Gauss quadrature approximation by Lanczos algorithm

一、 Gauss Quadrature

1.1 Gauss quadrature with weight function

For
I ( f ) = ∫ a b ρ ( x ) f ( x ) d x (1) I(f)=\int_{a}^{b} \rho(x) f(x)dx \tag{1} I(f)=abρ(x)f(x)dx(1)
If the following quadrature formula[2,p220]
∫ a b ρ ( x ) f ( x ) d x ≈ ∑ k = 0 n A k f ( x k ) (2) \int_{a}^{b}\rho(x)f(x)dx\approx \sum_{k=0}^{n}A_{k}f(x_{k}) \tag{2} abρ(x)f(x)dxk=0nAkf(xk)(2)
with the ( 2 n + 1 ) (2n+1) (2n+1) algebraic degree of precision, that is, when the degree of polynomial f ( x ) f(x) f(x) less than or equal to 2 n + 1 2n+1 2n+1, the left and right sides of equation (2) are strictly equal. Thus, it’s called the Gauss-type quadrature formula with weights ρ ( x ) \rho(x) ρ(x).
The nodes x k x_{k} xk are called Gauss points (nodes), A k A_{k} Aks are called quadrature coefficients about ρ ( x ) \rho(x) ρ(x) (weights or coefficients).

1.2 The caculation of the weights and nodes

(Note: This section is mainly about how to caculate the Gauss points and coefficents. The theoretical foundation can refer to [2, p223]. )

  • 1)Undeterminated coefficient method: As the Gaussian quadrature formula in equation (2) is accurately for any polynomial with degree at most 2 n + 1 2n+1 2n+1. Thus, for the polynomial sequence { 1 , x , x 2 , ⋯   , x 2 n + 1 } \{1,x,x^{2},\cdots,x^{2n+1}\} {1,x,x2,,x2n+1}, equation (2) are all exactly equal.
    Based on the above analysis, we can construct the following ( 2 n + 2 ) (2n+2) (2n+2) equations about quadrature nodes { x k } k = 0 n \{x_{k}\}_{k=0}^{n} {xk}k=0n and quadrature coefficient { A k } k = 0 n \{A_{k}\}_{k=0}^{n} {Ak}k=0n :
    { ∑ k = 0 n A k = ∫ a b 1   d x ∑ k = 0 n x k A k = ∫ a b x   d x ∑ k = 0 n x k 2 A k = ∫ a b x 2   d x ⋮ ∑ k = 0 n x k 2 n + 1 A k = ∫ a b x 2 n + 1   d x \left\{\begin{array}{l}\sum_{k=0}^{n} A_{k}=\int_{a}^{b} 1 \mathrm{~d} x \\ \sum_{k=0}^{n} x_{k} A_{k}=\int_{a}^{b} x \mathrm{~d} x \\ \sum_{k=0}^{n} x_{k}^{2} A_{k}=\int_{a}^{b} x^{2} \mathrm{~d} x \\ \vdots \\ \sum_{k=0}^{n} x_{k}^{2 n+1} A_{k}=\int_{a}^{b} x^{2 n+1} \mathrm{~d} x\end{array}\right. k=0nAk=ab1 dxk=0nxkAk=abx dxk=0nxk2Ak=abx2 dxk=0nxk2n+1Ak=abx2n+1 dx

Note: when n ≥ 2 n\geq 2 n2, it is difficult to solve the above system of equations directly.

  • 2)Orthogonal polynomials method: The basic idea mainly based on the following two theorem.

[2, Theorem 5.5.1, p221] Let I ( f ) = ∫ a b f ( x ) d x I(f)=\int_{a}^{b}f(x)\mathrm{d}x I(f)=abf(x)dx and its interpolation quadrature formula is I n ( f ) = ∑ k = 0 n A k f ( x k ) , I_{n}(f)=\sum_{k=0}^{n}A_{k}f(x_{k}), In(f)=k=0nAkf(xk), and denote W n + 1 ( x ) = ( x − x 0 ) ( x − x 1 ) ⋯ ( x − x n ) . W_{n+1}(x)=(x-x_{0})(x-x_{1})\cdots(x-x_{n}). Wn+1(x)=(xx0)(xx1)(xxn).
Then { x k } k = 0 n \{x_{k}\}_{k=0}^{n} {xk}k=0n are Gaussian nodes if and only if W n + 1 ( x ) W_{n+1}(x) Wn+1(x) is orthogonal to any polynomial p ( x ) p(x) p(x) with degree at most n n n, that is ∫ a b p ( x ) W n + 1 ( x ) d x = 0. \int_{a}^{b}p(x)W_{n+1}(x)\mathrm{d}x=0. abp(x)Wn+1(x)dx=0.

