2] + (δ(1))2[(A(1))2+ (B(1))2] + 2δ(1)δ(2)(−B(2), A(2)) · (−B(1), A(1))

= (δ(2))2[(A(2))2+ (B(2))2] + (δ(1))2[(A(1))2+ (B(1))2] + 2δ(1)δ(2)[(A(2))2+ (B(2))2]1/2[(A(1))2 + (B(1))2]1/2cos φ

≤ (δ(2))2[(A(2))2+ (B(2))2] + (δ(1))2[(A(1))2+ (B(1))2].

(3.29)

In view of (3.9) and (3.11), we have (A(k))2+ (B(k))2 = |bk| = | X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj− (E(k)+ iF(k))2|

≤ | X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj| + |(E(k)+ iF(k))2|.

(3.30)

By (3.3), (3.4), and (3.12), we can obtain

| X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj|2 = | 1 a(k)nn

A(k)ξ · ξ|2 (with ξ = (ξ0, 0))

= |a(k)nn|−2| cos θM(k)ξ · ξ + γ sin θN(k)ξ · ξ + i(− sin θM(k)ξ · ξ + γ cos θN(k)ξ · ξ)|2

= |a(k)nn|−2[(M(k)ξ · ξ)2+ γ2(N(k)ξ · ξ)2]

≤ Λ2(1 + γ2)|ξ|4

λ20(1 + γ2) = ˜λ−11 |ξ|4, where we have used the estimate

λ0(1 + γ2)1/2≤ |a(k)nn| ≤ Λ0(1 + γ2)1/2 (3.31)

(15)

in deriving the inequality above. We thus obtain

| X

1≤`,j≤n−1

a(k)`j a(k)nn

ξ`ξj| ≤ ˜λ−1/210|2. (3.32)

Furthermore, we can estimate

|(E(k)+ iF(k))2| = |( X

1≤j≤n−1

a(k)nj a(k)nn

ξj)2| = | X

1≤j≤n−1

a(k)nj a(k)nn

ξj|2

≤ ( X

1≤j≤n−1

|a(k)nj a(k)nn

|2)|ξ0|2 ≤ (n − 1)Λ20(1 + γ2) λ20(1 + γ2) |ξ0|2

= (n − 1)˜λ−110|2.

(3.33)

Substituting (3.32), (3.33) into (3.30) gives

(A(k))2+ (B(k))2 ≤ (˜λ−1/21 + (n − 1)˜λ−11 )|ξ0|2 ≤ nΛ20

λ200|2. (3.34) It follows from (3.29) and (3.34) that

(2)v2+ δ(1)v1k2 ≤ 2Λ2020

λ200|2. (3.35) Next, we want to estimate kη(2)u2+ η(1)u1k from below. As above, we have kη(2)u2+ η(1)u1k2 = (η(2))2[(A(2))2+ (B(2))2] + (η(1))2[(A(1))2+ (B(1))2]

+ 2η(1)η(2)[(A(2))2+ (B(2))2]1/2[(A(1))2+ (B(1))2]1/2cos φ. (3.36) Recall that B1 > 0, B2 < 0. Thus,

cos φ = A(1)A(2)+ B(1)B(2)

[(A(2))2+ (B(2))2]1/2[(A(1))2+ (B(1))2]1/2

= A(1)A(2)− |B(1)||B(2)|

[(A(2))2+ (B(2))2]1/2[(A(1))2+ (B(1))2]1/2

= 1 −|BA(1)(1)|

|B(2)| A(2)

(1 + (BA(2)(2))2)1/2(1 + (BA(1)(1))2)1/2. Notice that by (3.19) and (3.35)

0 ≤ |B(k)| A(k)

p(A(k))2+ (B(k))2

A(k)

√nΛλ0

00| p˜λ10| =

√nΛ20

λ20 := ˜λ2 ≥ 1.

參考文獻

相關文件

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

Abstract Based on a class of smoothing approximations to projection function onto second-order cone, an approximate lower order penalty approach for solving second-order cone

Based on a class of smoothing approximations to projection function onto second-order cone, an approximate lower order penalty approach for solving second-order cone

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,

For the proposed algorithm, we establish a global convergence estimate in terms of the objective value, and moreover present a dual application to the standard SCLP, which leads to

Abstract In this paper, we consider the smoothing Newton method for solving a type of absolute value equations associated with second order cone (SOCAVE for short), which.. 1

We propose a primal-dual continuation approach for the capacitated multi- facility Weber problem (CMFWP) based on its nonlinear second-order cone program (SOCP) reformulation.. The

Abstract We investigate some properties related to the generalized Newton method for the Fischer-Burmeister (FB) function over second-order cones, which allows us to reformulate