• 沒有找到結果。

Part I On the Numerical Solutions of Linear Systems

N/A
N/A
Protected

Academic year: 2022

Share "Part I On the Numerical Solutions of Linear Systems"

Copied!
22
0
0

加載中.... (立即查看全文)

全文

(1)

Part I

On the Numerical Solutions of

Linear Systems

(2)
(3)

Chapter 1 Introduction

1.1 Mathematical auxiliary, definitions and relations

1.1.1 Vectors and matrices

A∈ Km×n, where K = R or C ⇔ A = [aij] =



a11 · · · a1n

... ... ...

am1 · · · amn

.

• Product of matrices (Km×n× Kn×p → Km×p): C = AB, where cij = Pn

k=1aikbkj, i = 1,· · · , m, j = 1, · · · , p.

• Transpose (Rm×n→ Rn×m): C = AT, where cij = aji∈ R.

• Conjugate transpose (Cm×n→ Cn×m): C = A or C = AH, where cij = ¯aji ∈ C.

• Differentiation (Rm×n→ Rm×n): Let C(t) = (cij(t)). Then ˙C(t) = [ ˙cij(t)].

• If A, B ∈ Kn×n satisfy AB = I, then B is the inverse of A and is denoted by A−1. If A−1 exists, then A is said to be nonsingular; otherwise, A is singular. A is nonsingular if and only if det(A)6= 0.

• If A ∈ Km×n, x∈ Kn and y = Ax, then yi =Pn

j=1aijxj, i = 1,· · · , m.

• Outer product of x ∈ Km and y∈ Kn:

xy =



x11 · · · x1n

... . .. ...

xm1 · · · xmn

 ∈ Km×n.

• Inner product of x and y ∈ Kn: (x, y) := xTy =

Xn i=1

xiyi = yTx∈ R

(x, y) := xy = Xn

i=1

¯

xiyi = yx∈ C

(4)

• Sherman-Morrison Formula:

Let A∈ Rn×n be nonsingular, u, v∈ Rn. If vTA−1u6= −1, then

(A + uvT)−1 = A−1− (1 + vTA−1u)−1A−1uvTA−1. (1.1.1)

• Sherman-Morrison-Woodburg Formula:

Let A∈ Rn×n, be nonsingular U , V ∈ Rn×k. If (I + VTA−1U ) is invertible, then (A + U VT)−1 = A−1− A−1U (I + VTA−1U )−1VTA−1,

Proof of (1.1.1):

(A + uvT)[A−1− A−1uvTA−1/(1 + vTA−1u)]

= I + 1

1 + vTA−1u[uvTA−1(1 + vTA−1u)− uvTA−1− uvTA−1uvTA−1]

= I + 1

1 + vTA−1u[u(vTA−1u)vTA−1− uvTA−1uvTA−1] = I.

Example 1.1.1

A =





3 −1 1 1 1

0 1 2 2 2

0 −1 4 1 1

0 0 0 3 0

0 0 0 0 3





= B +





 0 0

−1 0 0





 0 1 0 0 0  .

1.1.2 Rank and orthogonality

Let A∈ Rm×n. Then

• R(A) = {y ∈ Rm| y = Ax for some x ∈ Rn } ⊆ Rm is the range space of A.

• N (A) = {x ∈ Rn| Ax = 0 } ⊆ Rn is the null space of A.

• rank(A) = dim [R(A)] = The number of maximal linearly independent columns of A.

• rank(A) = rank(AT).

• dim(N (A)) + rank(A) = n.

• If m = n, then A is nonsingular ⇔ N (A) = {0} ⇔ rank(A) = n.

• Let {x1,· · · , xp} ⊆ Rn. Then {x1,· · · , xp} is said to be orthogonal if xTi xj = 0, for i6= j and orthonormal if xTi xj = δij, where δij = 0 if i 6= j and δij = 1 if i = j.

• S={y ∈ Rm | yTx = 0, for x∈ S} = orthogonal complement of S.

• Rn=R(AT)⊕ N (A), Rm=R(A) ⊕ N (AT).

• R(AT)⊥ N (A), R(A)=N (AT).

(5)

1.1 Mathematical auxiliary, definitions and relations 5

A∈ Rn×n A∈ Cn×n

Symmetric: AT = A Hermitian: A = A(AH = A) skew-symmetric: AT =−A skew-Hermitian: A =−A

positive definite: xTAx > 0, x6= 0 positive definite: xAx > 0, x6= 0 non-negative definite: xTAx≥ 0 non-negative definite: xAx ≥ 0

indefinite: (xTAx)(yTAy) < 0 for some x, y indefinite: (xAx)(yAy) < 0 for some x, y orthogonal: ATA = In unitary: AA = In

normal: ATA = AAT normal: AA = AA positive: aij > 0

non-negative: aij ≥ 0.

