This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The degree of ill-posedness for some composition governed by the Cesàro operator

Yu Deng Faculty of Mathematics, Chemnitz University of Technology, 09107 Chemnitz, Germany [email protected] Hans-Jürgen Fischer 01159 Dresden, Germany [email protected]  and  Bernd Hofmann Faculty of Mathematics, Chemnitz University of Technology, 09107 Chemnitz, Germany [email protected]
(Date:
Corresponding author: B. Hofmann)
Abstract.

In this article, we consider the singular value asymptotics of compositions of compact linear operators mapping in the real Hilbert space of quadratically integrable functions over the unit interval. Specifically, the composition is given by the compact simple integration operator followed by the non-compact Cesàro operator possessing a non-closed range. We show that the degree of ill-posedness of that composition is two, which means that the Cesàro operator increases the degree of ill-posedness by the amount of oneone compared to the simple integration operator.

Keywords: Cesàro operator, linear inverse problem, degree of ill-posedness, composition operator, compact operator, singular value decomposition, twofold integration operator.

MSC: 47A52, 47B06, 65J20, 40G05

1. Introduction

This is a new paper in the series of articles [3, 7, 8] and recently [5] that are dealing with the degree of ill-posedness of linear operator equations

(1) Ax=y,A\,x\,=\,y\,,

where the compact linear operator A:XYA:X\to Y is factorized as

(2) A:XKZNY,\begin{CD}A:\;X@>{K}>{}>Z@>{N}>{}>Y\,,\end{CD}

for infinite dimensional separable Hilbert spaces X,YX,Y and ZZ. In this context, A=NKA=N\circ K denotes the composition of an injective compact linear operator K:XZK:X\to Z and a bounded injective and non-compact, but not continuously invertible, linear operator N:ZYN:~{}Z\to Y. The imposed requirements on KK and NN imply that the ranges (K)\mathcal{R}(K), (N)\mathcal{R}(N) and (A)\mathcal{R}(A) are infinite dimensional, but non-closed, subspaces of the corresponding Hilbert spaces. Following the concept of Nashed [12], the total equation (1) and the inner linear operator equation

(3) Kx=zK\,x\,=\,z

are ill-posed of type II due to the compactness of AA and KK, whereas the outer linear operator equation

(4) Nz=yN\,z\,=\,y

is ill-posed of type I, since NN is non-compact.

We recall here a definition of the interval and degree of ill-posedness along the lines of [6]:

Definition 1.

Let {σn(A)}n=1\{\sigma_{n}(A)\}_{n=1}^{\infty} the non-increasing sequence of singular values of the injective and compact linear operator A:XYA:X\to Y, tending to zero as nn\to\infty. Based on the well-defined interval of ill-posedness introduced as

[μ¯(A),μ¯(A)]=[lim infnlogσn(A)logn,lim supnlogσn(A)logn][0,],[\underline{\mu}(A),\overline{\mu}(A)]=\left[\liminf\limits_{n\to\infty}\frac{-\log\sigma_{n}(A)}{\log n}\,,\,\limsup\limits_{n\to\infty}\frac{-\log\sigma_{n}(A)}{\log n}\right]\subset[0,\infty],

we say that the operator AA, and respectively the associated operator equation (1), is ill-posed of degree μ=μ(A)(0,)\mu=\mu(A)\in(0,\infty) if μ=μ¯(A)=μ¯(A)\mu=\underline{\mu}(A)=\overline{\mu}(A), i.e., if the interval of ill-posedness degenerates into a single point.

The main objective of the above mentioned article series and of the present study is to learn whether the non-compact operator NN can amend the degree of ill-posedness of the compact operator KK by such a composition A=NKA=N\circ K. Such amendment would be impossible if NN were continuously invertible, or in other words if the outer operator equation (4) were well-posed. Since in our setting zero belongs to the spectrum of NN, the singular value asymptotics of AA can differ from that of KK, but due to the inequality

(5) σn(A)N(Z,Y)σn(K)(n)\sigma_{n}(A)\leq\|N\|_{\scriptscriptstyle{\mathcal{L}}(Z,Y)}\,\sigma_{n}(K)\qquad(n\in\mathbb{N})

only in the sense of a growing decay rate, which means a growing degree of ill-posedness of AA compared with KK. Note that the estimate (5) is an immediate consequence of the Courant–Fischer min-max principle for the characterization of singular values.

In this study, we will focus on one common Hilbert space X=Y=Z=L2(0,1)X=Y=Z=L^{2}(0,1), the space of quadratically integrable real functions over the unit interval of the real axis. We consider as operator KK the compact simple integration operator J:L2(0,1)L2(0,1)J:L^{2}(0,1)\to L^{2}(0,1) defined as

(6) [Jx](s):=0sx(t)𝑑t(0s1),[Jx](s):=\int_{0}^{s}x(t)dt\qquad(0\leq s\leq 1)\,,

where the singular system

{σn(J);un(J);vn(J)}n=1withJun(J)=σn(J)vn(J)(n)\{\sigma_{n}(J);u_{n}(J);v_{n}(J)\}_{n=1}^{\infty}\quad\mbox{with}\quad\ Ju_{n}(J)=\sigma_{n}(J)v_{n}(J)\quad(n\in\mathbb{N})

is of the form

(7) {2(2n1)π,2cos(n12)πt(0t1);2sin(n12)πt(0t1)}n=1.\left\{\frac{2}{(2n-1)\pi},\sqrt{2}\cos\left(n-\frac{1}{2}\right)\pi t\;(0\leq t\leq 1);\sqrt{2}\sin\left(n-\frac{1}{2}\right)\pi t\;(0\leq t\leq 1)\right\}_{n=1}^{\infty}\,.

