Brought to you by:
Paper

Some trigonometric polynomials with extremely small uniform norm and their applications

© 2020 Russian Academy of Sciences (DoM) and London Mathematical Society
, , Citation A. O. Radomskii 2020 Izv. Math. 84 361 DOI 10.1070/IM8887

1064-5632/84/2/361

Abstract

We construct orthogonal trigonometric polynomials satisfying a new spectral condition and such that their $L^{1}$-norms are bounded below and the uniform norm of their partial sums has extremely small order of growth. We obtain new results that relate the uniform norm and $\mathrm{QC}$-norm on subspaces of the vector space of trigonometric polynomials.

Export citation and abstract BibTeX RIS

§ 1. Introduction

We begin with some necessary definitions. The sets of positive integers, integers and real numbers are denoted by $\mathbb{N}$, $\mathbb{Z}$ and $\mathbb{R}$, respectively. Given $f\in L^{p}(0,2\pi)$, put

Given $f\in L^{1}(0,2\pi)$, we consider the Fourier coefficients

Given a real number $x$, we denote by $[x]$ its integer part and by $\lceil x\rceil$ the least integer $n$ such that $n\geq x$. We write $|A|$ for the cardinality of a finite set $A$. We denote the exact order of a non-zero trigonometric polynomial $T(x)$ by $\deg(T)$. The Lebesgue measure of $E\subset \mathbb{R}^{1}$ is denoted by $\operatorname{mes}(E)$.

If $x, y\in (0, 2\pi]$, then $\operatorname{dist}(x,y)$ stands for the distance around the circle, that is,

Let $\varepsilon >0$ be a real number and $X$ a subset of $(0,2\pi]$. We define $O_{\varepsilon}(X)$ as follows. Put

when $\varnothing \neq X\subset (0,2\pi]$ and $O_{\varepsilon}(X)=\varnothing$ when $X=\varnothing$. We also define $\operatorname{Conn}(X)$ for $X\subset (0,2\pi]$: put $\operatorname{Conn}(X)=k$ when $X$ is a non-empty subset of $(0,2\pi]$ having $k\in \mathbb{N}$ components with respect to the topology of the circle and $\operatorname{Conn}(X)=0$ when $X=\varnothing$.

Let $K_{0}(x)\equiv 1$. For a positive integer $n\geq 2$ put

The function $K_{n-1}(x)$ is called the Fejér kernel of order $n-1$. We mention some properties of the Fejér kernel. If $n$ is a positive integer, then $K_{n-1}(x)$ is a non-zero real positive even trigonometric polynomial, $\deg(K_{n-1})= n-1$, $K_{n-1}(0)=n$, $\|K_{n-1}\|_{\infty}=n$, $\|K_{n-1}\|_{1}=2\pi$, and the following inequality holds:

Equation (1.1)

Let $N$ be a positive integer. Given $x=(x_1,\dots, x_N)\in \mathbb{R}^{N}$, put

We denote by $(v)_i$, $1\leq i \leq N$, the $i$th coordinate of $v\in \mathbb{R}^{N}$. We write $\dim V$ for the dimension of a finite-dimensional vector space $V$ over $\mathbb{R}$.

Put

Let $n$ be a positive integer. We denote by $\mathrm{T}(n)$ the vector space of real trigonometric polynomials of the form

and by $E_n$ the vector space of real trigonometric polynomials of the form

It easy to see that $\mathrm{T}(n)$ and $E_n$ are finite dimensional and that $\dim \mathrm{T}(n)=2n+1$ and $\dim E_n=2^n$. It is also evident that

§ 2. Some trigonometric polynomials with extremely small uniform norm

We first note the following fact. Let $\alpha>0$ be a real number and $\{t_n(x)\}_{n=1}^{\infty}$ a sequence of orthogonal trigonometric polynomials such that $\|t_n\|_{1}\geq \alpha$, $n=1, 2, \dots$ . Then

Thus, under the assumptions above, the uniform norm of $\sum_{j=1}^{n}t_{j}(x)$ cannot have order of growth smaller than $\sqrt{n}$. We have the following result.

