Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Russian Mathematical Surveys
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Archive
Impact factor
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Uspekhi Mat. Nauk:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Russian Mathematical Surveys, 2024, Volume 79, Issue 6, Pages 975–990
DOI: https://doi.org/10.4213/rm10186e
(Mi rm10186)
 

On greedy approximation in complex Banach spaces

A. V. Gasnikovabc, V. N. Temlyakovdbef

a Ivannikov institute for System Programming of Russian Academy of Sciences, Moscow, Russia
b Steklov Mathematical Institute of Russian Academy of Sciences, Moscow, Russia
c Innopolis University, Innopolis, Russia
d University of South Carolina, Columbia, USA
e Lomonosov Moscow State University, Moscow, Russia
f Moscow Center for Fundamental and Applied Mathematics, Moscow, Russia
References:
Abstract: The general theory of greedy approximation with respect to arbitrary dictionaries is well developed in the case of real Banach spaces. Recently some results proved for the Weak Chebyshev Greedy Algorithm (WCGA) in the case of real Banach spaces were extended to the case of complex Banach spaces. In this paper we extend some of the results known in the real case for greedy algorithms other than the WCGA to the case of complex Banach spaces.
Bibliography: 25 titles.
Keywords: complex Banach space, greedy approximation, weak greedy algorithm with free relaxation, greedy algorithm with weakness and relaxation, incremental algorithm.
Funding agency Grant number
Ministry of Science and Higher Education of the Russian Federation 075-15-2024-529
This work was supported by Ministry of Science and Higher Education of the Russian Federation (grant no. 075-15-2024-529).
Received: 22.08.2024
Bibliographic databases:
Document Type: Article
UDC: 517.5+519.6
MSC: Primary 41A65; Secondary 46B15, 46B20
Language: English
Original paper language: English

1. Introduction

Sparse and greedy approximation theory is important in applications. This theory has actively been developing for about 40 years. In the beginning researchers were interested in approximation in Hilbert spaces (see, for instance, [12], [15], [17], [18], [1], [7], [4], and [2]; for a detailed history, see [22]). Later, this theory was extended to the case of real Banach spaces (see, for instance, [10], [20], [13], and [5]; for a detailed history see [22] and [23]) and to the case of convex optimization (see, for instance, [3], [11], [16], [19], [24], [25], [8], [14], and [6]). The reader can find some open problems in the book [22].

We study greedy approximation with respect to arbitrary dictionaries in Banach spaces. This general theory is well developed in the case of real Banach spaces (see the book [22]). Recently (see [9]) some results proved in the case of real Banach spaces were extended to the case of complex Banach spaces. The Weak Chebyshev Greedy Algorithm (WCGA) was studied in [9]. In this paper we discuss greedy algorithms other than the WCGA. For them we extend some results known in the real case to the case of complex Banach spaces. We use the same main ideas as in the real case: first, we prove a recurrent inequality between the norms fm and fm1 of the residuals of the algorithm; second, we analyze this recurrent inequality. The step from the real case to the complex case requires some technical modifications in the proofs of the corresponding recurrent inequalities. For the reader’s convenience we present a detailed analysis here.

Let X be a (real or complex) Banach space with norm . We say that a set D of elements (functions) from X is a dictionary if each gD has a norm bounded by one (g1) and the closure of spanD is X. It is convenient for us to consider, along with the dictionary D, its symmetrization. In the case of real Banach spaces we set

D±:={±g:gD}.
In the case of complex Banach spaces we set
D:={eiθg:gD, θ[0,2π)}.
In the above notation D symbol stands for the unit circle.

Following the standard notation, set

Ao1(D):={fX:f=j=1ajgj, j=1|aj|1, gjD, j=1,2,}
and denote the closure (in X) of Ao1(D) by A1(D). Then it is clear that A1(D) is the closure (in X) of the convex hull of D± in the real case and of D in the complex case. Also, denote by conv(D) the closure (in X) of the convex hull of D.

We introduce a new norm, associated with the dictionary D, in the dual space X by the formula

FD:=sup
Note that, clearly, in the real case \|F\|_{\mathcal D}=\sup_{\phi\in{\mathcal D}^\pm}F(\phi) and in the complex case \|F\|_{\mathcal D}= \sup_{\phi\in{\mathcal D}^{\circ}}\operatorname{Re}(F(\phi)).

In this paper we study greedy algorithms with respect to {\mathcal D}. For a non-zero element h\in X we let F_h denote a norming (peak) functional for h:

\begin{equation*} \|F_h\|=1,\qquad F_h(h)=\|h\|. \end{equation*} \notag
The existence of such a functional is guaranteed by the Hahn–Banach theorem.

2. Definitions and lemmas

We consider here approximation in uniformly smooth Banach spaces. For a Banach space X we define its modulus of smoothness

\begin{equation*} \rho(u):=\rho(u,X):=\sup_{\|x\|=\|y\|=1} \biggl(\frac{1}{2}(\|x+uy\|+\|x-uy\|)-1\biggr). \end{equation*} \notag
A uniformly smooth Banach space is one with the property
\begin{equation*} \lim_{u\to 0}\frac{\rho(u)}{u}=0. \end{equation*} \notag
It is easy to see that for any Banach space X its modulus of smoothness \rho(u) is an even convex function satisfying the inequalities
\begin{equation*} \max(0,u-1)\leqslant \rho(u)\leqslant u,\qquad u\in (0,\infty). \end{equation*} \notag
It is well known (see, for instance, [10]) that in the case X=L_p, 1\leqslant p < \infty, we have
\begin{equation} \rho(u,L_p) \leqslant \begin{cases} \dfrac{u^p}{p} & \text{if}\ \ 1\leqslant p\leqslant 2, \vphantom{\Biggl\}} \\ \dfrac{(p-1)u^2}{2} & \text{if}\ \ 2\leqslant p<\infty. \end{cases} \end{equation} \tag{2.1}

We note that from the definition of the modulus of smoothness we obtain the following inequality (in the real case see, for instance, [22], p. 336).

Lemma 2.1. Let x\ne0. Then

\begin{equation} 0\leqslant \|x+uy\|-\|x\|-\operatorname{Re}(uF_x(y))\leqslant 2\|x\|\rho\biggl(u\,\frac{\|y\|}{\|x\|}\biggr). \end{equation} \tag{2.2}

Proof. We have
\begin{equation*} \|x+uy\|\geqslant |F_x(x+uy)|\geqslant \operatorname{Re}(F_x(x+uy))=\|x\|+\operatorname{Re}(uF_x(y)). \end{equation*} \notag
This proves the left-hand inequality. Next, from the definition of the modulus of smoothness it follows that
\begin{equation} \|x+uy\|+\|x-uy\|\leqslant 2\|x\|\biggl(1+\rho\biggl(u\,\frac{\|y\|}{\|x\|}\biggr)\biggr). \end{equation} \tag{2.3}
Also,
\begin{equation} \|x-uy\|\geqslant |F_x(x-uy)| \geqslant \operatorname{Re}(F_x(x-uy))= \|x\|-\operatorname{Re}(uF_x(y)). \end{equation} \tag{2.4}
Combining (2.3) and (2.4) we obtain
\begin{equation*} \|x+uy\|\leqslant \|x\|+\operatorname{Re}(uF_x(y))+ 2\|x\|\rho\biggl(u\,\frac{\|y\|}{\|x\|}\biggr). \end{equation*} \notag
This proves the second inequality. \Box

We will use the following three simple and well-known lemmas. In the case of real Banach spaces these lemmas were proved, for instance, in [22], Chap. 6, pp. 342–343.