The asymptotics 111We use the notation anbna_{n}\asymp b_{n} for sequences of positive numbers ana_{n} and bnb_{n} satisfying inequalities c¯bnanc¯bn\underline{c}\,b_{n}\leq a_{n}\leq\overline{c}\,b_{n} with constants 0<c¯c¯<0<\underline{c}\leq\overline{c}<\infty for sufficiently large nn\in\mathbb{N}.

(8) σn(J)1n\sigma_{n}(J)\asymp\frac{1}{n}

shows that the degree of ill-posedness of JJ in the sense of Definition 1 is one.

Let us briefly mention the former results using JJ as compact operator in such a composition. In the papers [7, 8], it was shown that wide classes of bounded non-compact multiplication operators M:L2(0,1)L2(0,1)M:L^{2}(0,1)\to L^{2}(0,1) defined as

(9) [Mx](t):=m(t)x(t)(0t1),[Mx](t):=m(t)\,x(t)\qquad(0\leq t\leq 1)\,,

with a multiplier functions mL(0,1)m\in L^{\infty}(0,1) possessing essential zeros, do not amend the singular value asymptotics. This means σn(MJ)σn(J)\sigma_{n}(M\circ J)\asymp\sigma_{n}(J) and implies that the ill-posedness degree of A=MJA=M\circ J stays at one. A first example, where the degree of ill-posedness grows, was presented in [5] for A=HJA=H\circ J with the bounded non-compact Hausdorff operator H:L2(0,1)2()H:L^{2}(0,1)\to\ell^{2}(\mathbb{N}) defined as

(10) [Hz]j:=01tj1z(t)𝑑t(j=1,2,),[Hz]_{j}:=\int_{0}^{1}t^{j-1}z(t)dt\qquad(j=1,2,...)\,,

and we refer for properties of HH to the article [4]. In [5, §5] it could be shown that for some positive constants c¯\underline{c} and c¯\overline{c} the singular values behave as

(11) exp(c¯n)σn(HJ)c¯n3/2.\exp(-\underline{c}\,n)\leq\sigma_{n}(H\circ J)\leq\frac{\overline{c}}{n^{3/2}}\,.

Hence, the ill-posedness interval for the composition A=HJA=H\circ J is a subset of the interval [32,]\left[\frac{3}{2},\infty\right].

Now in the present study, we only consider as bounded non-compact and not continuously invertible operator NN the continuous Cesàro operator C:L2(0,1)L2(0,1)C:L^{2}(0,1)\to L^{2}(0,1) defined as

(12) [Cx](s):=1s0sx(t)𝑑t(0<s1).[Cx](s):=\frac{1}{s}\int_{0}^{s}x(t)dt\qquad(0<s\leq 1)\,.

We refer to [1] and [9, 10] for properties including boundedness and further discussions concerning this operator CC. The two properties of CC, which are most important for the present study, are outlined in the following Lemma 1. Its proof is given in the appendix.

Lemma 1.

The injective bounded linear operator C:L2(0,1)L2(0,1)C:L^{2}(0,1)\to L^{2}(0,1) from (12) is non-compact and not continuously invertible, i.e., the inverse operator C1:(C)L2(0,1)L2(0,1)C^{-1}:\mathcal{R}(C)\subset L^{2}(0,1)\to L^{2}(0,1) is unbounded and hence the range (C)\mathcal{R}(C) is not a closed subset of L2(0,1)L^{2}(0,1).

Precisely, in our composition the compact simple integration operator JJ from (6) is followed by the non-compact Cesàro operator CC from (12) as A:=CJ:L2(0,1)L2(0,1)A:=C\circ J:L^{2}(0,1)\to L^{2}(0,1). The compact composite operator AA can be written explicitly as

(13) [Ax](s):=1s0s(st)x(t)𝑑t=0sstsx(t)𝑑t(0s1).[Ax](s):=\frac{1}{s}\int_{0}^{s}(s-t)\,x(t)dt=\int_{0}^{s}\frac{s-t}{s}\,x(t)dt\qquad(0\leq s\leq 1)\,.

The paper is organized as follows. In Section 2 the relationship of the composite operator AA with the twofold integration operator is presented. In this way, the lower bound of the degree of the ill-posedness of AA can be determined. We analyse some properties of the Hilbert-Schmidt operator AA in Section 3 and one can establish its approximate decay rate numerically via calculating the eigenvalues of AAA^{*}A with symmetric kernel. Finally, with the aid of a suitable orthonormal basis in L2(0,1)L^{2}(0,1) we are able to identify the degree of ill-posedness of the composite operator A=CJA=C\circ J in Section 4.

2. Cross connections to the twofold integration operator

We recall the family of Riemann-Liouville fractional integral operators Jκ:L2(0,1)L2(0,1)J^{\kappa}:L^{2}(0,1)\to L^{2}(0,1) defined as

(14) [Jκx](s):=1Γ(κ)0s(st)κ1x(t)𝑑t(0s1).[J^{\kappa}x](s):=\frac{1}{\Gamma(\kappa)}\int_{0}^{s}(s-t)^{\kappa-1}x(t)dt\qquad(0\leq s\leq 1)\,.