Table 1.1: Some definitions for matrices.

1.1.3 Special matrices

Let A∈ Kn×n. Then the matrix A is

• diagonal if aij = 0, for i6= j. Denote D = diag(d1,· · · , dn)∈ Dnthe set of diagonal matrices;

• tridiagonal if aij = 0,|i − j| > 1;

• upper bi-diagonal if aij= 0, i > j or j > i + 1;

• (strictly) upper triangular if aij = 0, i > j (i≥ j);

• upper Hessenberg if aij = 0, i > j + 1.

(Note: the lower case is the same as above.)

Sparse matrix: n1+r, where r < 1 (usually between 0.2∼ 0.5). If n = 1000, r = 0.9, then n1+r = 501187.

Example 1.1.2 If S is skew-symmetric, then I− S is nonsingular and (I − S)−1(I + S) is orthogonal (Cayley transformation of S).

1.1.4 Eigenvalues and Eigenvectors

Definition 1.1.1 Let A∈ Cn×n. Then λ∈ C is called an eigenvalue of A, if there exists x6= 0, x ∈ Cn with Ax = λx and x is called an eigenvector corresponding to λ.

Notations:

σ(A) := Spectrum of A = The set of eigenvalues of A.

ρ(A) := Radius of A = max{|λ| : λ ∈ σ(A)}.

• λ ∈ σ(A) ⇔ det(A− λI) = 0.

• p(λ) = det(λI − A) = The characteristic polynomial of A.

• p(λ) =Qs

i=1(λ− λi)m(λi), where λi 6= λj (for i6= j) and Ps

i=1m(λi) = n.

(6)

• m(λi) = The algebraic multiplicity of λi.

• n(λi) = n− rank(A − λiI) = The geometric multiplicity of λi.

• 1 ≤ n(λi)≤ m(λi).

Definition 1.1.2 If there is some i such that n(λi) < m(λi), then A is called degenerated.

The following statements are equivalent:

(1) There are n linearly independent eigenvectors;

(2) A is diagonalizable, i.e., there is a nonsingular matrix T such that T−1AT ∈ Dn; (3) For each λ∈ σ(A), it holds that m(λ) = n(λ).

If A is degenerated, then eigenvectors and principal vectors derive the Jordan form of A.

(See Gantmacher: Matrix Theory I, II)

Theorem 1.1.1 (Schur) (1) Let A ∈ Cn×n. There is a unitary matrix U such that UAU (= U−1AU ) is upper triangular.

(2) Let A ∈ Rn×n. There is an orthogonal matrix Q such that QTAQ(= Q−1AQ) is quasi-upper triangular, i.e., an upper triangular matrix possibly with nonzero subdiagonal elements in non-consecutive positions.

(3) A is normal if and only if there is a unitary U such that UAU = D diagonal.

(4) A is Hermitian if and only if A is normal and σ(A)⊆ R.

(5) A is symmetric if and only if there is an orthogonal U such that UTAU = D diagonal and σ(A)⊆ R.

1.2 Norms and eigenvalues

Let X be a vector space over K = R or C.

Definition 1.2.1 (Vector norms) Let N be a real-valued function defined on X (N : X → R+). Then N is a (vector) norm, if

N1: N (αx) =|α|N(x), α ∈ K, for x ∈ X;

N2: N (x + y)≤ N(x) + N(y), for x, y ∈ X;

N3: N (x) = 0 if and only if x = 0.

The usual notation is kxk = N(x).

(7)

1.2 Norms and eigenvalues 7 Example 1.2.1 Let X = Cn, p ≥ 1. Then kxkp = (Pn

i=1|xi|p)1/p is a p-norm. Espe- cially,

kxk1 = Xn

i=1

|xi| ( 1-norm),

kxk2 = ( Xn

i=1

|xi|2)1/2 (2-norm = Euclidean-norm), kxk= max

1≤i≤n|xi| (∞-norm = maximum norm).

Lemma 1.2.1 N (x) is a continuous function in the components x1,· · · , xn of x.

Proof:

|N(x) − N(y)| ≤ N(x − y) ≤ Xn

j=1

|xj − yj|N(ej)

≤ kx − yk

Xn j=1

N (ej).

Theorem 1.2.1 (Equivalence of norms) Let N and M be two norms on Cn. Then there are constants c1, c2 > 0 such that

c1M (x)≤ N(x) ≤ c2M (x), for all x ∈ Cn.

Proof: Without loss of generality (W.L.O.G.) we can assume that M (x) =kxkand N is arbitrary. We claim that

c1kxk≤ N(x) ≤ c2kxk, equivalently,

c1 ≤ N(z) ≤ c2,∀ z ∈ S = {z ∈ Cn|kzk= 1}.