Lemma 2.2. Let X be a uniformly smooth Banach space and L be a finite-dimensional subspace of X. For any f\in X\setminus L let f_L denote the best approximant of f in L. Then

\begin{equation*} F_{f-f_L}(\phi)=0 \end{equation*} \notag
for any \phi \in L.

Proof. Let us assume the contrary: there is \phi \in L such that \|\phi\|=1 and
\begin{equation*} |F_{f-f_L}(\phi)|=\beta >0. \end{equation*} \notag
Denote by \nu the complex conjugate of \operatorname{sign}(F_{f-f_L}(\phi)), where \operatorname{sign} z:=z/|z| for z\ne 0. Then \nu F_{f-f_L}(\phi)=|F_{f-f_L}(\phi)|. For any \lambda\geqslant 0, from the definition of \rho(u) we obtain
\begin{equation} \|f-f_L-\lambda \nu\phi\|+\|f-f_L+\lambda \nu\phi\| \leqslant 2\|f-f_L\|\biggl(1+\rho\biggl(\frac{\lambda}{\|f-f_L\|}\biggr)\biggr). \end{equation} \tag{2.5}
Next,
\begin{equation} \|f-f_L+\lambda \nu\phi\| \geqslant |F_{f-f_L}(f-f_L+\lambda\nu\phi)|= \|f-f_L\|+\lambda\beta. \end{equation} \tag{2.6}
Combining (2.5) and (2.6) we obtain
\begin{equation} \|f-f_L-\lambda\nu\phi\|\leqslant \|f-f_L\|\biggl(1-\frac{\lambda\beta}{\|f-f_L\|}+ 2\rho\biggl(\frac{\lambda}{\|f-f_L\|}\biggr)\!\biggr). \end{equation} \tag{2.7}
Taking into account that \rho(u)=o(u), we find \lambda'>0 such that
\begin{equation*} \biggl(1-\frac{\lambda'\beta}{\|f-f_L\|}+ 2\rho\biggl(\frac{\lambda'}{\|f-f_L\|}\biggr)\!\biggr)<1. \end{equation*} \notag
Then (2.7) gives
\begin{equation*} \|f-f_L-\lambda'\nu\phi\| < \|f-f_L\|, \end{equation*} \notag
which contradicts the assumption that f_L \in L is the best approximant of f. \Box

Remark 2.1. The condition F_{f-f_L}(\phi)=0 for all \phi \in L is also a sufficient condition for f_L\in L to be a best approximant of f in L.

Indeed, for any g \in L we have

\begin{equation*} \|f-f_L\|=F_{f-f_L}(f-f_L)=F_{f-f_L}(f-g) \leqslant \|f-g\|. \end{equation*} \notag

Lemma 2.3. For any bounded linear functional F and any dictionary {\mathcal D} we have

\begin{equation*} \|F\|_{\mathcal D}:=\sup_{g\in {\mathcal D}}|F(g)|= \sup_{f\in A_1({\mathcal D})}|F(f)|. \end{equation*} \notag

Proof. The inequality
\begin{equation*} \sup_{g\in {\mathcal D}}|F(g)| \leqslant \sup_{f\in A_1({\mathcal D})} |F(f)| \end{equation*} \notag
is obvious because {\mathcal D} \subset A_1({\mathcal D}). We prove the reverse inequality. Take any f\in A_1({\mathcal D}). Then for any \varepsilon >0 there exist g_1^\varepsilon,\dots,g_N^\varepsilon \in {\mathcal D} and numbers a_1^\varepsilon,\dots,a_N^\varepsilon such that |a_1^\varepsilon|+\cdots+|a_N^\varepsilon| \leqslant 1 and
\begin{equation*} \biggl\|f-\sum_{i=1}^Na_i^\varepsilon g_i^\varepsilon\biggr\| \leqslant \varepsilon. \end{equation*} \notag
Thus,
\begin{equation*} |F(f)| \leqslant \|F\|\varepsilon+ \biggl|F\biggl(\,\sum_{i=1}^Na_i^\varepsilon g_i^\varepsilon\biggr)\biggr| \leqslant \varepsilon \|F\|+\sup_{g\in {\mathcal D}}|F(g)|, \end{equation*} \notag
which proves the lemma. \Box

Lemma 2.4. For any bounded linear functional F and any dictionary {\mathcal D} we have

\begin{equation*} \sup_{g\in {\mathcal D}}\operatorname{Re}(F(g))= \sup_{f\in \operatorname{conv}({\mathcal D})} \operatorname{Re}(F(f)). \end{equation*} \notag

Proof. The inequality
\begin{equation*} \sup_{g\in {\mathcal D}}\operatorname{Re}(F(g)) \leqslant \sup_{f\in \operatorname{conv}({\mathcal D})} \operatorname{Re}(F(f)) \end{equation*} \notag
is obvious because {\mathcal D} \subset \operatorname{conv}({\mathcal D}). We prove the reverse inequality. Take any f\in \operatorname{conv}({\mathcal D}). Then for any \varepsilon >0 there exist g_1^\varepsilon,\dots,g_N^\varepsilon \in {\mathcal D} and non-negative numbers a_1^\varepsilon,\dots,a_N^\varepsilon such that a_1^\varepsilon+\cdots+a_N^\varepsilon=1 and
\begin{equation*} \biggl\|f-\sum_{i=1}^Na_i^\varepsilon g_i^\varepsilon\biggr\| \leqslant \varepsilon. \end{equation*} \notag
Thus,
\begin{equation*} \operatorname{Re}(F(f)) \leqslant \|F\|\varepsilon+\operatorname{Re} \biggl(F\biggl(\,\sum_{i=1}^Na_i^\varepsilon g_i^\varepsilon\biggr)\!\biggr) \leqslant \varepsilon \|F\|+\sup_{g\in {\mathcal D}} \operatorname{Re}(F(g)), \end{equation*} \notag
which proves the lemma. \Box

3. Main results

In this section we discuss three greedy-type algorithms and prove convergence and rate of convergence results for those algorithms. In the case of real Banach spaces these algorithms and the corresponding results on their convergence and rate of convergence are known. We show here how to modify these algorithms to the case of complex Banach spaces.

3.1. WGAFR

The following algorithm — WGAFR — can be defined in the same way in both the real and complex cases. We present the definition in the complex case.

The Weak Greedy Algorithm with Free Relaxation (WGAFR). Let \tau:=\{t_m\}_{m=1}^\infty, t_m\in[0,1], be a weakness sequence. We define f_0:=f and G_0:=0. Then for each m\geqslant 1 we have the following inductive definition.