[2, Theorem 5.5.3, p223] Let { q n ( x ) } n = 0 ∞ \{q_{n}(x)\}_{n=0}^{\infty} {qn(x)}n=0 is an orthogonal polynomilas sequence on [ a , b ] [a,b] [a,b], thus q n ( x ) q_{n}(x) qn(x) has n n n different zeros on [ a , b ] [a,b] [a,b].

` As the orthogonal polynomials sequence { q k ( x ) } k = 0 n + 1 \{q_{k}(x)\}_{k=0}^{n+1} {qk(x)}k=0n+1 form a basis for a polynomial space of degree n+1, and q n ( x ) q_{n}(x) qn(x) is orthogoanl to any polynomial of degree at most n on the interval [a,b].
Thus, for a given weight function ρ ( x ) \rho(x) ρ(x) (can be ρ ( x ) = 1 \rho(x)=1 ρ(x)=1), if we can find a orthogonal polynomials sequence { q 0 ( x ) , q 1 ( x ) , ⋯   , q n ( x ) , q n + 1 ( x ) } \{q_{0}(x),q_{1}(x),\cdots,q_{n}(x),q_{n+1}(x)\} {q0(x),q1(x),,qn(x),qn+1(x)} related to the weight funciton ρ ( x ) \rho(x) ρ(x).
Where q n + 1 ( x ) q_{n+1}(x) qn+1(x) is exact degree n + 1 n+1 n+1, suppose its zeros are t 0 , ⋯   , t n t_{0},\cdots,t_{n} t0,,tn, then they are also the Gauss points { x k } k = 0 n \{x_{k}\}_{k=0}^{n} {xk}k=0n, their coresponding coefficients { A k } k = 0 n \{A_{k}\}_{k=0}^{n} {Ak}k=0n are derived by
A k = ∫ a b ρ ( x ) l k ( x ) d x , (3) A_{k}=\int_{a}^{b}\rho(x)l_{k}(x)dx, \tag{3} Ak=abρ(x)lk(x)dx,(3)
where l k ( t ) l_{k}(t) lk(t) are base function of interpolation
l k ( t ) = ∏ j = 0 , j ≠ k n t − t j t k − t j . l_{k}(t)=\prod_{j=0,j\neq k}^{n}\frac{t-t_{j}}{t_{k}-t_{j}}. lk(t)=j=0,j=kntktjttj.

Note: Using the zeros of orthogoanl polynomials to construct the Gauss quadrature formula, which is only effective for some special weight functions, such as 1 1 1(Legendre), 1 1 − x 2 \frac{1}{\sqrt{1-x^{2}}} 1x2 1(Chebyshev 1), 1 − x 2 \sqrt{1-x^{2}} 1x2 (Chebyshev 2) and e − x 2 e^{-x^{2}} ex2(Hermite), et.al.
However, for the general weight function, the above undertermined coefficient method is usually adopted.


二、Lanczos algorithm

2.1 Krylov subspace

  • Krylov subspace: A A A is real matrix (unsymmetric) of order n n n. Let v v v be a given vector and
    K k = ( v , A v , ⋯   , A k − 1 v ) K_{k}=\left(v, \quad A v, \quad \cdots, \quad A^{k-1} v\right) Kk=(v,Av,,Ak1v)
    be the Krylov matrix of dimension n × k n\times k n×k. The subspace spanned by the columns of matrix K k K_{k} Kk is called a Krylov subspace and denoted by K k ( A , v ) K_{k}(A,v) Kk(A,v) or K ( A , v ) K(A,v) K(A,v).