From Lemma 1.2.1, N is continuous on S (closed and bounded). By maximum and minimum principle, there are c1, c2 ≥ 0 and z1, z2 ∈ S such that

c1 = N (z1)≤ N(z) ≤ N(z2) = c2.

If c1 = 0, then N (z1) = 0, and thus, z1 = 0. This contradicts that z1 ∈ S.

Remark 1.2.1 Theorem 1.2.1 does not hold in infinite dimensional space.

Definition 1.2.2 (Matrix-norms) Let A∈ Cm×n. A real-valued functionk·k : Cm×n→ R+ satisfying

N1: kαAk = |α|kAk;

N2: kA + Bk ≤ kAk + kBk ;

(8)

N3: kAk = 0 if and only if A = 0;

N4: kABk ≤ kAkkBk ;

N5: kAxkv ≤ kAkkxkv (matrix and vector norms are compatible for some k · kv) is called a matrix norm. If k · k satisfies N1 to N4, then it is called a multiplicative or algebra norm.

Example 1.2.2 (Frobenius norm) Let kAkF = [Pn

i,j=1|ai,j|2]1/2. kABkF = (X

i,j

|X

k

aikbkj|2)12

≤ (X

i,j

{X

k

|aik|2}{X

k

|bkj|2})12 (Cauchy-Schwartz Ineq.)

= (X

i

X

k

|aik|2)12(X

j

X

k

|bkj|2)12

= kAkFkBkF. (1.2.1)

This implies that N4 holds. Furthermore, by Cauchy-Schwartz inequality we have kAxk2 = (X

i

|X

j

aijxj|2)12

≤ X

i

(X

j

|aij|2)(X

j

|xj|2)

!12

= kAkFkxk2. (1.2.2)

This implies that N5 holds. Also, N1, N2 and N3 hold obviously. (Here, kIkF =√ n).

Example 1.2.3 (Operator norm) Given a vector norm k · k. An associated (induced) matrix norm is defined by

kAk = sup

x6=0

kAxk

kxk = max

x6=0

kAxk

kxk . (1.2.3)

Then N5 holds immediately. On the other hand,

k(AB)xk = kA(Bx)k ≤ kAkkBxk

≤ kAkkBkkxk (1.2.4)

for all x6= 0. This implies that

kABk ≤ kAkkBk. (1.2.5)

It holds N 4. (Here kIk = 1).

(9)

1.2 Norms and eigenvalues 9 In the following, we represent and verify three useful matrix norms:

kAk1 = sup

x6=0

kAxk1

kxk1 = max

1≤j≤n

Xn i=1

|aij| (1.2.6)

kAk= sup

x6=0

kAxk

kxk

= max

1≤i≤n

Xn j=1

|aij| (1.2.7)

kAk2 = sup

x6=0

kAxk2

kxk2

=p

ρ(AA) (1.2.8)

Proof of (1.2.6):

kAxk1 = X

i

|X

j

aijxj| ≤X

i

X

j

|aij||xj|

= X

j

|xj|X

i

|aij|.

Let C1 := P

i|aik| = maxjP

i|aij|. Then kAxk1 ≤ C1kxk1, thus kAk1 ≤ C1. On the other hand, kekk1 = 1 and kAekk1 =Pn

i=1|aik| = C1. Proof of (1.2.7):

kAxk = max

i |X

j

aijxj|

≤ maxi X

j

|aijxj|

≤ max

i

X

j

|aij|kxk

≡ X

j

|akj|kxk

≡ Ckxk.

This implies that kAk≤ C. If A = 0, there is nothing to prove. Assume that A6= 0 and the k-th row of A is nonzero. Define z = [zj]∈ Cn by

zj = ( ¯a

kj

|akj| if akj 6= 0, 1 if akj = 0.

Then kzk= 1 and akjzj =|akj|, for j = 1, . . . , n. It follows that kAk≥ kAzk = max

i |X

j

aijzj| ≥ |X

j

akjzj| = Xn

j=1

|akj| ≡ C. Thus, kAk ≥ max1≤i≤nPn

j=1|aij| ≡ C.

Proof of (1.2.8): Let λ1 ≥ λ2 ≥ · · · ≥ λn ≥ 0 be the eigenvalues of AA. There are mutually orthonormal vectors vj, j = 1, . . . , n such that (AA)vj = λjvj. Let x = P

jαjvj. Since kAxk22 = (Ax, Ax) = (x, AAx), kAxk22 = X

j

αjvj,X

j

αjλjvj

!

=X

j

λjj|2 ≤ λ1kxk22.

(10)

Therefore, kAk22 ≤ λ1. Equality follows by choosing x = v1 and kAv1k22 = (v1, λ1v1) = λ1. So, we have kAk2 =p

ρ(AA).