(1) \varphi_m \in {\mathcal D} is any element satisfying

\begin{equation*} |F_{f_{m-1}}(\varphi_m)| \geqslant t_m\|F_{f_{m-1}}\|_{\mathcal D}. \end{equation*} \notag

(2) Find w_m\in {\mathbb C} and \lambda_m\in{\mathbb C} such that

\begin{equation*} \|f-((1-w_m)G_{m-1}+\lambda_m\varphi_m)\|= \inf_{\lambda,w\in{\mathbb C}}\|f-((1-w)G_{m-1}+\lambda\varphi_m)\| \end{equation*} \notag
and define
\begin{equation*} G_m:=(1-w_m)G_{m-1}+\lambda_m\varphi_m. \end{equation*} \notag

(3) Let

\begin{equation*} f_m:=f-G_m. \end{equation*} \notag

We begin with the main lemma. In the case of real Banach spaces this lemma is known (see, for instance, [22], p. 350, Lemma 6.20).

Lemma 3.1. Let X be a uniformly smooth complex Banach space with modulus of smoothness \rho(u). Take a number \varepsilon\geqslant 0 and two elements f and f^\varepsilon of X such that

\begin{equation*} \|f-f^\varepsilon\| \leqslant \varepsilon,\qquad \frac{f^\varepsilon}{A(\varepsilon)} \in A_1({\mathcal D}) \end{equation*} \notag
for some number A(\varepsilon)\geqslant \varepsilon. Then for the residual of the WGAFR after m iterations we have
\begin{equation*} \|f_m\| \leqslant \|f_{m-1}\|\inf_{\lambda\geqslant0} \biggl(1-\frac{\lambda t_m}{A(\varepsilon)} \biggl(1-\frac{\varepsilon}{\|f_{m-1}\|}\biggr)+ 2\rho\biggl(\frac{5\lambda}{\|f_{m-1}\|}\biggr)\!\biggr), \end{equation*} \notag
for m=1,2,\dots .

Proof. Denote by \nu the complex conjugate of \operatorname{sign} F_{f_{m-1}}(\varphi_m), where \operatorname{sign} z:=z/|z| for z\ne 0. Then \nu F_{f_{m-1}}(\varphi_m)=|F_{f_{m-1}}(\varphi_m)|. By the definition of f_m
\begin{equation*} \|f_m\|\leqslant \inf_{\lambda\geqslant0,w\in{\mathbb C}}\|f_{m-1}+ wG_{m-1}-\lambda\nu\varphi_m\|. \end{equation*} \notag
As in the proof of Lemma 2.2 we use the inequality
\begin{equation} \begin{aligned} \, \nonumber &\|f_{m-1}+wG_{m-1}-\lambda\nu\varphi_m\|+\|f_{m-1}-wG_{m-1}+ \lambda\nu\varphi_m\| \\ &\qquad \leqslant 2\|f_{m-1}\|\biggl(1+\rho\biggl(\frac{\|wG_{m-1}- \lambda\nu\varphi_m\|}{\|f_{m-1}\|}\biggr)\!\biggr) \end{aligned} \end{equation} \tag{3.1}
and estimate for \lambda\geqslant0:
\begin{equation*} \begin{aligned} \, &\|f_{m-1}-wG_{m-1}+\lambda\nu\varphi_m\|\geqslant \operatorname{Re}(F_{f_{m-1}}(f_{m-1}-wG_{m-1}+\lambda\nu\varphi_m)) \\ &\qquad\geqslant \|f_{m-1}\|-\operatorname{Re}(F_{f_{m-1}}(wG_{m-1}))+ \lambda t_m\sup_{g\in{\mathcal D}}|F_{f_{m-1}}(g)| \\ &\qquad=\|f_{m-1}\|-\operatorname{Re}(F_{f_{m-1}}(wG_{m-1}))+ \lambda t_m\sup_{\phi\in A_1({\mathcal D})}|F_{f_{m-1}}(\phi)|\quad (\text{by Lemma 2.3}) \\ &\qquad\geqslant \|f_{m-1}\|-\operatorname{Re}(F_{f_{m-1}}(wG_{m-1}))+ \frac{\lambda t_m}{A(\varepsilon)}|F_{f_{m-1}}(f^\varepsilon)| \\ &\qquad\geqslant\|f_{m-1}\|-\operatorname{Re}(F_{f_{m-1}}(wG_{m-1}))+ \frac{\lambda t_m}{A(\varepsilon)}(\operatorname{Re}(F_{f_{m-1}}(f))- \varepsilon). \end{aligned} \end{equation*} \notag
We set w^*:=\lambda t_m/A(\varepsilon) and obtain
\begin{equation} \|f_{m-1}-w^*G_{m-1}+\lambda\nu\varphi_m\|\geqslant\|f_{m-1}\|+ \frac{\lambda t_m}{A(\varepsilon)}(\|f_{m-1}\|-\varepsilon). \end{equation} \tag{3.2}
Combining (3.1) and (3.2) we get that
\begin{equation*} \|f_m\| \leqslant \|f_{m-1}\|\inf_{\lambda\geqslant0} \biggl(1-\frac{\lambda t_m}{A(\varepsilon)} \biggl(1-\frac{\varepsilon}{\|f_{m-1}\|}\biggr) +2\rho\biggl(\frac{\|w^*G_{m-1}-\lambda\nu\varphi_m\|} {\|f_{m-1}\|}\biggr)\!\biggr). \end{equation*} \notag
We now estimate
\begin{equation*} \|w^*G_{m-1}-\lambda\nu\varphi_m\| \leqslant w^*\|G_{m-1}\|+\lambda. \end{equation*} \notag
Next,
\begin{equation*} \|G_{m-1}\|=\|f-f_{m-1}\|\leqslant 2\|f\|\leqslant 2(\|f^\varepsilon\|+\varepsilon)\leqslant2(A(\varepsilon)+\varepsilon). \end{equation*} \notag
Thus, under the assumption A(\varepsilon)\geqslant\varepsilon we obtain
\begin{equation*} w^*\|G_{m-1}\|\leqslant \frac{2\lambda t_m(A(\varepsilon)+\varepsilon)} {A(\varepsilon)} \leqslant 4\lambda. \end{equation*} \notag
Finally,
\begin{equation*} \|w^*G_{m-1}-\lambda\varphi_m\|\leqslant 5\lambda. \end{equation*} \notag
This completes the proof of the lemma. \Box

Remark 3.1. It follows from the definition of the \operatorname{WGAFR} that the sequence \{\|f_m\|\} is non-increasing.

We now prove a convergence theorem for an arbitrary uniformly smooth Banach space. The modulus of smoothness \rho(u) of a uniformly smooth Banach space is an even convex function such that \rho(0)=0 and \lim_{u\to0}\rho(u)/u=0. The function s(u):=\rho(u)/u, s(0):=0, associated with \rho(u) is a continuous increasing function on [0,\infty). Therefore, the inverse function s^{-1}(\,\cdot\,) is well defined. The following theorem is known in the case of real Banach spaces (see, for instance, [22], p. 352, Theorem 6.22).