For all κ>0\kappa>0, the linear operators JκJ^{\kappa} are injective and compact. We know (cf. [18] and references therein) that the singular value asymptotics

(15) σn(Jκ)1nκ\sigma_{n}(J^{\kappa})\asymp\frac{1}{n^{\kappa}}

holds true. The solution of the equation Jκx=yJ^{\kappa}x=y can be seen as the κ\kappa’s fractional derivative of yy such that degree of ill-posedness of this equation is κ\kappa and grows with the order of differentiation. Besides the simple integration operator JJ from (8) for κ=1\kappa=1, the twofold integration operator

(16) [J2x](s):=0s(st)x(t)𝑑t(0s1)[J^{2}x](s):=\int_{0}^{s}(s-t)x(t)dt\qquad(0\leq s\leq 1)

for κ=2\kappa=2 (see further details in [15, Section 11.5]) plays some prominent role in our study. Obviously, we can write for the composite operator AA from (13) on the one hand

(17) [Ax](s)=[J2x](s)/s(0<s1),[Ax](s)=[J^{2}x](s)/s\qquad(0<s\leq 1)\,,

and on the other hand

(18) J2=MAfor multiplication operator[Mx](s)=sx(s).J^{2}=M\circ A\quad\mbox{for multiplication operator}\quad[Mx](s)=s\,x(s).

We mention here that formula (17) shows that the range (A)\mathcal{R}(A) of A:L2(0,1)L2(0,1)A:L^{2}(0,1)\to L^{2}(0,1) is a subset of the space of continuous functions over [0,1][0,1]. Namely, (J2)\mathcal{R}(J^{2}) is a subset of the Sobolev space H2(0,1)H^{2}(0,1), which is continuously embedded in C1[0,1]C^{1}[0,1] and contains only Lipschitz continuous functions. Thus, [Ax](s)[Ax](s) can be continuously extended to s[0,1]s\in[0,1]. An application of formula (5) yields for the singular values

σn(J2)M(L2(0,1))σn(A)σn(A)C(L2(0,1))σn(J)(n).\sigma_{n}(J^{2})\leq\|M\|_{\scriptscriptstyle\mathcal{L}(L^{2}(0,1))}\,\sigma_{n}(A)\leq\sigma_{n}(A)\leq\|C\|_{\scriptscriptstyle\mathcal{L}(L^{2}(0,1))}\,\sigma_{n}(J)\qquad(n\in\mathbb{N})\,.

Together with (15) we obtain that there exist positive constants K1K_{1} and K2K_{2} such that

(19) K1n2σn(A)K2n(n).\frac{K_{1}}{n^{2}}\leq\sigma_{n}(A)\leq\frac{K_{2}}{n}\qquad(n\in\mathbb{N})\,.

Consequently, we know at this point only that the interval of ill-posedness of the operator AA from (13) is a subset of the interval [1,2][1,2]. However, taking into account that AA is a Hilbert-Schmidt operator, we will be able to improve the order of the upper bound of (19) in Section 4.

3. Hilbert-Schmidt property and kernel smoothness

As one sees from (13), A=CJ:L2(0,1)L2(0,1)A=C\circ J:L^{2}(0,1)\to L^{2}(0,1) is linear Volterra integral operator with quadratically integrable kernel and hence a Hilbert-Schmidt operator with Hilbert-Schmidt norm square

(20) AHS2=010s(sts)2𝑑t𝑑s=16.\|A\|_{HS}^{2}=\int\limits_{0}^{1}\int\limits_{0}^{s}\left(\frac{s-t}{s}\right)^{2}dtds=\frac{1}{6}\,.

Taking into account that the adjoint operator A:L2(0,1)L2(0,1)A^{*}:L^{2}(0,1)\to L^{2}(0,1) is of the form

[Ay](t)=t1stsy(s)𝑑s(0t1),[A^{*}y](t)=\int\limits_{t}^{1}\frac{s-t}{s}\,y(s)ds\quad(0\leq t\leq 1)\,,

we derive the structure of the symmetric kernel k(s,t)k(s,t) of the self-adjoint Fredholm integral operator

[AAw](t)=01k(t,s)w(s)𝑑s(0t,s1)[A^{*}A\,w](t)=\int_{0}^{1}k(t,s)\,w(s)ds\quad(0\leq t,s\leq 1)

as

(21) k(t,s)\displaystyle k(t,s) =max(t,s)1(τtτ)(τsτ)𝑑τ\displaystyle=\int\limits_{\max(t,s)}^{1}\left(\frac{\tau-t}{\tau}\right)\,\left(\frac{\tau-s}{\tau}\right)\,d\tau
={1sts+tlns+slns+t(0<ts1)1stt+tlnt+slnt+s(0<s<t1).\displaystyle={\scriptstyle\left\{\begin{array}[]{ll}1-st-s+t\ln s+s\ln s+t&\quad(0<t\leq s\leq 1)\\ 1-st-t+t\ln t+s\ln t+s&\quad(0<s<t\leq 1)\end{array}\right.}\,.

It is well-known that decay rates of the singular values of a compact linear operator grows in general with the smoothness of the kernel. Unfortunately, the kernel (21) is continuous on the unit square only with the exception of the origin (0,0)(0,0), where a pole arises. Therefore, usually applied assertions on kernel smoothness (cf., e.g., [2, 16, 17]) cannot be exploited to estimate the asymptotics of the eigenvalues λn(AA)\lambda_{n}(A^{*}A) and in the same manner the asymptotics of the singular values σn(A)=λn(AA)\sigma_{n}(A)=\sqrt{\lambda_{n}(A^{*}A)} of AA. On the other hand, by twice differentiation of the function [AAw](t)(0t1)[A^{*}Aw](t)\,(0\leq t\leq 1) as