Note: The natural basis of the Krylov subspace K ( A , v ) K(A,v) K(A,v) given by the columns of the Krylov matirx K k K_{k} Kk is badly conditioned when k k k is large.

2.2 Arnoldi algorithm

Arnoldi algorithm constructs an orthonormal basis of the Krylov subspace K ( A , v ) K(A,v) K(A,v) by applying a variant of the Gram-Schmidt orthogonalization process.

  • Arnoldi algorithm: Set vectors v ( j + 1 ) = A v ( j ) v^{(j+1)}=Av^{(j)} v(j+1)=Av(j) with v ( 1 ) = v v^{(1)}=v v(1)=v, then K ( A , v ) K(A,v) K(A,v) is spanned by the vectors { v ( j ) } j = 1 k \{v^{(j)}\}_{j=1}^{k} {v(j)}j=1k.For constucting orthogonal basis vectors v j v^{j} vj, instead of orthogonalizing A j v A^{j}v Ajv against the previous vectors, orthogonalize A v j Av^{j} Avj.
    Starting from v 1 = v v^{1}=v v1=v(normalized,that is v 1 = v / ∥ v ∥ v^{1}=v/\|v\| v1=v/∥v),the ( j + 1 ) (j+1) (j+1)st vector of the basis is computed by using the previous vectors,
  • Projection: h i , j = ( A v j , v i ) , i = 1 , ⋯   , j , h_{i,j}=(Av^{j},v^{i}),i=1,\cdots,j, hi,j=(Avj,vi),i=1,,j,
  • Orthogonalize: v ~ j + 1 = A v j − ∑ i = 1 j h i , j v i , \tilde{v}^{j+1}=Av^{j}-\sum_{i=1}^{j}h_{i,j}v^{i}, v~j+1=Avji=1jhi,jvi,
  • Length: h j + 1 , j = ∥ v ~ j + 1 ∥ h_{j+1,j}=\|\tilde{v}^{j+1}\| hj+1,j=v~j+1,(if h j + 1 , j = 0 h_{j+1,j}=0 hj+1,j=0 then stop)
  • Normalize: v j + 1 = v ~ j + 1 h j + 1 , j v^{j+1}=\frac{\tilde{v}^{j+1}}{h_{j+1,j}} vj+1=hj+1,jv~j+1

As h j + 1 , j v j + 1 = A v j − ∑ i = 1 j h i , j v i h_{j+1,j}v^{j+1}=Av^{j}-\sum_{i=1}^{j}h_{i,j}v^{i} hj+1,jvj+1=Avji=1jhi,jvi, we have
A v j = ∑ i = 1 j h i , j v i + h j + 1 , j v j + 1 . Av^{j}=\sum_{i=1}^{j}h_{i,j}v^{i}+h_{j+1,j}v^{j+1}. Avj=i=1jhi,jvi+hj+1,jvj+1.
If we collect the vectors v j , j = 1 , ⋯   , k v^{j},j=1,\cdots,k vj,j=1,,k in a matrix V k V_{k} Vk, that is V k = [ v 1 , v 2 , ⋯   , v k ] V_{k}=[v_{1},v_{2},\cdots,v_{k}] Vk=[v1,v2,,vk] (thus V k V_{k} Vk is column orthogonal matrix), the relations defining the vectors v k + 1 v^{k+1} vk+1 can be written in the matrix form as
A V k = V k H k + h k + 1 , k v k + 1 ( e k ) T , (4) A V_{k}=V_{k} H_{k}+h_{k+1, k} v^{k+1}\left(e^{k}\right)^{T}, \tag{4} AVk=VkHk+hk+1,kvk+1(ek)T,(4)
where H k H_{k} Hk is an upper Hessenberg matrix with elements h i , j h_{i,j} hi,j, h i , j = 0 , j = 1 , ⋯   , i − 2 , i > 2 h_{i,j}=0,j=1,\cdots,i-2,i>2 hi,j=0,j=1,,i2,i>2.
H k = [ h 1 , 1 h 1 , 2 h 1 , 3 ⋯ h 1 , k h 2 , 1 h 2 , 2 h 2 , 3 ⋯ h 2 , k 0 h 3 , 2 h 3 , 3 ⋯ h 3 , k 0 0 h 4 , 3 ⋯ h 4 , k ⋮ ⋮ ⋮ ⋱ ⋮ 0 0 0 h k , k − 1 h k , k ] H_{k}=\left[ \begin{array}{lllll} h_{1,1} & h_{1,2} & h_{1,3} & \cdots & h_{1,k}\\ h_{2,1} & h_{2,2} & h_{2,3} & \cdots & h_{2,k}\\ 0 & h_{3,2} & h_{3,3} & \cdots & h_{3,k}\\ 0 & 0 & h_{4,3} & \cdots & h_{4,k}\\ \vdots & \vdots & \vdots & \ddots & \vdots\\ 0 & 0 & 0 & h_{k,k-1} & h_{k,k}\\ \end{array} \right ] Hk= h1,1h2,1000h1,2h2,2h3,200h1,3h2,3h3,3h4,30hk,k1h1,kh2,kh3,kh4,khk,k