Theorem 3.1. Let X be a uniformly smooth complex Banach space with modulus of smoothness \rho(u). Assume that a sequence \tau:=\{t_k\}_{k=1}^\infty satisfies the following condition: for each \theta >0 we have

\begin{equation} \sum_{m=1}^\infty t_m s^{-1}(\theta t_m)=\infty. \end{equation} \tag{3.3}
Then for any f\in X we have for the WGAFR
\begin{equation*} \lim_{m\to \infty} \|f_m\|=0. \end{equation*} \notag

Proof. This proof only uses Lemma 3.1, which provides a relation between the residuals \|f_m\| and \|f_{m-1}\|. It repeats the corresponding proof in the real case. For completeness we present this simple proof here. By Remark 3.1 \{\|f_m\|\} is a non-increasing sequence. Therefore,
\begin{equation*} \lim_{m\to \infty}\|f_m\|=\beta. \end{equation*} \notag
We prove that \beta=0 by contradiction. Assume the contrary: \beta>0. Then for any m we have
\begin{equation*} \|f_m\| \geqslant \beta. \end{equation*} \notag
We set \varepsilon=\beta/2 and find f^\varepsilon such that
\begin{equation*} \|f-f^\varepsilon\| \leqslant \varepsilon \quad \text{and}\quad \frac{f^\varepsilon}{A(\varepsilon)} \in A_1({\mathcal D}), \end{equation*} \notag
for some A(\varepsilon)\geqslant\varepsilon. Then by Lemma 3.1 we obtain
\begin{equation*} \|f_m\| \leqslant \|f_{m-1}\|\inf_{\lambda\geqslant0} \biggl(1-\frac{\lambda t_m}{2A(\varepsilon)}+ 2\rho\biggl(\frac{5\lambda}{\beta}\biggr)\!\biggr). \end{equation*} \notag
Let us specify \theta:=\beta/(40A(\varepsilon)) and take \lambda=\beta s^{-1}(\theta t_m)/5. Then
\begin{equation*} \|f_m\| \leqslant \|f_{m-1}\|\bigl(1-2\theta t_m s^{-1}(\theta t_m)\bigr). \end{equation*} \notag
The assumption
\begin{equation*} \sum_{m=1}^\infty t_m s^{-1}(\theta t_m)=\infty \end{equation*} \notag
implies that
\begin{equation*} \|f_m\| \to 0 \quad \text{as} \ \ m\to \infty. \end{equation*} \notag
We obtain a contradiction, which proves the theorem. \Box

Remark 3.2. The reader can find some necessary conditions on the weakness sequence \tau for the convergence of the WCGA, for instance, in [22], p. 346, Proposition 6.13.

We now proceed to results on the rate of convergence. The following theorem is known in the case of real Banach spaces (see, for instance, [22], p. 353, Theorem 6.23).

Theorem 3.2. Let X be a uniformly smooth Banach space with modulus of smoothness \rho(u)\leqslant \gamma u^q, 1<q\leqslant 2. Take a number \varepsilon\geqslant 0 and two elements f and f^\varepsilon of X such that

\begin{equation*} \|f-f^\varepsilon\| \leqslant \varepsilon,\qquad \frac{f^\varepsilon}{A(\varepsilon)} \in A_1({\mathcal D}), \end{equation*} \notag
for some number A(\varepsilon)>0. Then for the residual of the WGAFR after m iterations we have
\begin{equation*} \|f_m\| \leqslant \max\biggl(2\varepsilon, C(q,\gamma)(A(\varepsilon)+ \varepsilon)\biggl(1+\sum_{k=1}^mt_k^p\biggr)^{\!-1/p}\,\biggr),\qquad p:=\frac{q}{q-1}\,. \end{equation*} \notag

Proof. This proof only uses Lemma 3.1, which provides a relation between the residuals \|f_m\| and \|f_{m-1}\|. It repeats the corresponding proof in the real case. For completeness we present this proof here. It is clear that it suffices to consider the case A(\varepsilon)\geqslant\varepsilon. Otherwise, \|f_m\|\leqslant\|f\|\leqslant\|f^\varepsilon\|+ \varepsilon\leqslant2\varepsilon. Also assume that \|f_m\|>2\varepsilon (otherwise Theorem 3.2 holds trivially). Then, by Remark 3.1 we have \|f_k\|>2\varepsilon for all k=0,1,\dots,m. By Lemma 3.1 we obtain
\begin{equation} \|f_k\| \leqslant \|f_{k-1}\|\inf_{\lambda\geqslant0} \biggl(1-\frac{\lambda t_k}{2A(\varepsilon)}+ 2\gamma\biggl(\frac{5\lambda}{\|f_{k-1}\|}\biggr)^{\!q}\,\biggr). \end{equation} \tag{3.4}
We choose \lambda from the equation
\begin{equation*} \frac{\lambda t_k}{4A(\varepsilon)}= 2\gamma \biggl(\frac{5\lambda}{\|f_{k-1}\|}\biggr)^{\!q}, \end{equation*} \notag
which implies that
\begin{equation*} \lambda=\|f_{k-1}\|^{q/(q-1)}\,5^{-q/(q-1)} (8\gamma A(\varepsilon))^{-1/(q-1)}t_k^{1/(q-1)}. \end{equation*} \notag
Define
\begin{equation*} A_q:=4(8\gamma)^{1/(q-1)}\,5^{q/(q-1)}. \end{equation*} \notag
Using notation p:=q/(q-1), we deduce from (3.4)
\begin{equation*} \|f_k\| \leqslant \|f_{k-1}\|\biggl(1-\frac{1}{4}\, \frac{\lambda t_k}{A(\varepsilon)}\biggr)= \|f_{k-1}\|\biggl(1-\frac{t_k^p\|f_{k-1}\|^p}{A_qA(\varepsilon)^p}\biggr). \end{equation*} \notag
Raising both sides of this inequality to power p and taking into account the inequality x^r\leqslant x for r\geqslant 1, 0\leqslant x\leqslant 1, we obtain
\begin{equation} \|f_k\|^p \leqslant \|f_{k-1}\|^p \biggl(1-\frac{t^p_k\|f_{k-1}\|^p}{A_qA(\varepsilon)^p}\biggr). \end{equation} \tag{3.5}
We will need the following simple lemma, various versions of which are well known (see, for example, [22], p. 91).

Lemma 3.2. Given a number C_1>0 and a sequence \{a_k\}_{k=1}^\infty, a_k >0, k=1,2,\dots, assume that \{x_m\}_{m=0}^\infty is a sequence of non-negative numbers satisfying the inequalities

\begin{equation*} x_0 \leqslant C_1, \qquad x_{m+1} \leqslant x_m(1-x_m a_{m+1}), \quad m=1,2,\dots\,. \end{equation*} \notag
Then for each m we have
\begin{equation*} x_m \leqslant \biggl(\frac{1}{C_1}+\sum_{k=1}^{m} a_k\biggr)^{-1}. \end{equation*} \notag

Proof. The proof is by induction on m. For m=0 the statement is true by assumption. We prove that the inequality
\begin{equation*} x_m \leqslant \biggl(\frac{1}{C_1}+\sum_{k=1}^{m} a_k\biggr)^{\!-1} \quad\text{implies that} \quad x_{m+1} \leqslant\biggl(\frac{1}{C_1}+\sum_{k=1}^{m+1} a_k\biggr)^{\!-1}. \end{equation*} \notag
If x_{m+1}=0, this statement is obvious. Assume therefore that x_{m+1} > 0. Then we have
\begin{equation*} x_{m+1}^{-1} \geqslant x_m^{-1}(1-x_m a_{m+1})^{-1} \geqslant x_m^{-1}(1+x_m a_{m+1})=x_m^{-1}+a_{m+1} \geqslant \frac{1}{C_1}+\sum_{k=1}^{m+1} a_k, \end{equation*} \notag
which proves the required inequality. \Box