(t)\displaystyle(t) =t1sts0ssτsw(τ)𝑑τ𝑑s\displaystyle=\int_{t}^{1}\frac{s-t}{s}\int_{0}^{s}\frac{s-\tau}{s}w(\tau)d\tau ds
[AAw](t)\displaystyle[A^{*}Aw]^{\prime}(t) =t1(1s)0ssτsw(τ)𝑑τ𝑑s\displaystyle=\int_{t}^{1}(-\frac{1}{s})\int_{0}^{s}\frac{s-\tau}{s}w(\tau)d\tau ds
=t10ssτs2w(τ)𝑑τ𝑑s\displaystyle=-\int_{t}^{1}\int_{0}^{s}\frac{s-\tau}{s^{2}}w(\tau)d\tau ds
[AAw]′′(t)\displaystyle[A^{*}Aw]^{\prime\prime}(t) =0ttτt2w(τ)𝑑τ,\displaystyle=\int_{0}^{t}\frac{t-\tau}{t^{2}}w(\tau)d\tau,

we have an integro-differential final value problem

(22) 0ttτt2w(τ)𝑑τ=λw′′(t)(0<t<1),w(1)=w(1)=0.\int\limits_{0}^{t}\frac{t-\tau}{t^{2}}\,w(\tau)d\tau=\lambda\,w^{\prime\prime}(t)\;\;(0<t<1),\quad w(1)=w^{\prime}(1)=0\,.

Achieving an explicit analytical solution of the eigenvalues λn(AA)\lambda_{n}(A^{*}A) and corresponding eigenfunctions wL2(0,1)w\in L^{2}(0,1) seems to be very difficult.

However, we are still able to calculate numerical approximations of the eigenvalues λn(AA)\lambda_{n}(A^{*}A) from (22) for small nn. We have made use of the technique of finite difference discretization with a specific rectangular rule. In this context, the unit interval [0,1][0,1] had been divided into \ell partitions with the uniform length h=1/h=1/\ell. The function values w(τ)w(\tau) are represented by discrete values wi:=w(ih)(i=0,1,,)w_{i}:=w(i*h)\>(i=0,1,\dots,\ell). The discrete counterpart of its second derivative w′′(t)w^{\prime\prime}(t) is considered as wj+12wj+wj1h2(j=1,2,,1)\frac{w_{j+1}-2w_{j}+w_{j-1}}{h^{2}}\;(j=1,2,...,\ell-1). The equation (22) can be written in a discrete form as

hi=0j1(ji)h(jh)2wi=λwj+12wj+wj1h2j=1,,1.h\sum_{i=0}^{j-1}\frac{(j-i)h}{(jh)^{2}}w_{i}=\lambda\frac{w_{j+1}-2w_{j}+w_{j-1}}{h^{2}}\qquad j=1,\dots,\ell-1.

with the boundary conditions of as w=0w_{\ell}=0 and w=w1w_{\ell}=w_{\ell-1}.

Now we need to find those positive values λ\lambda such that the determinant of the (+1)×(+1)(\ell+1)\times(\ell+1) square matrix

(h2λ2λλ00000024h214h2λ2λλ0000039h229h219h2λ2λλ0000ii2h2i1i2h21i2h2λ2λλ0001(1)2h22(1)2h22(1)2h21(1)2h2λ2λλ0000000010000011)\left(\begin{array}[]{lllllllll}h^{2}-\lambda&2\lambda&-\lambda&0&0&0&0&0&0\\ \tfrac{2}{4}h^{2}&\tfrac{1}{4}h^{2}-\lambda&2\lambda&-\lambda&0&0&0&0&0\\ \tfrac{3}{9}h^{2}&\tfrac{2}{9}h^{2}&\tfrac{1}{9}h^{2}-\lambda&2\lambda&-\lambda&0&0&0&0\\ \vdots&\ddots&\ddots&\cdots&\cdots&\cdots&\ddots&\ddots&\vdots\\ \tfrac{i}{i^{2}}h^{2}&\tfrac{i-1}{i^{2}}h^{2}&\cdots&\tfrac{1}{i^{2}}h^{2}-\lambda&2\lambda&-\lambda&0&0&0\\ \vdots&\ddots&\ddots&\cdots&\cdots&\cdots&\ddots&\ddots&\vdots\\ \tfrac{\ell-1}{(\ell-1)^{2}}h^{2}&\tfrac{\ell-2}{(\ell-1)^{2}}h^{2}&\cdots&\tfrac{2}{(\ell-1)^{2}}h^{2}&\tfrac{1}{(\ell-1)^{2}}h^{2}-\lambda&2\lambda&-\lambda&0&0\\ 0&0&\cdots&\cdots&0&0&0&0&1\\ 0&0&\cdots&\cdots&0&0&0&1&-1\\ \end{array}\right)

equals zero, which gives the sequence of discretized eigenvalues λn(n=1,2,,+1)\lambda_{n}\;(n=1,2,...,\ell+1) in non-increasing order. The following Figure 1 displays in a double logarithmic representation the decay of calculated eigenvalues for the case =20\ell=20. From that curve an asymptotics of the form λn(AA)n4\lambda_{n}(A^{*}A)\asymp n^{-4} can be predicted, which would coincide with the lower bound of the inequality chain (19).

Refer to caption
Figure 1. Double logarithmic representation of the computationally approximated eigenvalues λn(AA)\lambda_{n}(A^{*}A) for small indices nn.

This computational prediction will be confirmed by the analytical study of the subsequent section.

4. Improved upper bounds

To reduce the asymptotics gap, which has been opened by the inequality chain (19), we reuse the Hilbert-Schmidt operator technique introduced in Section 5 of the recent article [5]. First, we briefly summarize the main ideas of this approach. For the Hilbert-Schmidt operator A:L2(0,1)L2(0,1)A:L^{2}(0,1)\to L^{2}(0,1) with singular system {σn(A),un=un(A),vn=vn(A)}n=1\{\sigma_{n}(A),u_{n}=u_{n}(A),v_{n}=v_{n}(A)\}_{n=1}^{\infty} we have for the Hilbert-Schmidt norm square

AHS2=n=1σn2(A).\|A\|_{HS}^{2}=\sum\limits_{n=1}^{\infty}\sigma_{n}^{2}(A)\,.