2.3 Lanczos algorithm

Multiplying equation (4) by V k T V_{k}^{T} VkT and using orthogonality, we have
H k = V k T A V k . H_{k}=V_{k}^{T}A V_{k}. Hk=VkTAVk.
If matrix A A A is symmetric, then the Hessenberg matrix H k H_{k} Hk is tridiagonal and denoted by J k J_{k} Jk, that is J i , j = 0 , j = i + 2 , ⋯   , k . J_{i,j}=0,j=i+2,\cdots,k. Ji,j=0,j=i+2,,k. (or h i , j = 0 , i = 1 , ⋯   , j − 2 , j + 2 , ⋯   , k h_{i,j}=0,i=1,\cdots,j-2,j+2,\cdots,k hi,j=0,i=1,,j2,j+2,,k). This implies that the new vector v k + 1 v^{k+1} vk+1 can be computed by using only the two previous v k v^{k} vk and v k − 1 v^{k-1} vk1.
h j + 1 , j v j + 1 = A v j − ∑ i = 1 j h i , j v i = A v j − h j − 1 , j v j − 1 − h j , j v j . h_{j+1,j}v^{j+1}=Av^{j}-\sum_{i=1}^{j}h_{i,j}v^{i}=Av^{j}-h_{j-1,j}v^{j-1}-h_{j,j}v^{j}. hj+1,jvj+1=Avji=1jhi,jvi=Avjhj1,jvj1hj,jvj.
In matrix form, set η i \eta_{i} ηi denote the nonzero off-diagonal entries of J k J_{k} Jk,
A V k = V k J k + η k v k + 1 ( e k ) T . (5) A V_{k}=V_{k} J_{k}+\eta_{k} v^{k+1}\left(e^{k}\right)^{T}. \tag{5} AVk=VkJk+ηkvk+1(ek)T.(5)

Equation (5) describes in matrix form the elegant Lanczos algorithm.
For simple notation system, starting from a nonzero vector v 1 = v / ∥ v ∥ , α 1 = ( A v 1 , v 1 ) , v ~ 2 = A v 1 − α 1 v 1 v^{1}=v/\|v\|,\alpha_{1}=(Av^{1},v^{1}),\tilde{v}^{2}=Av^{1}-\alpha_{1}v^{1} v1=v/∥v,α1=(Av1,v1),v~2=Av1α1v1, and then for k = 2 , 3 , ⋯   , k=2,3,\cdots, k=2,3,,
η k − 1 = ∥ v ~ k ∥ , v k = v ~ k η k − 1 , α k = ( v k , A v k ) = ( v k ) T A v k , v ~ k + 1 = A v k − α k v k − η k − 1 v k − 1 . \begin{array}{c} \eta_{k-1}=\|\tilde{v}^{k}\|,\\ v^{k}=\frac{\tilde{v}^{k}}{\eta_{k-1}},\\ \alpha_{k}=\left(v^{k}, A v^{k}\right)=\left(v^{k}\right)^{T} A v^{k},\\ \tilde{v}^{k+1}=A v^{k}-\alpha_{k} v^{k}-\eta_{k-1} v^{k-1}.\\ \end{array} ηk1=v~k,vk=ηk1v~k,αk=(vk,Avk)=(vk)TAvk,v~k+1=Avkαkvkηk1vk1.
J k = [ α 1 η 1 0 ⋯ 0 η 1 α 2 η 2 ⋯ 0 0 η 2 α 3 ⋯ 0 0 0 η 3 ⋯ 0 ⋮ ⋮ ⋮ ⋱ ⋮ 0 0 0 η k − 1 α k ] J_{k}=\left[ \begin{array}{lllll} \alpha_{1} & \eta_{1} & 0 & \cdots & 0\\ \eta_{1} & \alpha_{2} & \eta_{2} & \cdots & 0\\ 0 & \eta_{2} & \alpha_{3} & \cdots & 0\\ 0 & 0 & \eta_{3} & \cdots & 0\\ \vdots & \vdots & \vdots & \ddots & \vdots\\ 0 & 0 & 0 & \eta_{k-1} & \alpha_{k}\\ \end{array} \right ] Jk= α1η1000η1α2η2000η2α3η30ηk10000αk