We use Lemma 3.2 for x_k:=\|f_k\|^p. Using the estimate \|f\| \leqslant A(\varepsilon)+\varepsilon, we set C_1:= A(\varepsilon)+\varepsilon. We specify a_k:=t^p_k/(A_qA(\varepsilon)^p). Then (3.5) guarantees that we can apply Lemma 3.2. Note that A_q>1. Then Lemma 3.2 gives

\begin{equation} \|f_m\|^p \leqslant A_q(A(\varepsilon)+ \varepsilon)^p\biggl(1+\sum_{k=1}^m t_k^p\biggr)^{\!-1}, \end{equation} \tag{3.6}
which implies that
\begin{equation*} \|f_m\|\leqslant C(q,\gamma)(A(\varepsilon)+ \varepsilon)\biggl(1+\sum_{k=1}^m t_k^p\biggr)^{\!-1/p}. \end{equation*} \notag
Theorem 3.2 is proved. \Box

3.2. GAWR

In this subsection we study the Greedy Algorithm with Weakness parameter t and Relaxation {\mathbf r} (\operatorname{GAWR}(t,{\mathbf r})), where {\mathbf r}:=\{r_j\}_{j=1}^\infty, r_j\in [0,1), j=1,2,\dots . In addition to the acronym \operatorname{GAWR}(t,{\mathbf r}), we will use the abbreviated acronym \operatorname{GAWR} for the name of this algorithm. We give a general definition of the algorithm in the case of the weakness sequence \tau.

The Greedy Algorithm with Weakness and Relaxation \operatorname{GAWR}(\tau,\mathbf r). Let \tau:=\{t_m\}_{m=1}^\infty, t_m\in[0,1], be a weakness sequence. We define f_0:=f and G_0:=0. Then for each m\geqslant 1 we give the following inductive definition.

(1) \varphi_m \in {\mathcal D} is any element satisfying

\begin{equation*} |F_{f_{m-1}}(\varphi_m)| \geqslant t_m \| F_{f_{m-1}}\|_{\mathcal D}. \end{equation*} \notag

(2) Find \lambda_m \in {\mathbb C} such that

\begin{equation*} \|f-((1-r_m)G_{m-1}+\lambda_m\varphi_m)\|= \inf_{\lambda\in {\mathbb C}}\|f-((1-r_m)G_{m-1}+\lambda\varphi_m)\| \end{equation*} \notag
and define
\begin{equation*} G_m:=(1-r_m)G_{m-1}+\lambda_m\varphi_m. \end{equation*} \notag

(3) Set

\begin{equation*} f_m:=f-G_m. \end{equation*} \notag

We now prove results on convergence and rate of convergence for the \operatorname{GAWR}(t,{\mathbf r}), that is, for the \operatorname{GAWR}(\tau,{\mathbf r}) with \tau =\{t_j\}_{j=1}^\infty, t_j=t \in (0,1], j=1,2,\dots . We begin with an analogue of Lemma 3.1. In the case of real Banach spaces Lemma 3.3 is known (see, for instance, [22], p. 355, Lemma 6.24).

Lemma 3.3. Let X be a uniformly smooth complex Banach space with modulus of smoothness \rho(u). Take a number \varepsilon\geqslant 0 and two elements f and f^\varepsilon of X such that

\begin{equation*} \|f-f^\varepsilon\| \leqslant \varepsilon,\qquad \frac{f^\varepsilon}{A(\varepsilon)} \in A_1({\mathcal D}), \end{equation*} \notag
for some number A(\varepsilon)>0. Then for the \operatorname{GAWR}(t,\mathbf r) we have the inequality
\begin{equation*} \|f_m\| \leqslant \|f_{m-1}\|\biggl(1-r_m\biggl(1-\frac{\varepsilon} {\|f_{m-1}\|}\biggr)+2\rho\biggl(\frac{r_m(\|f\|+A(\varepsilon)/t)}{(1-r_m) \|f_{m-1}\|}\biggr)\!\biggr), \end{equation*} \notag
for m=1,2,\dots .

Proof. As in the above proof of Lemma 3.1, we denote by \nu the complex conjugate of \operatorname{sign} F_{f_{m-1}}(\varphi_m), where \operatorname{sign}z:=z/|z| for z\ne 0. Then \nu F_{f_{m-1}}(\varphi_m)=|F_{f_{m-1}}(\varphi_m)|. The definition of f_m implies that
\begin{equation*} \|f_m\|\leqslant \inf_{\lambda\geqslant0}\|f-((1-r_m)G_{m-1}+ \lambda\nu\varphi_m)\|. \end{equation*} \notag
For any \lambda we have
\begin{equation} f-((1-r_m)G_{m-1}+\lambda\nu \varphi_m)= (1-r_m)f_{m-1}+r_mf-\lambda\nu \varphi_m, \end{equation} \tag{3.7}
and from the definition of \rho(u) with u=\|r_mf-\lambda\nu\varphi_m\|/((1-r_m)\|f_{m-1}\|) we obtain
\begin{equation} \begin{aligned} \, \nonumber &\|(1-r_m)f_{m-1}+r_mf-\lambda\nu \varphi_m\|+\|(1-r_m)f_{m-1}-r_mf+ \lambda\nu \varphi_m\| \\ &\qquad\leqslant 2(1-r_m)\|f_{m-1}\| \biggl(1+\rho\biggl(\frac{\|r_mf-\lambda\nu\varphi_m\|} {(1-r_m)\|f_{m-1}\|}\biggr)\!\biggr). \end{aligned} \end{equation} \tag{3.8}
We have
\begin{equation} \begin{aligned} \, \nonumber &\|(1-r_m)f_{m-1}-r_mf+\lambda \nu\varphi_m\|\geqslant \operatorname{Re}\bigl(F_{f_{m-1}}((1-r_m)f_{m-1}-r_mf+ \lambda\nu \varphi_m)\bigr) \\ &\qquad=(1-r_m)\|f_{m-1}\|-r_m\operatorname{Re}(F_{f_{m-1}}(f))+ \lambda \operatorname{Re}(\nu F_{f_{m-1}}(\varphi_m)). \end{aligned} \end{equation} \tag{3.9}
From the definition of \varphi_m we obtain
\begin{equation} \nu F_{f_{m-1}}(\varphi_m)=|F_{f_{m-1}}(\varphi_m)| \geqslant t\sup_{g\in{\mathcal D}}|F_{f_{m-1}}(g)|. \end{equation} \tag{3.10}
From Lemma 2.3 we deduce that
\begin{equation} \begin{aligned} \, \nonumber \sup_{g\in{\mathcal D}} |F_{f_{m-1}}(g)|&= \sup_{\phi\in A_1({\mathcal D})}|F_{f_{m-1}}(\phi)| \\ &\geqslant \frac{1}{A(\varepsilon)}|F_{f_{m-1}}(f^\varepsilon)|\geqslant \frac{1}{A(\varepsilon)}(|F_{f_{m-1}}(f)|-\varepsilon). \end{aligned} \end{equation} \tag{3.11}
Combining (3.10) and (3.11) we obtain
\begin{equation} |F_{f_{m-1}}(\varphi_m)|\geqslant \frac{t}{A(\varepsilon)}(|F_{f_{m-1}}(f)|-\varepsilon). \end{equation} \tag{3.12}
We now choose \lambda:=\lambda^*:=r_mA(\varepsilon)/t. Then for this \lambda we derive from (3.9) and (3.12) the inequality
\begin{equation} \|(1-r_m)f_{m-1}-r_mf+\lambda^*\nu \varphi_m\|\geqslant (1-r_m)\|f_{m-1}\|-r_m\varepsilon. \end{equation} \tag{3.13}
Relations (3.8) and (3.13) imply that
\begin{equation*} \begin{aligned} \, &\|(1-r_m)f_{m-1}+r_mf-\lambda^*\nu \varphi_m\| \\ &\qquad\leqslant (1-r_m)\|f_{m-1}\|+r_m\varepsilon+ 2(1-r_m)\|f_{m-1}\|\,\rho\biggl(\frac{r_m(\|f\|+A(\varepsilon)/t)} {(1-r_m)\|f_{m-1}\|}\biggr). \end{aligned} \end{equation*} \notag
The lemma is proved.