We start with the following proposition, which can be found with a complete proof as Proposition 3 in [5] when taking into account Theorem 15.5.5 from [14].

Proposition 1.

Let {ei}i=1\{e_{i}\}_{i=1}^{\infty} denote an arbitrary orthonormal basis in L2(0,1)L^{2}(0,1) and QnQ_{n} denote the orthogonal projections onto the nn-dimensional subspace span{e1,,en}\operatorname{span}\left\{e_{1},\dots,e_{n}\right\} of L2(0,1)L^{2}(0,1). Moreover, let SnS_{n} denote the orthogonal projection onto the specific nn-dimensional subspace span{u1,,un}\operatorname{span}\left\{u_{1},\dots,u_{n}\right\} of the first nn singular functions of AA. Then we have that

(23) i=n+1σi2(A)=A(ISn)HS2A(IQn)HS2=i=n+1AeiL2(0,1)2.\sum_{i=n+1}^{\infty}\sigma^{2}_{i}(A)=\|A(I-S_{n})\|_{HS}^{2}\leq\|A(I-Q_{n})\|_{HS}^{2}=\sum_{i=n+1}^{\infty}\|Ae_{i}\|^{2}_{L^{2}(0,1)}\,.

The following technical lemma and its proof can be found in [5, Lemma 4].

Lemma 2.

Let {si}i=1\{s_{i}\}_{i=1}^{\infty} be a non-increasing sequence of positive numbers, and let ω\omega be a positive number. Suppose that there is a constant K<K<\infty such that i=n+1si2Kn2ω\sum\limits_{i=n+1}^{\infty}s_{i}^{2}\leq Kn^{-2\omega} for n=1,2,n=1,2,\dots. Then there also exists a constant K^<\hat{K}<\infty such that si2K^i(1+2ω)s_{i}^{2}\leq\hat{K}i^{-(1+2\omega)} for i=1,2,.i=1,2,\dots.

Furthermore, there is another useful property of orthogonal polynomials, shown in the proof of Lemma 7.4 in [13, Chapter 2, p.69–70].

Lemma 3.

Let {pj}j=0,1,\{p_{j}\}_{j=0,1,\dots} be a system of orthogonal polynomials with respect to some measure ν\nu and satisfy the orthogonality relation

(24) pj(t)pk(t)𝑑ν(t)=hjδjk,\int p_{j}(t)p_{k}(t)d\nu(t)=h_{j}\delta_{jk},

where the value hjh_{j} depends on the index jj and δjk\delta_{jk} denotes the Kronecker delta, equal to 11 if j=kj=k and to 0 otherwise. Then, the polynomials qjq_{j} defined by

(25) qj(t)=pj(t)pj(τ)tτ𝑑ν(τ)q_{j}(t)=\int\frac{p_{j}(t)-p_{j}(\tau)}{t-\tau}\,d\nu(\tau)

possess the degree of j1j-1. Furthermore, we have the identity

pi(t)pi(τ)tτ=j=0i1pj(τ)qi(τ)pi(τ)qj(τ)hjpj(t).\frac{p_{i}(t)-p_{i}(\tau)}{t-\tau}=\sum_{j=0}^{i-1}\frac{p_{j}(\tau)q_{i}(\tau)-p_{i}(\tau)q_{j}(\tau)}{h_{j}}\,p_{j}(t).

Based on the above Proposition 1 as well as on Lemmas 2 and 3 we can now improve the inequality chain (19) for updating the asymptotics of the singular values σn(A)\sigma_{n}(A) of the composite operator A=CJA=C\circ J under consideration.

Theorem 1.

For the composition A=CJ:L2(0,1)L2(0,1)A=C\circ J:L^{2}(0,1)\to L^{2}(0,1) of the simple integration operator J:L2(0,1)L2(0,1)J:L^{2}(0,1)\to L^{2}(0,1) from (6) followed by the continuous Cesàro operator C:L2(0,1)L2(0,1)C:L^{2}(0,1)\to L^{2}(0,1) from (12), the asymptotics

(26) σn(A)1n2\sigma_{n}(A)\asymp\frac{1}{n^{2}}

holds true. In other words, the operator AA possesses exactly the degree two of ill-posedness in the sense of Definition 1.

Proof.

We are going to apply Proposition 1 for A=CJA=C\circ J with the specific orthonormal basis {Pi}i=1\{P_{i}\}_{i=1}^{\infty} in L2(0,1)L^{2}(0,1) of normalized shifted Legendre Polynomials, which are defined by means of the standard orthogonal Legendre Polynomials {Li(t)}i=0\{L_{i}(t)\}_{i=0}^{\infty} on [1,1][-1,1] with the relationship

Pi+1(t)=2i+1Li(2t1).P_{i+1}(t)=\sqrt{2i+1}\,L_{i}(2t-1)\,.

In this proof, we set t~:=2t1\tilde{t}:=2t-1, s~:=2s1\tilde{s}:=2s-1 for t,s[0,1]t,s\in[0,1] and we will use the well-known three-term recurrence relation

(27) (i+1)Li+1(t~)=(2i+1)t~Li(t~)iLi1(t~),(i+1)\,L_{i+1}(\tilde{t})=(2i+1)\,\tilde{t}\,L_{i}(\tilde{t})-i\,L_{i-1}(\tilde{t}),

as well as the initial conditions

L0(t~)=1,L1(t~)=t~.L_{0}(\tilde{t})=1,\qquad L_{1}(\tilde{t})=\tilde{t}.