三、Two important theory

3.1 Theorem 1

Theorem1 [1,Th.4.1]: Let χ k ( λ ) \chi_{k}(\lambda) χk(λ) be the determinant of J k − λ I J_{k}-\lambda I JkλI(which is a monic polynomial); then
v k = p k ( A ) v 1 , p k ( λ ) = ( − 1 ) k − 1 χ k − 1 ( λ ) η 1 ⋯ η k − 1 , k > 1 , p 1 ≡ 1 v^{k}=p_{k}(A) v^{1}, \quad p_{k}(\lambda)=(-1)^{k-1} \frac{\chi_{k-1}(\lambda)}{\eta_{1} \cdots \eta_{k-1}}, k>1, \quad p_{1} \equiv 1 vk=pk(A)v1,pk(λ)=(1)k1η1ηk1χk1(λ),k>1,p11
The polynomials p k p_{k} pk of degree k − 1 k-1 k1 are called the normalized Lanczos polynomials.

This theorem describes the most important property of Lanczos alg orithm:The Lanczos vectors v k v^{k} vk are given as a polynomial in matrix A A A applied to the initial vector v 1 v^{1} v1.

It is easy to verify that (expand on the last row of J k + 1 J_{k+1 } Jk+1
det ⁡ ( J k + 1 ) = α k + 1 det ⁡ ( J k ) − η k 2 det ⁡ ( J k − 1 ) , (6) \det(J_{k+1}) = \alpha_{k+1}\det(J_{k})-\eta_{k}^{2}\det(J_{k-1}), \tag{6} det(Jk+1)=αk+1det(Jk)ηk2det(Jk1),(6)
and
det ⁡ ( J k + 1 − λ I ) = ( α k + 1 − λ ) det ⁡ ( J k − λ I ) − η k 2 det ⁡ ( J k − 1 − λ I ) . (7) \det(J_{k+1}-\lambda I) = (\alpha_{k+1}-\lambda)\det(J_{k}-\lambda I)-\eta_{k}^{2}\det(J_{k-1}-\lambda I). \tag{7} det(Jk+1λI)=(αk+1λ)det(JkλI)ηk2det(Jk1λI).(7)
From the expression of polynomial p k ( λ ) p_{k}(\lambda) pk(λ) in Theorem 1, we know that the Lanczos polynomials satisfy a scalar three-term recurrence,
η k p k + 1 ( λ ) = ( λ − α k ) p k ( λ ) − η k − 1 p k − 1 ( λ ) , k = 1 , 2 , … , (8) \eta_{k} p_{k+1}(\lambda)=\left(\lambda-\alpha_{k}\right) p_{k}(\lambda)-\eta_{k-1} p_{k-1}(\lambda), k=1,2, \ldots, \tag{8} ηkpk+1(λ)=(λαk)pk(λ)ηk1pk1(λ),k=1,2,,(8)
with initial conditions p 0 ≡ 0 , p 1 ≡ 1 p_{0} \equiv 0, p_{1} \equiv 1 p00,p11.