We begin with a convergence result. It is known that the version of Lemma 3.3 for the real case implies the corresponding convergence result — Theorem 3.3 below — in the real case (see, for instance, [22], pp. 355–357). That same proof derives Theorem 3.3 from Lemma 3.3. We do not present it here.

Theorem 3.3. Let the sequence {\mathbf r}:=\{r_k\}_{k=1}^\infty, r_k\in[0,1), satisfy the conditions

\begin{equation*} \sum_{k=1}^\infty r_k =\infty,\qquad r_k\to 0\quad\textit{as}\ \ k\to\infty. \end{equation*} \notag
Then the \operatorname{GAWR}(t,{\mathbf r}) converges in any uniformly smooth Banach space for each f\in X and for all dictionaries {\mathcal D}.

The situation with results on the rate of convergence is very similar to the situation with the convergence result described above. The real-case proof (see, for instance, [22], pp. 357–358) derives Theorem 3.4 from Lemma 3.3. We do not present it here.

Theorem 3.4. Let X be a uniformly smooth Banach space with modulus of smoothness \rho(u)\leqslant \gamma u^q, 1<q\leqslant 2. Let {\mathbf r}:=\{2/(k+2)\}_{k=1}^\infty. Consider the \operatorname{GAWR}(t,{\mathbf r}). Then for a pair of functions f, f^\varepsilon satisfying

\begin{equation*} \|f-f^\varepsilon\|\leqslant \varepsilon,\qquad \frac{f^\varepsilon}{A(\varepsilon)} \in A_1({\mathcal D}) \end{equation*} \notag
we have
\begin{equation*} \|f_m\|\leqslant \varepsilon+C(q,\gamma)\biggl(\|f\|+ \frac{A(\varepsilon)}{t}\biggr)m^{-1+1/q}. \end{equation*} \notag

3.3. Incremental algorithm

We proceed to one more greedy-type algorithm (see [21]). Let \varepsilon=\{\varepsilon_n\}_{n=1}^\infty , \varepsilon_n> 0, n=1,2,\dots . The following greedy algorithm was introduced and studied in [21] (see also [22], pp. 361–363) in the case of real Banach spaces.

The Incremental Algorithm with schedule \varepsilon (\operatorname{IA}(\varepsilon)). Let f\in\operatorname{conv}({\mathcal D}). Denote f_0^{i,\varepsilon}:=f and G_0^{i,\varepsilon}:=0. Then for each m\geqslant 1 we have the following inductive definition.

(1) \varphi_m^{i,\varepsilon} \in {\mathcal D} is any element satisfying

\begin{equation*} F_{f_{m-1}^{i,\varepsilon}}(\varphi_m^{i,\varepsilon}-f) \geqslant -\varepsilon_m. \end{equation*} \notag

(2) Define

\begin{equation*} G_m^{i,\varepsilon}:=\biggl(1-\frac{1}{m}\biggr)G_{m-1}^{i,\varepsilon}+ \frac{\varphi_m^{i,\varepsilon}}{m}\,. \end{equation*} \notag

(3) Let

\begin{equation*} f_m^{i,\varepsilon}:=f-G_m^{i,\varepsilon}. \end{equation*} \notag
It is clear from the definition of the \operatorname{IA}(\varepsilon) that
\begin{equation} G_m^{i,\varepsilon}=\frac{1}{m}\sum_{j=1}^m \varphi_j^{i,\varepsilon}. \end{equation} \tag{3.14}

We now modify the definition of \operatorname{IA}(\varepsilon) to make it suitable for complex Banach spaces.

Incremental Algorithm (complex) with schedule \varepsilon (\operatorname{IAc}(\varepsilon)). Let f\in A_1({\mathcal D}). Denote f_0^{{\rm c},\varepsilon}:= f and G_0^{{\rm c},\varepsilon}:=0. Then for each m\geqslant 1 we have the following inductive definition.

(1) \varphi_m^{{\rm c},\varepsilon} \in {\mathcal D}^\circ is any element satisfying

\begin{equation*} \operatorname{Re}(F_{f_{m-1}^{{\rm c},\varepsilon}} (\varphi_m^{{\rm c},\varepsilon}-f))\geqslant -\varepsilon_m. \end{equation*} \notag
Denote by \nu_m the complex conjugate of \operatorname{sign} F_{f_{m-1}^{{\rm c},\varepsilon}}(\varphi_m^{{\rm c},\varepsilon}), where \operatorname{sign} z:=z/|z| for z\ne 0.

(2) Define

\begin{equation*} G_m^{{\rm c},\varepsilon}:=\biggl(1-\frac{1}{m}\biggr) G_{m-1}^{{\rm c},\varepsilon}+ \frac{\nu_m\varphi_m^{{\rm c},\varepsilon}}{m}\,. \end{equation*} \notag

(3) Let

\begin{equation*} f_m^{{\rm c},\varepsilon}:=f-G_m^{{\rm c},\varepsilon}. \end{equation*} \notag

It follows from the definition of the \operatorname{IAc}(\varepsilon) that

\begin{equation} G_m^{{\rm c},\varepsilon}= \frac{1}{m}\sum_{j=1}^m \nu_j\varphi_j^{{\rm c},\varepsilon},\qquad |\nu_j|=1, \quad j=1,2,\dots\,. \end{equation} \tag{3.15}

Note that by the definition of \nu_m we have

\begin{equation*} F_{f_{m-1}^{{\rm c},\varepsilon}}(\nu_m\varphi_m^{{\rm c},\varepsilon})= |F_{f_{m-1}^{{\rm c},\varepsilon}}(\varphi_m^{{\rm c},\varepsilon})|, \end{equation*} \notag
and therefore, by Lemma 2.3
\begin{equation*} \begin{aligned} \, \sup_{\phi \in {\mathcal D}^\circ}\operatorname{Re} \bigl(F_{f_{m-1}^{{\rm c},\varepsilon}}(\phi)\bigr)&= \sup_{g \in {\mathcal D}}\bigl|F_{f_{m-1}^{{\rm c},\varepsilon}}(g)\bigr|= \sup_{\phi \in A_1({\mathcal D})} \bigl|F_{f_{m-1}^{{\rm c},\varepsilon}}(\phi)\bigr| \\ &\geqslant \bigl|F_{f_{m-1}^{{\rm c},\varepsilon}}(f)\bigr| \geqslant \operatorname{Re}\bigl(F_{f_{m-1}^{{\rm c},\varepsilon}}(f)\bigr). \end{aligned} \end{equation*} \notag
This means that we can always run the \operatorname{IAc}(\varepsilon) for f\in A_1({\mathcal D}).