Firstly, we obtain

[JPi+1](s)=0sPi+1(t)𝑑t=Li+1(s~)Li1(s~)22i+1.[JP_{i+1}](s)=\int_{0}^{s}P_{i+1}(t)dt=\frac{L_{i+1}(\tilde{s})-L_{i-1}(\tilde{s})}{2\sqrt{2i+1}}.

Then we define

fi(t):=Li(t~)Li(1)2t=Li(t~)Li(1)t~+1f_{i}({t}):=\frac{L_{i}(\tilde{t})-L_{i}(-1)}{2t}=\frac{L_{i}(\tilde{t})-L_{i}(-1)}{\tilde{t}+1}

and verify

[CPi+1](s)=1s0sPi+1(t)𝑑t=fi+1(s)fi1(s)2i+1.[CP_{i+1}](s)=\frac{1}{s}\int_{0}^{s}P_{i+1}(t)dt=\frac{f_{i+1}({s})-f_{i-1}({s})}{\sqrt{2i+1}}.

Applying Lemma 3 and identifying the orthogonal polynomials {pj}j=0,1,\{p_{j}\}_{j=0,1,\dots} as {Lj}j=0,1,\{L_{j}\}_{j=0,1,\dots} on [1,1][-1,1] such that hj=22j+1h_{j}=\frac{2}{2j+1} and Lj(1)=(1)jL_{j}(-1)=(-1)^{j}, one can check that qj(x)q_{j}(x) from (25) satisfies the recursion relation (27). Moreover, the equation

qj(1)=2(1)j1Hjq_{j}(-1)=2(-1)^{j-1}H_{j}

holds true, where Hj:=k=1j1kH_{j}:=\sum_{k=1}^{j}\frac{1}{k}. Consequently, we have

fi(t)=j=0i1(1)i+j1(2j+1)(HiHj)Lj(t~)f_{i}({t})=\sum_{j=0}^{i-1}(-1)^{i+j-1}(2j+1)(H_{i}-H_{j})L_{j}(\tilde{t})

and

ci(s)=[CPi+1](s)=(1)i2i+1i(i+1)j=0i1(1)j(2j+1)Lj(s~)+2i+1i+1Li(s~).c_{i}(s)=[CP_{i+1}](s)=\frac{(-1)^{i}\sqrt{2i+1}}{i(i+1)}\sum_{j=0}^{i-1}(-1)^{j}(2j+1)L_{j}(\tilde{s})+\frac{\sqrt{2i+1}}{i+1}L_{i}(\tilde{s}).

Finally, we calculate

ai(s)=[APi+1](s)=[CJPi+1](s)=ci+1(s)2i+3ci1(s)2i122i+1a_{i}(s)=[AP_{i+1}](s)=[CJP_{i+1}](s)=\frac{\frac{c_{i+1}(s)}{\sqrt{2i+3}}-\frac{c_{i-1}({s})}{\sqrt{2i-1}}}{2\sqrt{2i+1}}

and obtain the complex expression

ai(s)\displaystyle a_{i}(s) =(1)i2i+1(i1)i(i+1)(i+2):=k1j=0i2(1)j(2j+1)Lj(s~)\displaystyle=\underbrace{(-1)^{i}\frac{\sqrt{2i+1}}{(i-1)i(i+1)(i+2)}}_{:=k_{1}}\sum_{j=0}^{i-2}(-1)^{j}(2j+1)L_{j}(\tilde{s})
+i24i22i(i+1)(i+2)2i+1:=k2Li1(s~)\displaystyle+\underbrace{\frac{i^{2}-4i-2}{2i(i+1)(i+2)\sqrt{2i+1}}}_{:=k_{2}}L_{i-1}(\tilde{s})
2i+12(i+1)(i+2):=k3Li(s~)+12(i+2)2i+1:=k4Li+1(s~).\displaystyle-\underbrace{\frac{\sqrt{2i+1}}{2(i+1)(i+2)}}_{:=k_{3}}L_{i}(\tilde{s})+\underbrace{\frac{1}{2(i+2)\sqrt{2i+1}}}_{:=k_{4}}L_{i+1}(\tilde{s})\,.

Since all terms above are orthogonal, the squared L2L^{2}-norm for aia_{i} attains the form

APi+1L2(0,1)2\displaystyle\|AP_{i+1}\|_{L^{2}(0,1)}^{2} =aiL2(0,1)2=01ai2(s)𝑑s=1211ai2(s~)𝑑s~\displaystyle=\|a_{i}\|_{L^{2}(0,1)}^{2}=\int_{0}^{1}a_{i}^{2}(s)ds=\frac{1}{2}\int_{-1}^{1}a_{i}^{2}(\tilde{s})d\tilde{s}
=k12j=0i2(2j+1)+k2212i1+k3212i+1+k4212i+3\displaystyle=k_{1}^{2}\sum_{j=0}^{i-2}(2j+1)+k_{2}^{2}\frac{1}{2i-1}+k_{3}^{2}\frac{1}{2i+1}+k_{4}^{2}\frac{1}{2i+3}
=2i+1i2(i+1)2(i+2)2+(i24i2)24i2(i+1)2(i+2)2(2i+1)(2i1)\displaystyle=\frac{2i+1}{i^{2}(i+1)^{2}(i+2)^{2}}+\frac{(i^{2}-4i-2)^{2}}{4i^{2}(i+1)^{2}(i+2)^{2}(2i+1)(2i-1)}
+14(i+1)2(i+2)2+14(i+2)2(2i+1)(2i+3)\displaystyle+\frac{1}{4(i+1)^{2}(i+2)^{2}}+\frac{1}{4(i+2)^{2}(2i+1)(2i+3)}
=32i(i+1)(2i1)(2i+3).\displaystyle=\frac{3}{2i(i+1)(2i-1)(2i+3)}\,.

and

APiL2(0,1)2=32i(i1)(2i3)(2i+1).\|AP_{i}\|_{L^{2}(0,1)}^{2}=\frac{3}{2i(i-1)(2i-3)(2i+1)}.