3.2 Theorem 2

Theorem2 [1, Th.4.2]:
Consider the Lanczos vectors v k v^{k} vk. There exists a measure α ( λ ) \alpha(\lambda) α(λ) such that
( v k , v l ) = ⟨ p k , p l ⟩ = ∫ a b p k ( λ ) p l ( λ ) d α ( λ ) , (9) \left(v^{k}, v^{l}\right)=\left\langle p_{k}, p_{l}\right\rangle=\int_{a}^{b} p_{k}(\lambda) p_{l}(\lambda) d \alpha(\lambda), \tag{9} (vk,vl)=pk,pl=abpk(λ)pl(λ)dα(λ),(9)
where α ≤ λ 1 = λ m i n \alpha\leq \lambda_{1}=\lambda_{min} αλ1=λmin and b ≥ λ n = λ m a x b\geq \lambda_{n}=\lambda_{max} bλn=λmax, λ m i n \lambda_{min} λmin and λ m a x \lambda_{max} λmax being the smallest and largest eigenvalues of A A A, and p i p_{i} pi are the Lanczos polynomials associated with A A A and v 1 v^{1} v1.
α ( λ ) = { 0 ,  if  λ < λ 1 ∑ j = 1 i [ v ^ j ] 2 ,  if  λ i ≤ λ < λ i + 1 ∑ j = 1 n [ v ^ j ] 2 ,  if  λ n ≤ λ (10) \alpha(\lambda)=\left\{\begin{array}{ll}0, & \text { if } \lambda<\lambda_{1} \\ \sum_{j=1}^{i}\left[\hat{v}_{j}\right]^{2}, & \text { if } \lambda_{i} \leq \lambda<\lambda_{i+1} \\ \sum_{j=1}^{n}\left[\hat{v}_{j}\right]^{2}, & \text { if } \lambda_{n} \leq \lambda\end{array}\right. \tag{10} α(λ)= 0,j=1i[v^j]2,j=1n[v^j]2, if λ<λ1 if λiλ<λi+1 if λnλ(10)
where v ^ = Q T v 1 \hat{v}=Q^{T}v^{1} v^=QTv1, v 1 = v / ∥ v ∥ v^{1}=v/\|v\| v1=v/∥v, and the spectral decomposition of A A A is A = Q Λ Q T A=Q\Lambda Q^{T} A=QΛQT.

That is, normalized Lanczos polynomials sequence { p k ( λ ) } \{p_{k}(\lambda)\} {pk(λ)} is orthogonal in the sense of measure α ( λ ) \alpha(\lambda) α(λ).

Note: For the sake of simplicity, here suppose that the eigenvalues of A A A are distinct.

四、Gauss quadrature by Lanczos algorithm

For symetric matrix A A A with eigenvalue decomposition A = Q Λ Q T A=Q\Lambda Q^{T} A=QΛQT, f f f is smooth funciton, v v v is random unit vector, then
v T f ( A ) v = v T Q f ( Λ ) Q T v v^{T}f(A)v=v^{T}Qf(\Lambda)Q^{T}v vTf(A)v=vTQf(Λ)QTv
set v ^ = Q T v \hat{v}=Q^{T}v v^=QTv, then
v T f ( A ) v = v ^ T f ( Λ ) v = ∑ j = 1 m f ( λ i ) [ v ^ j ] 2 v^{T}f(A)v=\hat{v}^{T}f(\Lambda)v=\sum_{j=1}^{m}f(\lambda_{i})[\hat{v}_{j}]^{2} vTf(A)v=v^Tf(Λ)v=j=1mf(λi)[v^j]2
Consider the above sum as Riemann-Stieltjes integral
∑ j = 1 m f ( λ i ) [ v ^ j ] 2 = ∫ a b f ( t ) d α ( t ) . (11) \sum_{j=1}^{m}f(\lambda_{i})[\hat{v}_{j}]^{2}=\int_{a}^{b}f(t)d\alpha(t). \tag{11} j=1mf(λi)[v^j]2=abf(t)dα(t).(11)
where the measure α ( t ) \alpha(t) α(t) is defined as:
α ( t ) = { 0 ,  if  t < λ 1 = a ∑ j = 1 i [ v ^ j ] 2 ,  if  λ i ≤ t < λ i + 1 ∑ j = 1 m [ v ^ j ] 2 ,  if  b = λ m ≤ t . \alpha(t)=\left\{\begin{array}{ll}0, & \text { if } t<\lambda_{1}=a \\ \sum_{j=1}^{i}\left[\hat{v}_{j}\right]^{2}, & \text { if } \lambda_{i} \leq t<\lambda_{i+1} \\ \sum_{j=1}^{m}\left[\hat{v}_{j}\right]^{2}, & \text { if } b=\lambda_{m} \leq t.\end{array}\right. α(t)= 0,j=1i[v^j]2,j=1m[v^j]2, if t<λ1=a if λit<λi+1 if b=λmt.