Theorem 3.5. Let X be a uniformly smooth complex Banach space with modulus of smoothness \rho(u)\leqslant \gamma u^q, 1<q\leqslant 2. Define

\begin{equation*} \varepsilon_n:=K_1\gamma ^{1/q}n^{-1/p},\qquad p=\frac{q}{q-1}\,,\quad n=1,2,\dots\,. \end{equation*} \notag
Then for any f\in A_1({\mathcal D}) we have
\begin{equation*} \|f_m^{{\rm c},\varepsilon}\| \leqslant C(K_1) \gamma^{1/q}m^{-1/p},\qquad m=1,2,\dots\,. \end{equation*} \notag

Proof. We use the abbreviated notation f_m:=f_m^{{\rm c},\varepsilon}, \varphi_m:=\nu_m\varphi_m^{{\rm c},\varepsilon}, and G_m:=G_m^{{\rm c},\varepsilon}. Writing
\begin{equation*} f_m=f_{m-1}-\frac{\varphi_m-G_{m-1}}{m}\,, \end{equation*} \notag
we immediately obtain the trivial estimate
\begin{equation} \|f_m\|\leqslant \|f_{m-1}\|+\frac{2}{m}\,. \end{equation} \tag{3.16}
Since
\begin{equation} f_m=\biggl(1-\frac{1}{m}\biggr)f_{m-1}-\frac{\varphi_m-f}{m} =\biggl(1-\frac{1}{m}\biggr)\biggl(f_{m-1}-\frac{\varphi_m-f}{m-1}\biggr), \end{equation} \tag{3.17}
from Lemma 2.1 for y=\varphi_m-f and u=-1/(m-1) we obtain the inequality
\begin{equation} \biggl\|f_{m-1}-\frac{\varphi_m-f}{m-1}\biggr\|\leqslant \|f_{m-1}\|\biggl(1+2\rho\biggl(\frac{2}{(m-1)\|f_{m-1}\|}\biggr)\!\biggr)+ \frac{\varepsilon_m}{m-1}\,. \end{equation} \tag{3.18}
Using the definition of \varepsilon_m and the assumption \rho(u) \leqslant \gamma u^q, we make the following observation. There exists a constant C(K_1) such that if
\begin{equation} \|f_{m-1}\|\geqslant C(K_1)\gamma^{1/q}(m-1)^{-1/p}, \end{equation} \tag{3.19}
then
\begin{equation} 2\rho\biggl(\frac{2}{(m-1)\|f_{m-1}\|}\biggr)+ \frac{\varepsilon_m}{(m-1)\|f_{m-1}\|}\leqslant \frac{1}{4m}\,, \end{equation} \tag{3.20}
and therefore, by (3.17) and (3.18)
\begin{equation} \|f_m\| \leqslant \biggl(1-\frac{3}{4m}\biggr)\|f_{m-1}\|. \end{equation} \tag{3.21}

We now need the following technical lemma (see, for instance, [22], p. 357).

Lemma 3.4. Let the sequence \{a_n\}_{n=1}^\infty have the following property: given positive numbers \alpha < \gamma \leqslant 1 and A > a_1, we have

\begin{equation} a_n\leqslant a_{n-1}+A(n-1)^{-\alpha}\quad\textit{for all}\ \ n\geqslant2, \end{equation} \tag{3.22}
and if for some v \geqslant2 we have
\begin{equation*} a_v \geqslant Av^{-\alpha}, \end{equation*} \notag
then
\begin{equation} a_{v+1} \leqslant a_v\biggl(1-\frac{\gamma}{v}\biggr). \end{equation} \tag{3.23}

Then there exists a constant C(\alpha,\gamma) such that

\begin{equation*} a_n \leqslant C(\alpha,\gamma)An^{-\alpha}\quad\textit{for all}\ \ n=1,2,\dots\,. \end{equation*} \notag

Taking (3.16) into account we apply Lemma 3.4 to the sequence a_n=\|f_n\|, n=1,2,\dots, for \alpha=1/p, \beta =3/4 and complete the proof of Theorem 3.5.

We now consider one more modification of \operatorname{IA}(\varepsilon), which is suitable for complex Banach spaces and provides a convex combination of dictionary elements. In the case of real Banach spaces it coincides with the \operatorname{IA}(\varepsilon).

The Incremental Algorithm (complex and convex) with schedule \varepsilon (\operatorname{IAcc}(\varepsilon)). Let f\in \operatorname{conv}({\mathcal D}). Set f_0^{{\rm cc},\varepsilon}:=f and G_0^{{\rm cc},\varepsilon}:=0. Then, for each m\geqslant 1 we have the following inductive definition.

(1) \varphi_m^{{\rm cc},\varepsilon} \in {\mathcal D} is any element satisfying

\begin{equation*} \operatorname{Re}\bigl(F_{f_{m-1}^{{\rm cc},\varepsilon}} (\varphi_m^{{\rm cc},\varepsilon}-f)\bigr) \geqslant -\varepsilon_m. \end{equation*} \notag

(2) Define

\begin{equation*} G_m^{{\rm cc},\varepsilon}:= \biggl(1-\frac{1}{m}\biggr)G_{m-1}^{{\rm cc},\varepsilon}+ \frac{\varphi_m^{{\rm cc},\varepsilon}}{m}\,. \end{equation*} \notag

(3) Let

\begin{equation*} f_m^{{\rm cc},\varepsilon}:=f-G_m^{{\rm cc},\varepsilon}. \end{equation*} \notag

It follows from the definition of the \operatorname{IAcc}(\varepsilon) that

\begin{equation} G_m^{{\rm cc},\varepsilon}= \frac{1}{m}\sum_{j=1}^m \varphi_j^{{\rm cc},\varepsilon}. \end{equation} \tag{3.24}

Note that by Lemma 2.4

\begin{equation*} \sup_{g \in {\mathcal D}} \operatorname{Re}\bigl(F_{f_{m-1}^{{\rm cc},\varepsilon}}(g)\bigr)= \sup_{\phi \in \operatorname{conv}({\mathcal D})}\operatorname{Re} \bigl(F_{f_{m-1}^{{\rm cc},\varepsilon}}(\phi)\bigr) \geqslant \operatorname{Re}\bigl(F_{f_{m-1}^{{\rm cc},\varepsilon}}(f)\bigr). \end{equation*} \notag
This means that we can always run the \operatorname{IAcc}(\varepsilon) for f\in \operatorname{conv}({\mathcal D}).