Example 1.2.4 (Dual norm) Let 1p +1q = 1. Then k · kp =k · kq, (p =∞, q = 1). (It concludes from the application of the H¨older inequality that |yx| ≤ kxkpkykq.)

Theorem 1.2.2 Let A∈ Cn×n. Then for any operator norm k · k, it holds ρ(A)≤ kAk.

Moreover, for any ε > 0, there exists an operator norm k · kε such that k · kε ≤ ρ(A) + ε.

Proof: Let |λ| = ρ(A) ≡ ρ and x be the associated eigenvector with kxk = 1. Then, ρ(A) =|λ| = kλxk = kAxk ≤ kAkkxk = kAk.

On the other hand, there is a unitary matrix U such that A = URU , where R is upper triangular. Let Dt = diag(t, t2, . . . , tn). Compute

DtRD−1t =







λ1 t−1r12 t−2r13 · · · t−n+1r1n

λ2 t−1r23 · · · t−n+2r2n

λ3 ...

. .. t−1rn−1,n λn





 .

For t > 0 sufficiently large, the sum of all absolute values of the off-diagonal elements of DtRDt−1 is less than ε. So, it holdskDtRDt−1k1 ≤ ρ(A) + ε for sufficiently large t(ε) > 0.

Define k · kε for any B by

kBkε = kDtU BUD−1t k1

= k(UD−1t )−1B(U Dt−1)k1. This implies that

kAkε =kDtRDt−1k ≤ ρ(A) + ε.

Remark 1.2.2

kUAV kF = kAkF (by kUAkF = q

kUa1k22+· · · + kUank22), (1.2.9) kUAV k2 = kAk2 (by ρ(AA) = ρ(AA)), (1.2.10) where U and V are unitary.

Theorem 1.2.3 (Singular Value Decomposition (SVD)) Let A∈ Cm×n. Then there exist unitary matrices U = [u1,· · · , um]∈ Cm×m and V = [v1,· · · , vn]∈ Cn×n such that

UAV = diag(σ1,· · · , σp) = Σ,

where p = min{m, n} and σ1 ≥ σ2 ≥ · · · ≥ σp ≥ 0. (Here, σi denotes the i-th largest singular value of A).

(11)

1.2 Norms and eigenvalues 11 Proof: There are x ∈ Cn, y ∈ Cm with kxk2 = kyk2 = 1 such that Ax = σy, where σ = kAk2 (kAk2 = supkxk2=1kAxk2). Let V = [x, V1] ∈ Cn×n, and U = [y, U1] ∈ Cm×m be unitary. Then

A1 ≡ UAV =

 σ w 0 B

 . Since

A1

 σ w



2 2

≥ (σ2+ ww)2, it follows that

kA1k22 ≥ σ2+ ww from A1

 σ w



2

2

 σ w



2 2

≥ σ2+ ww.

But σ2 =kAk22 =kA1k22, it implies w = 0. Hence, the theorem holds by induction.

Remark 1.2.3 kAk2 =p

ρ(AA) = σ1 = The maximal singular value of A.

Let A = U ΣV. Then we have

kABCkF = kUΣVBCkF =kΣVBCkF

≤ σ1kBCkF =kAk2kBCkF. This implies

kABCkF ≤ kAk2kBkFkCk2. (1.2.11) In addition, by (1.2.2) and (1.2.11), we get

kAk2 ≤ kAkF ≤√

nkAk2. (1.2.12)

Theorem 1.2.4 Let A∈ Cn×n. The following statements are equivalent:

(1) lim

m→∞Am= 0;

(2) lim

m→∞Amx = 0 for all x;

(3) ρ(A) < 1.

Proof: (1) ⇒ (2): Trivial. (2) ⇒ (3): Let λ ∈ σ(A), i.e., Ax = λx, x 6= 0. This implies Amx = λmx → 0, as λm → 0. Thus |λ| < 1, i.e., ρ(A) < 1. (3) ⇒ (1): There is a norm k · k with kAk < 1 (by Theorem 1.2.2). Therefore, kAmk ≤ kAkm → 0, i.e., Am → 0.

Theorem 1.2.5 It holds that

ρ(A) = lim

k→∞kAkk1/k where k k is an operator norm.

(12)

Proof: Since

ρ(A)k = ρ(Ak)≤ kAkk ⇒ ρ(A) ≤ kAkk1/k,

for k = 1, 2, . . .. If ε > 0, then ˜A = [ρ(A) + ε]−1A has spectral radius < 1 and by Theorem 1.2.4 we have k ˜Akk → 0 as k → ∞. There is an N = N(ε, A) such that k ˜Akk < 1 for all k ≥ N. Thus, kAkk ≤ [ρ(A) + ε]k, for all k≥ N or kAkk1/k ≤ ρ(A) + ε for all k ≥ N. Since ρ(A) ≤ kAkk1/k, and k, ε are arbitrary, limk→∞kAkk1/k exists and equals ρ(A).