The integral in Equation (11) can be estimated using the Gauss quadrature,
∫ a b f ( t ) d α ( t ) ≈ ∑ k = 1 m A k f ( x k ) . (12) \int_{a}^{b}f(t)d\alpha(t)\approx \sum_{k=1}^{m}A_{k}f(x_{k}).\tag{12} abf(t)dα(t)k=1mAkf(xk).(12)
As stated in the Theorem2, { p k } k = 1 m + 1 \{p_{k}\}_{k=1}^{m+1} {pk}k=1m+1 are orthonormal polynomial and the measure funciton is also α ( t ) \alpha(t) α(t).
Thus, the Gaussian quadrature nodes in (12) can be derived by caculating the zeros of polynomial p m + 1 ( x ) p_{m+1}(x) pm+1(x) with degree of m m m.
We can rewirte the three-term recurrence in Equation (8) in matrix form. Precisely, let p ( λ ) = [ p 1 ( λ ) , p 2 ( λ ) , ⋯   , p k ( λ ) ] T \mathbf{p}(\lambda)=[p_{1}(\lambda),p_{2}(\lambda),\cdots,p_{k}(\lambda)]^{T} p(λ)=[p1(λ),p2(λ),,pk(λ)]T and p 0 ( λ ) = 0 p_{0}(\lambda)=0 p0(λ)=0, then
λ p ( λ ) = p ( λ ) J k + η k p k + 1 ( λ ) e k T . (13) \lambda \mathbf{p}(\lambda)=\mathbf{p}(\lambda) J_{k}+\eta_{k}p_{k+1}(\lambda)e_{k}^{T}. \tag{13} λp(λ)=p(λ)Jk+ηkpk+1(λ)ekT.(13)
Then the zeros of p k + 1 ( λ ) p_{k+1}(\lambda) pk+1(λ) are equal to the eigenvalues of J k J_{k} Jk.
Thus, the Gaussian quadrature nodes in (12) can be derived by caculating the eigenvalues of tridiagonal Jacobi matrix J m J_{m} Jm, where J m J_{m} Jm is derived by m-steps Lanczos algorithm associate with A A A and v v v.

Note: p k + 1 ( λ ) p_{k+1}(\lambda) pk+1(λ) is polynomial of degree k.

In conclusion, the Gauss point prolbem in Equation(12) can be get by the eigenvalues of J k J_{k} Jk which derived by Lanczos algorithm for matrix A A A and v v v.


References:
[1] Golub, G. H. and Meurant, G. Matrices, Moments and Quadrature with Applications, Princeton University Press, 2010.
[2] 孙志忠,袁慰平,闻震初. 数值分析(第3版),东南大学出版社.

评论
添加红包

请填写红包祝福语或标题

红包个数最小为10个

红包金额最低5元

当前余额3.43前往充值 >
需支付:10.00
成就一亿技术人!
领取后你会自动成为博主和红包主的粉丝 规则
hope_wisdom
发出的红包
实付
使用余额支付
点击重新获取
扫码支付
钱包余额 0

抵扣说明:

1.余额是钱包充值的虚拟货币,按照1:1的比例进行支付金额的抵扣。
2.余额无法直接购买下载,可以购买VIP、付费专栏及课程。

余额充值