Now we can apply Proposition 1 immediately and derive that

i=n+1σi2(A)i=n+1APiL2(0,1)2=18n32nKn3\sum_{i=n+1}^{\infty}\sigma_{i}^{2}(A)\leq\sum_{i=n+1}^{\infty}\|AP_{i}\|_{L^{2}(0,1)}^{2}=\frac{1}{8n^{3}-2n}\leq Kn^{-3}

for a constant K<K<\infty. According to Lemma 2 by identifying sis_{i} as σi(A)\sigma_{i}(A), there exists a positive constant K^\hat{K} such that σi2(A)K^2i4\sigma_{i}^{2}(A)\leq\hat{K}^{2}\,i^{-4} and consequently

σi(A)K^i2.\sigma_{i}(A)\leq\hat{K}i^{-2}.

Taking into account the estimates of (19) with focus on the lower bound, this shows the asymptotics

σn(A)1n2\sigma_{n}(A)\asymp\frac{1}{n^{2}}

and completes the proof of the theorem. ∎

It is interesting to notice that the set {Pi}i=1\{P_{i}\}_{i=1}^{\infty} of the (shifted) Legendre polynomials as orthonormal basis seems to be sufficiently close to the eigensystem {ui(A)}i=1\{u_{i}(A)\}_{i=1}^{\infty} as part of the singular system of the Hilbert-Schmidt operator AA. As the following remark indicates, the corresponding eigensystem {ui(J)}i=1\{u_{i}(J)\}_{i=1}^{\infty} from (7) of the integration operator JJ does not reach the best result to determine the upper bound of the rate σn(A)\sigma_{n}(A).

Remark 1.

Applying Proposition 1 for A=CJA=C\circ J with the specific orthonormal basis {ei}i=1\{e_{i}\}_{i=1}^{\infty} in L2(0,1)L^{2}(0,1) of the form

ei(t):=ui(J)=2cos((i12)πt)(0t1)e_{i}(t):=u_{i}(J)=\sqrt{2}\cos((i-\frac{1}{2})\pi t)\quad(0\leq t\leq 1)

(see the singular system (7)), it is only possible to obtain the degree of ill-posedness for operator AA as

(28) K¯1n2σn(A)K¯2n3/2(n),\frac{\bar{K}_{1}}{n^{2}}\leq\sigma_{n}(A)\leq\frac{\bar{K}_{2}}{n^{3/2}}\qquad(n\in\mathbb{N})\,,

here, where K¯1\bar{K}_{1} and K¯2\bar{K}_{2} are some positive constants.

Remark 2.

In [7] it was shown that multiplication operators M:L2(0,1)L2(0,1)M:L^{2}(0,1)\to L^{2}(0,1) from (9) with multiplier functions m(t)=tη(η>0)m(t)=t^{\eta}\;(\eta>0) do not change the ill-posedness degree of JJ when they occur in a composition MJM\circ J. From the present study, we can see that this effect is also observable for the composition J2=M(CJ)J^{2}=M\circ(C\circ J) (cf. (18)). This means that such multiplication operator MM in the special case η=1\eta=1 also does not amend the ill-posedness degree when moving from to A=CJA=C\circ J to J2=MAJ^{2}=M\circ A. In coincidence, the following Figure 2

Refer to caption
Figure 2. Singular values of operators A,J,A,J, and J2J^{2}

illustrates the logarithmic plot of singular values of ×\ell\times\ell discretization matrices with =104\ell=10^{4} of the operators A=CJA=C\circ J, JJ and J2J^{2}, which are calculated based on the MATLAB routine svd.

Acknowledgment

YD and BH are supported by the German Science Foundation (DFG) under grant HO 1454/13-1 (Project No. 453804957).

Appendix

Proof of Lemma 1: To prove the non-compactness of CC it is enough to find a sequence {xn}n=1\{x_{n}\}_{n=1}^{\infty} in L2(0,1)L^{2}(0,1) such that xn0x_{n}\rightharpoonup 0 (weak convergence in L2(0,1)L^{2}(0,1)), but CxnL2(0,1)↛0\|Cx_{n}\|_{L^{2}(0,1)}\not\to 0 as nn\to\infty. In this context, we use the bounded sequence xn(t)=nχ(0,1n](t)(0t1)x_{n}(t)=\sqrt{n}\,\chi_{(0,\frac{1}{n}]}(t)\;(0\leq t\leq 1) with xnL2(0,1)=1\|x_{n}\|_{L^{2}(0,1)}=1 for all nn\in\mathbb{N}. Then we have, for all 0<s10<s\leq 1 and sufficiently large nNn\in N, that

0sxn(t)𝑑t=01/nn𝑑t=1n,\int_{0}^{s}x_{n}(t)\,dt=\int_{0}^{1/n}\sqrt{n}\,dt=\frac{1}{\sqrt{n}}\,,

which tends to zero as nn\to\infty. This shows (cf., e.g., [11, Satz 10, p. 151]) the claimed weak convergence. On the other hand, we have

[Cxn](s)={n(0<s1n)1ns(0<1ns1).[Cx_{n}](s)=\left\{\begin{array}[]{ll}\sqrt{n}&\quad(0<s\leq\frac{1}{n})\\ \frac{1}{\sqrt{n}s}&\quad(0<\frac{1}{n}\leq s\leq 1)\end{array}\right.\,.