Theorem 1.2.6 Let A∈ Cn×n, and ρ(A) < 1. Then (I− A)−1 exists and (I− A)−1 = I + A + A2+· · · .

Proof: Since ρ(A) < 1, the eigenvalues of (I − A) are nonzero. Therefore, by Theorem 2.5, (I− A)−1 exists and

(I− A)(I + A + A2+· · · + Am) = I − Am → 0.

Corollary 1.2.1 If kAk < 1, then (I − A)−1 exists and k(I − A)−1k ≤ 1

1− kAk Proof: Since ρ(A)≤ kAk < 1 (by Theorem 1.2.2),

k(I − A)−1k = k X

i=0

Aik ≤ X

i=0

kAki = (1− kAk)−1.

Theorem 1.2.7 (Without proof ) For A∈ Kn×n the following statements are equivalent:

(1) There is a multiplicative norm p with p(Ak)≤ 1, k = 1, 2, . . ..

(2) For each multiplicative norm p the power p(Ak) are uniformly bounded, i.e., there exists a M (p) <∞ such that p(Ak)≤ M(p), k = 0, 1, 2, . . ..

(3) ρ(A)≤ 1 and all eigenvalue λ with |λ| = 1 are not degenerated. (i.e., m(λ) = n(λ).) (See Householder’s book: The theory of matrix, pp.45-47.)

In the following we prove some important inequalities of vector norms and matrix norms.

(13)

1.2 Norms and eigenvalues 13 (a) It holds that

1≤ kxkp

kxkq ≤ n(q−p)/pq, (p≤ q). (1.2.13) Proof: Claim kxkq ≤ kxkp, (p≤ q): It holds

kxkq = kxkp

x kxkp

q

=kxkp

x kxkp

q

≤ Cp,qkxkp,

where

Cp,q = max

kekp=1kekq, e = (e1,· · · , en)T. We now show that Cp,q≤ 1. From p ≤ q, we have

kekqq = Xn

i=1

|ei|q ≤ Xn

i=1

|ei|p = 1 (by |ei| ≤ 1).

Hence, Cp,q ≤ 1, thus kxkq ≤ kxkp.

To prove the second inequality: Let α = q/p > 1. Then the Jensen inequality holds for the convex function ϕ(x):

ϕ(

Z

f dµ)≤ Z

(ϕ◦ f)dµ, µ(Ω) = 1.

If we take ϕ(x) = xα, then we have Z

|f|qdx = Z

(|f|p)q/pdx≥

Z

|f|pdx

q/p

with |Ω| = 1. Consider the discrete measure Pn i=11

n = 1 and f (i) =|xi|. It follows that

Xn i=1

|xi|q1 n ≥

Xn i=1

|xi|p1 n

!q/p

. Hence, we have

n1qkxkq ≥ n1pkxkp. Thus,

n(q−p)/pqkxkq ≥ kxkp. (b) It holds that

1≤ kxkp

kxk ≤ n1p. (1.2.14)

Proof: Let q→ ∞ and lim

q→∞kxkq =kxk:

kxk =|xk| = (|xk|q)1q ≤ Xn

i=1

|xi|q

!1q

=kxkq.

(14)

On the other hand, we have

kxkq = Xn

i=1

|xi|q

!1q

≤ (nkxkq)1q ≤ n1qkxk

which implies that limq→∞kxkq =kxk. (c) It holds that

1≤j≤nmax kajkp ≤ kAkp ≤ n(p−1)/p max

1≤j≤nkajkp, (1.2.15) where A = [a1,· · · , an]∈ Rm×n.

Proof: The first inequality holds obviously. To show the second inequality, for kykp = 1 we have

kAykp ≤ Xn j=1

|yj|kajkp ≤ Xn

j=1

|yj| max

j kajkp

= kyk1max

j kajkp ≤ n(p−1)/pmax

j kajkp (by (1.2.13)).

(d) It holds that

maxi,j |aij| ≤ kAkp ≤ n(p−1)/pm1/pmax

i,j |aij|, (1.2.16) where A∈ Rm×n.

Proof: By (1.2.14) and (1.2.15) immediately.

(e) It holds that

m(1−p)/pkAk1 ≤ kAkp ≤ n(p−1)/pkAk1. (1.2.17) Proof: By (1.2.15) and (1.2.13) immediately.

(f ) By H¨older inequality, we have (see Appendix later!)

|yx| ≤ kxkpkykq, where 1p +1q = 1 or

max{|xy| : kykq= 1} = kxkp. (1.2.18) Then it holds that

kAkp =kATkq. (1.2.19)

Proof: By (1.2.18) we have

kxkmaxp=1kAxkp = max

kxkp=1 max

kykq=1|(Ax)Ty|

= max

kykq=1 max

kxkp=1|xT(ATy)| = max

kykq=1kATykq =kATkq.