In the same way as Theorem 3.5 was proved above, one can prove the following theorem.

Theorem 3.6. Let X be a uniformly smooth Banach space with modulus of smoothness \rho(u)\leqslant \gamma u^q, 1<q\leqslant 2. Define

\begin{equation*} \varepsilon_n:=K_1\gamma ^{1/q}n^{-1/p},\qquad p=\frac{q}{q-1}\,,\quad n=1,2,\dots\,. \end{equation*} \notag
Then for any f\in \operatorname{conv}({\mathcal D}) we have
\begin{equation*} \|f_m^{{\rm cc},\varepsilon}\| \leqslant C(K_1)\gamma^{1/q}m^{-1/p},\qquad m=1,2,\dots\,. \end{equation*} \notag


Bibliography

1. A. R. Barron, “Universal approximation bounds for superposition of a sigmoidal functions”, IEEE Trans. Inform. Theory, 39:3 (1993), 930–945  crossref  mathscinet  zmath
2. A. R. Barron, A. Cohen, W. Dahmen, and R. DeVore, “Approximation and learning by greedy algorithms”, Ann. Statist., 36:1 (2008), 64–94  crossref  mathscinet  zmath
3. K. L. Clarkson, “Coresets, sparse greedy approximation, and the Frank–Wolfe algorithm”, ACM Trans. Algorithms, 6:4 (2010), 63, 30 pp.  crossref  mathscinet  zmath
4. G. Davis, S. Mallat, and M. Avellaneda, “Adaptive greedy approximations”, Constr. Approx., 13:1 (1997), 57–98  crossref  mathscinet  zmath
5. A. V. Dereventsov and V. N. Temlyakov, “A unified way of analyzing some greedy algorithms”, J. Funct. Anal., 277:12 (2019), 108286, 30 pp.  crossref  mathscinet  zmath
6. A. Dereventsov and V. Temlyakov, “Biorthogonal greedy algorithms in convex optimization”, Appl. Comput. Harmon. Anal., 60 (2022), 489–511  crossref  mathscinet  zmath
7. R. A. DeVore and V. N. Temlyakov, “Some remarks on greedy algorithms”, Adv. Comput. Math., 5:2-3 (1996), 173–187  crossref  mathscinet  zmath
8. R. A. DeVore and V. N. Temlyakov, “Convex optimization on Banach spaces”, Found. Comput. Math., 16:2 (2016), 369–394  crossref  mathscinet  zmath
9. S. Dilworth, G. Garrigós, E. Hernández, D. Kutzarova, and V. Temlyakov, “Lebesgue-type inequalities in greedy approximation”, J. Funct. Anal., 280:5 (2021), 108885, 37 pp.  crossref  mathscinet  zmath
10. M. J. Donahue, L. Gurvits, C. Darken, and E. Sontag, “Rate of convex approximation in non-Hilbert spaces”, Constr. Approx., 13:2 (1997), 187–220  crossref  mathscinet  zmath
11. M. A. T. Figueiredo, R. D. Nowak, and S. J. Wright, “Gradient projection for sparse reconstruction: application to compressed sensing and other inverse problems”, IEEE J. Sel. Top. Signal Process., 1:4 (2007), 586–597  crossref  adsnasa
12. J. H. Friedman and W. Stuetzle, “Projection pursuit regression”, J. Amer. Statist. Assoc., 76:376 (1981), 817–823  crossref  mathscinet
13. M. Ganichev and N. J. Kalton, “Convergence of the weak dual greedy algorithm in L_p-spaces”, J. Approx. Theory, 124:1 (2003), 89–95  crossref  mathscinet  zmath
14. Zheming Gao and G. Petrova, “Rescaled pure greedy algorithm for convex optimization”, Calcolo, 56:2 (2019), 15, 14 pp.  crossref  mathscinet  zmath
15. P. J. Huber, “Projection pursuit”, Ann. Statist., 13:2 (1985), 435–475  crossref  mathscinet  zmath
16. M. Jaggi, “Revisiting Frank–Wolfe: projection-free sparse convex optimization”, Proceedings of the 30th international conference on machine learning, Proceedings of Machine Learning Research (PMLR), 28, no. 1, 2013, 427–435, https://proceedings.mlr.press/v28/jaggi13.html
17. L. K. Jones, “On a conjecture of Huber concerning the convergence of projection pursuit regression”, Ann. Statist., 15:2 (1987), 880–882  crossref  mathscinet  zmath
18. L. K. Jones, “A simple lemma on greedy approximation in Hilbert space and convergence rates for projection pursuit regression and neural network training”, Ann. Statist., 20:1 (1992), 608–613  crossref  mathscinet  zmath
19. S. Shalev-Shwartz, N. Srebro, and Tong Zhang, “Trading accuracy for sparsity in optimization problems with sparsity constraints”, SIAM J. Optim., 20:6 (2010), 2807–2832  crossref  mathscinet  zmath
20. V. N. Temlyakov, “Greedy algorithms in Banach spaces”, Adv. Comput. Math., 14:3 (2001), 277–292  crossref  mathscinet  zmath
21. V. N. Temlyakov, “Greedy-type approximation in Banach spaces and applications”, Constr. Approx., 21:2 (2005), 257–292  crossref  mathscinet  zmath
22. V. Temlyakov, Greedy approximation, Cambridge Monogr. Appl. Comput. Math., 20, Cambridge Univ. Press, Cambridge, 2011, xiv+418 pp.  crossref  mathscinet  zmath
23. V. Temlyakov, Multivariate approximation, Cambridge Monogr. Appl. Comput. Math., 32, Cambridge Univ. Press, Cambridge, 2018, xvi+534 pp.  crossref  mathscinet  zmath
24. A. Tewari, P. K. Ravikumar, and I. S. Dhillon, “Greedy algorithms for structurally constrained high dimensional problems”, Adv. Neural Inf. Process. Syst., 24, NIPS 2011, MIT Press, Cambridge, MA, 2011, 882–890
25. Tong Zhang, “Sequential greedy approximation for certain convex optimization problems”, IEEE Trans. Inform. Theory, 49:3 (2003), 682–691  crossref  mathscinet  zmath

Citation: A. V. Gasnikov, V. N. Temlyakov, “On greedy approximation in complex Banach spaces”, Russian Math. Surveys, 79:6 (2024), 975–990
Citation in format AMSBIB
\Bibitem{GasTem24}
\by A.~V.~Gasnikov, V.~N.~Temlyakov
\paper On greedy approximation in complex Banach spaces
\jour Russian Math. Surveys
\yr 2024
\vol 79
\issue 6
\pages 975--990
\mathnet{http://mi.mathnet.ru/eng/rm10186}
\crossref{https://doi.org/10.4213/rm10186e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4867087}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024RuMaS..79..975G}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001443210000006}
Linking options:
  • https://www.mathnet.ru/eng/rm10186
  • https://doi.org/10.4213/rm10186e
  • https://www.mathnet.ru/eng/rm/v79/i6/p39
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Успехи математических наук Russian Mathematical Surveys
    Statistics & downloads:
    Abstract page:239
    Russian version PDF:18
    English version PDF:19
    Russian version HTML:26
    English version HTML:36
    References:34
    First page:22
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2025