Hence

CxnL2(0,1)2=01/nn𝑑s+1/n11ns2𝑑s2asn,\|Cx_{n}\|_{L^{2}(0,1)}^{2}=\int_{0}^{1/n}nds+\int_{1/n}^{1}\frac{1}{ns^{2}}ds\to 2\quad\text{as}\;\;n\to\infty\,,

and CC is not compact.

In order to prove the unboundedness of C1C^{-1}, we can exploit the sequence xn(t)=ncos(nt)(0t1)x_{n}(t)=\sqrt{n}\,\cos(nt)\;(0\leq t\leq 1) together with the associated sequence yn(s):=[Cxn](s)=sin(ns)ns(0s1)y_{n}(s):=[Cx_{n}](s)=\frac{\sin(ns)}{\sqrt{n}s}\;(0\leq s\leq 1) possessing the limit

ynL2(0,1)2=CxnL2(0,1)2=01sin2(ns)ns2π2<asn.\|y_{n}\|^{2}_{L^{2}(0,1)}=\|Cx_{n}\|^{2}_{L^{2}(0,1)}=\int_{0}^{1}\frac{\sin^{2}(ns)}{ns^{2}}\to\sqrt{\frac{\pi}{2}}<\infty\quad\text{as}\;\;n\to\infty\,.

Now the property

xnL2(0,1)2=C1ynL2(0,1)2=01ncos2(nt)𝑑t=n+sin(n)cos(n)2\|x_{n}\|^{2}_{L^{2}(0,1)}=\|C^{-1}y_{n}\|^{2}_{L^{2}(0,1)}=\int_{0}^{1}n\cos^{2}(nt)dt=\frac{n+\sin(n)\cos(n)}{2}\to\infty

for nn\to\infty indicates that C1C^{-1} cannot be bounded, which completes the proof of the lemma.

References

  • [1] A. Brown, P. R. Halmos and A. L. Shields, Cesàro operators, Acta Sci. Math. (Szeged), 26:125–137, 1965.
  • [2] S.-H. Chang, A generalization of a theorem of Hille and Tamarkin with applications, Proc. London Math. Soc. (3)2:22–29, 1952.
  • [3] M. Freitag and B. Hofmann, Analytical and numerical studies on the influence of multiplication operators for the ill-posedness of inverse problems, J. Inverse Ill-Posed Probl., 13(2):123–148, 2005.
  • [4] D. Gerth, B. Hofmann, C. Hofmann and S. Kindermann, The Hausdorff moment problem in the light of ill-posedness of type I, Eurasian Journal of Mathematical and Computer pplications, 9(2):57–87, 2021.
  • [5] B. Hofmann and P. Mathé, The degree of ill-posedness of composite linear ill-posed problems with focus on the impact of the non-compact Hausdorff moment operator, Electronic Transactions on Numerical Analysis, 57:1-16, 2022.
  • [6] B. Hofmann and U. Tautenhahn, On ill-posedness measures and space change in Sobolev scales, Journal for Analysis and its Applications (ZAA), 16(4):979–1000, 1997.
  • [7] B. Hofmann and L. von Wolfersdorf, Some results and a conjecture on the degree of ill-posedness for integration operators with weights, Inverse Problems, 21(2):427–433, 2005.
  • [8] B. Hofmann and L. von Wolfersdorf, A new result on the singular value asymptotics of integration oprators with weights, J. Integral Equations Appl., 21(2):281–295, 2009.
  • [9] M. Lacruz, F. León-Saavedra, S. Petrovic and O. Zabeti, Extended eigenvalues for Cesàro operators, J. Math. Anal. Appl., 429(2):623–657, 2015.
  • [10] G. M. Leibowitz, Spectra of finite range Cesàro operators, Acta Sci. Math. (Szeged), 35:27–29, 1973.
  • [11] L. A. Ljusternik and W. I. Sobolew, Elemente der Funktionalanalysis, Mathematical Textbooks and Monographs, Part I: Mathematical Textbooks, Vol. 8, Akademie-Verlag, Berlin, 1968.
  • [12] M. Z. Nashed, A new approach to classification and regularization of ill-posed operator equations, In: Inverse and Ill-posed Problems Sankt Wolfgang, 1986, Vol. 4 of Notes Rep. Math. Sci. Engrg. (Eds.: H. W. Engl and C. W. Groetsch), Academic Press, Boston, 1987, pp. 53–75.
  • [13] E. M. Nikishin and V. N. Sorokin, Rational Approximations and Orthogonality, Translations of Mathematical Monographs, Vol. 92, American Mathematical Society, Providence, 1991.
  • [14] A. Pietsch, Operator Ideals, Mathematical Monographs, Vol. 16, Dt. Verlag der Wissenschaften, Berlin, 1978.
  • [15] R. Ramlau, Ch. Koutschan and B. Hofmann, On the singular value decomposition of nn-fold integration operators, In: Inverse Problems and Related Topics, Vol. 310 of Springer Proc. Math. Stat. (Eds.: J. Cheng, S. Lu and M. Yamamoto), Springer, Singapore, 2020, pp. 237–256.
  • [16] J. B. Reade, Eigenvalues of Lipschitz kernels, Math. Proc. Cambridge Philos. Soc., 93(1):135–140, 1983.
  • [17] J. B. Reade, Eigenvalues of positive definite kernels, SIAM J. Math. Anal., 14(1):152–157, 1983.
  • [18] Vu Kim Tuan and R. Gorenflo, Asymptotics of singular values of fractional integral operators, Inverse Problems, 10(4):949–955, 1994.