(15)

1.2 Norms and eigenvalues 15 (g) It holds that

n1pkAk≤ kAkp ≤ m1pkAk. (1.2.20) Proof: By (1.2.17) and (1.2.19), we get

m1pkAk = m1pkATk1 = m1−1qkATk1

= m(q−1)/qkATk1 ≥ kATkq =kAkp. (h) It holds that

kAk2 ≤q

kAkpkAkq, (1 p +1

q = 1). (1.2.21)

Proof: By (1.2.19) we have

kAkpkAkq =kATkqkAkq ≥ kATAkq ≥ kATAk2.

The last inequality holds by the following statement: Let S be a symmetric matrix.

Then kSk2 ≤ kSk, for any matrix operator norm k k. Since |λ| ≤ kSk, kSk2 =p

ρ(SS) = p

ρ(S2) = max

λ∈σ(S)|λ| = |λmax|.

This implies, kSk2 ≤ kSk.

(i) For A∈ Rm×n and q ≥ p ≥ 1, it holds that

n(p−q)/pqkAkq ≤ kAkp ≤ m(q−p)/pqkAkq. (1.2.22) Proof: By (1.2.13), we get

kAkp = max

kxkp=1kAxkp ≤ max

kxkq≤1m(q−p)/pqkAxkq

= m(q−p)/pqkAkq.

Appendix: To show H¨ older inequality and (1.2.18)

Taking ϕ(x) = ex in Jensen’s inequality we have exp

Z

f dµ



≤ Z

efdµ.

Let Ω = finite set ={p1, . . . , pn}, µ({pi}) = n1, f (pi) = xi. Then exp

1

n(x1+· · · + xn)



≤ 1

n(ex1 +· · · + exn) . Taking yi = exi, we have

(y1· · · yn)1/n ≤ 1

n(y1 +· · · + yn).

(16)

Taking µ({pi}) = qi > 0, Pn

i=1qi = 1 we have

y1q1· · · yqnn ≤ q1y1+· · · + qnyn. (1.2.23) Let αi = xi/kxkp, βi = yi/kykq, where x = [x1,· · · , xn]T, y = [y1,· · · , yn]T, α = [α1,· · · , αn]T and β = [β1,· · · , βn]T. By (1.2.23) we have

αiβi ≤ 1

pi + 1 qβiq. Since kαkp = 1, kβkq = 1, it holds

Xn i=1

αiβi ≤ 1 p +1

q = 1.

Thus,

|xTy| ≤ kxkpkykq.

To show max{|xTy|; kxkp = 1} = kykq. Taking xi = yiq−1/kykq/pq we have kxkpp =

Pn

i=1|yi|(q−1)p kykqq = 1.

Note (q− 1)p = q. Then

Xn i=1

xTi yi

=

Pn i=1|yi|q

kykq/pq = kykqq

kykq/pq =kykq. The following two properties are useful in the following sections.

(i) There exists ˆz with kˆzkp = 1 such that kykq = ˆzTy. Let z = ˆz/kykq. Then we have zTy = 1 andkzkp = kyk1q.

(ii) From the duality, we have kyk = (kyk) = maxkuk=1|yTu| = yTz andˆ kˆzk = 1.

Let z = ˆz/kyk. Then we have zTy = 1 and kzk = kyk1 .

1.3 The Sensitivity of Linear System Ax = b

1.3.1 Backward error and Forward error

Let x = F (a). We define backward and forward errors in Figure 1.1. In Figure 1.1, ˆ

x + ∆x = F (a + ∆a) is called a mixed forward-backward error, where |∆x| ≤ ε|x|,

|∆a| ≤ η|a|.

Definition 1.3.1 (i) An algorithm is backward stable, if for all a, it produces a computed ˆ

x with a small backward error, i.e., ˆx = F (a + ∆a) with ∆a small.

(ii) An algorithm is numerical stable, if it is stable in the mixed forward-backward error sense, i.e., ˆx + ∆x = F (a + ∆a) with both ∆a and ∆x small.

(17)

1.3 The Sensitivity of Linear System Ax = b 17

 

  

  

 

   

 



 

  

 

Figure 1.1: Relationship between backward and forward errors.

(iii) If a method which produces answers with forward errors of similar magnitude to those produced by a backward stable method, is called a forward stable.

Remark 1.3.1 (i) Backward stable ⇒ forward stable, not vice versa!

(ii) Forward error ≤ condition number × backward error Consider

ˆ

x− x = F (a + ∆a) − F (a) = F0(a)∆a + F00(a + θ∆a)

2 (∆a)2, θ∈ (0, 1).

Then we have

ˆ x− x

x =

aF0(a) F (a)

∆a

a + O (∆a)2 .

The quantity C(a) = aFF (a)0(a)

is called the condition number of F. If x or F is a vector, then the condition number is defined in a similar way using norms and it measures the maximum relative change, which is attained for some, but not all ∆a.

 `Apriori error estimate ! Pposteriori error estimate !`

1.3.2 An SVD Analysis

Let A =Pn

i=1σiuiviT = U ΣVT be a singular value decomposition (SVD) of A. Then x = A−1b = (U ΣVT)−1b =

Xn i=1

uiTb σi

vi.

If cos(θ) =| unTb | / k b k2 and

(A− εunvnT)y = b + ε(unTb)un, σn > ε≥ 0.

(18)

Then we have

k y − x k2≥ ( ε

σn)k x k2 cos(θ).

Let E = diag{0, · · · , 0, ε}. Then it holds

(Σ− E)VTy = UTb + ε(unTb)en. Therefore,

y− x = V (Σ − E)−1UTb + ε(unTb)(σn− ε)−1vn− V Σ−1UTb

= V ((Σ− E)−1− Σ−1)UTb + ε(unTb)(σn− ε)−1vn

= V (Σ−1E(Σ− E)−1)UTb + ε(unTb)(σn− ε)−1vn

= V diag



0,· · · , 0, ε σnn− ε)



UTb + ε(unTb)(σn− ε)−1vn

= ε

σnn− ε)vn(unTb) + ε(unTb)(σn− ε)−1vn

= unTbvn( ε

σnn− ε) + ε(σn− ε)−1)

= ε(1 + σn)

σnn− ε)unTbvn.

From the inequality kxk2 ≤ kbk2kA−1k2 we have k y − x k2

k x k2 ≥ | unTb | σεn(1+σσ−ε)

k b k2 ≥ | unTb | k b k2

ε σn

.

Theorem 1.3.1 A is nonsingular and k A−1E k= r < 1. Then A + E is nonsingular and k (A + E)−1− A−1 k≤k E k k A−1 k2 /(1− r).

Proof:: Since A is nonsingular, A+E = A(I−F ), where F = −A−1E. Sincek F k= r < 1, it follows that I− F is nonsingular (by Corollary 1.2.1) and k (I − F )−1 k< 1−r1 . Then

(A + E)−1 = (I− F )−1A−1 =⇒k (A + E)−1 k≤ kA−1k 1− r and

(A + E)−1− A−1 =−A−1E(A + E)−1. It follows that

k (A + E)−1− A−1 k≤k A−1 kk E kk (A + E)−1 k≤ k A−1 k2k E k 1− r .

Lemma 1.3.1 Let

 Ax = b,

(A + ∆A)y = b + ∆b,

wherek ∆A k≤ δ k A k and k ∆b k≤ δkbk. If δκ(A) = r < 1, then A+∆A is nonsingular and kykkxk1+r1−r, where κ(A) =kAkkA−1k.

(19)

1.3 The Sensitivity of Linear System Ax = b 19 Proof: Since k A−1∆A k< δkA−1kkAk = r < 1, it follows that A + ∆A is nonsingular.

From the equality (I + A−1∆A)y = x + A−1∆b follows that

kyk ≤ k (I + A−1∆A)−1 k (kxk + δkA−1kkbk)

≤ 1

1− r(kxk + δkA−1kkbk)

= 1

1− r(kxk + rk b k kAk).

From kbk =k Ax k≤ kAkkxk follows the lemma.

1.3.3 Normwise Forward Error Bound

Theorem 1.3.2 If the assumption of Lemma 1.3.1 holds, then kx−ykkxk1−r κ(A).

Proof:: Since y− x = A−1∆b− A−1∆Ay, we have

k y − x k≤ δkA−1kkbk + δkA−1kkAkkyk.

So by Lemma 1.3.1 it holds k y − x k

kxk ≤ δκ(A) kbk

kAkkxk+ δκ(A)kyk kxk

≤ δκ(A)(1 + 1 + r

1− r) = 2δ

1− rκ(A).

1.3.4 Componentwise Forward Error Bound

Theorem 1.3.3 Let Ax = b and (A + ∆A)y = b + ∆b, where | ∆A |≤ δ | A | and

| ∆b |≤ δ | b |. If δκ(A) = r < 1, then (A + ∆A) is nonsingular and ky−xkkxk

1−r k|

A−1 || A |k. Here k | A−1 || A | k is called a Skeel condition number.

Proof:: Since k ∆A k≤ δkAk and k ∆b k≤ δkbk, the assumptions of Lemma 1.3.1 are satisfied in ∞-norm. So, A + ∆A is nonsingular and kykkxk1+r1−r.

Since y− x = A−1∆b− A−1∆Ay, we have

| y − x | ≤ | A−1 || ∆b | + | A−1 || ∆A || y |

≤ δ | A−1 || b | +δ | A−1 || A || y |

≤ δ | A−1 || A | (| x | + | y |).

By taking ∞-norm, we have

k y − x k ≤ δ k| A−1 || A |k (kxk+ 1 + r

1− rkxk)

= 2δ

1− rk| A−1 || A |k.

(20)

1.3.5 Derivation of Condition Number of Ax = b

Let

(A + εF )x(ε) = b + εf with x(0) = x.

Then we have ˙x(0) = A−1(f − F x) and x(ε) = x + ε ˙x(0) + o(ε2). Therefore, k x(ε) − x k

kxk ≤ εkA−1k{k f k

kxk +k F k} + o(ε2).

Define condition number κ(A) :=kAkkA−1k. Then we have k x(ε) − x k

kxk ≤ κ(A)(ρA+ ρb) + o(ε2), where ρA = εkF k/kAk and ρb = εkfk/kbk.

1.3.6 Normwise Backward Error

Theorem 1.3.4 Let y be the computed solution of Ax = b. Then the normwise backward error bound

η(y) := min

ε|(A + ∆A)y = b + ∆b, k∆Ak ≤ εkAk, k∆bk ≤ εkbk is given by

η(y) = krk

kAkkyk + kbk, (1.3.24)

where r = b− Ay is the residual.

Proof: The right hand side of (1.3.24) is a upper bound of η(y). This upper bound is attained for the perturbation (by construction!)

∆Amin = kAkkykrzT

kAkkyk + kbk, ∆bmin =− kbk

kAkkyk + kbkr, where z is the dual vector of y, i.e. zTy = 1 and kzk = kyk1 .

Check:

k∆Amink = η(y)kAk, or

k∆Amink = kAkkykkrzTk kAkkyk + kbk =

 krk

kAkkyk + kbk

 kAk.

That is, to prove

krzTk = krk kyk. Since

krzTk = max

kuk=1k(rzT)uk = krk max

kuk=1|zTu| = krkkzk =krk 1 kyk, we have done. Similarly, k∆bmink = η(y)kbk.

(21)

1.3 The Sensitivity of Linear System Ax = b 21

1.3.7 Componentwise Backward Error

Theorem 1.3.5 The componentwise backward error bound ω(y) := min

ε|(A + ∆A)y = b + ∆b, |∆A| ≤ ε|A|, |∆b| ≤ ε|b|

is given by

ω(y) = max

i

|r|i (A|y| + b)i

, (1.3.25)

where r = b− Ay. (note: ξ/0 = 0 if ξ = 0; ξ/0 = ∞ if ξ 6= 0.)

Proof: The right hand side of (1.3.25) is a upper bound for ω(y). This bound is at- tained for the perturbations ∆A = D1AD2 and ∆b =−D1b, where D1 = diag(ri/(A|y| + b)i) and D2 = diag(sign(yi)).

Remark 1.3.2 Theorems 1.3.4 and 1.3.5 are posterior error estimation approach.

1.3.8 Determinants and Nearness to Singularity

Bn =





1 −1 · · · −1 1 . .. ...

1 −1

0 1



, Bn−1 =





1 1 · · · 2n−2 . .. ... ...

. .. 1

0 1



.

Then det(Bn) = 1, κ(Bn) = n2n−1, σ30(B30)≈ 10−8.

Dn=



10−1 0

. ..

0 10−1

 .

Then det(Dn) = 10−n, κp(Dn) = 1 and σn(Dn) = 10−1.

(22)

參考文獻

相關文件

In this paper, we propose a practical numerical method based on the LSM and the truncated SVD to reconstruct the support of the inhomogeneity in the acoustic equation with

If x or F is a vector, then the condition number is defined in a similar way using norms and it measures the maximum relative change, which is attained for some, but not all

“Find sufficiently accurate starting approximate solution by using Steepest Descent method” + ”Compute convergent solution by using Newton-based methods”. The method of

How does drama help to develop English language skills.. In Forms 2-6, students develop their self-expression by participating in a wide range of activities

In this paper, we build a new class of neural networks based on the smoothing method for NCP introduced by Haddou and Maheux [18] using some family F of smoothing functions.

Matrix model recursive formulation of 1/N expansion: all information encoded in spectral curve ⇒ generates topological string amplitudes... This is what we

OGLE-III fields Cover ~ 100 square degrees.. In the top figure we show the surveys used in this study. We use, in order of preference, VVV in red, UKIDSS in green, and 2MASS in

We give some numerical results to illustrate that the first pass of Algorithm RRLU(r) fails but the second pass succeeds in revealing the nearly rank