Theorem 1.  Let $\lambda>1$ be a real number, $f\colon [1, +\infty)\to \mathbb{R}$ a positive non-increasing function, and $\{m_n\}_{n=1}^{\infty}$ a sequence of positive integers such that

Also, let a sequence $\{d_n\}_{n=1}^{\infty}$ of positive integers satisfy the condition $1\leq d_n\leq m_{n+1}-m_n$, $n=1,2,\dots$ . Then there is a sequence $\{\delta_{n}(x)\}_{n=1}^{\infty}$ of complex trigonometric polynomials such that

where $\alpha>0$ is an absolute constant.

Remark 1.  Theorem 1 is a generalization of a result in [1], where the corresponding assertion is obtained in the case when $f(n)\equiv1$. The proof of Theorem 1 is based on ideas in [1]–[3].

Proof of Theorem 1. A celebrated result of Carleson and Hunt (see [4]) implies that if $g\in L^{2}(0,2\pi)$, and

then

Equation (2.1)

where $c_{H}>0$ is an absolute constant. It is easy to see that $c_{H}\geq 1$. Note that if $n>j$ are positive integers, then

Equation (2.2)

Put $\alpha=\sqrt{72c_{H}}$. We will construct trigonometric polynomials $\delta_n(x)$, $n=1, 2, \dots$, such that

Equation (2.3)

Equation (2.4)

It is obvious that this will prove the theorem. There is no loss of generality in assuming that $d_j \leq m_j$, $j=1,2,\dots$ . Put

Note that $\widetilde{d}_{j}$ is a positive integer and

Equation (2.5)

We proceed to construct by induction trigonometric polynomials $\delta_{n}(x)$, $n=1, 2, \dots$, satisfying (2.3) and then show that (2.4) also holds. Put $\delta_{1}(x)=e^{i m_1 x}$. It is clear that it satisfies (2.3). Assume that $n\geq 2$ is a positive integer and $\delta_{1}(x),\dots, \delta_{n-1}(x)$ have been constructed. We introduce the notation

Put

Equation (2.6)

Equation (2.7)

Equation (2.8)

Equation (2.9)

Equation (2.10)

When $n-\tau_n<1$, the union in (2.8) is defined as the empty set.

We claim that

Equation (2.11)

Indeed, if $\widetilde{B}_{n}=\varnothing$, then $|\Lambda_n|=d_n$ and so (2.11) holds. Suppose that $\widetilde{B}_{n}\neq\varnothing$. Then

is not empty and

We will first find an upper bound for $\operatorname{mes}(B_n)$. Define

It is evident that

Then we have (see (2.6), (2.7))

Applying Chebyshev's inequality and (2.1), and then using the orthogonality of the polynomials $\delta_j$ and the bound $\|\delta_j\|_{\infty}\leq 6$, $j=1,\dots,n-1$, we obtain

Equation (2.12)

Let $j\in W$. Then

The trigonometric polynomial $S_j (x)$ is non-zero since the spectra of $\delta_{k}(x)$, $k=1,\dots, j$, are pairwise disjoint and $\delta_{1}(x)=e^{i m_1 x}$. Thus there is an $x_0 \in (0,2\pi]$ such that $S_{j}(x_0)\neq 0$. Therefore $|S_{j}(x_0)|>0$ and so $|S_{j}(x_0)|^{2}>0$. Hence $|S_{j}(x)|^{2}$ is a non-zero real trigonometric polynomial. It is easy to see that $\deg (|S_j|^2)\leq m_j +2[(d_j-1)/2]$. Put $T_{j}(x)=|S_{j}(x)|^{2}-\alpha^{2}n$. It is obvious that

Equation (2.13)

and $A_{n}^{j}\neq\varnothing$ since $j\in W$. Therefore there is an $x_1\in (0,2\pi]$ such that $T_{j}(x_1)>0$. Hence $T_{j}(x)$ is a non-zero real trigonometric polynomial and

It follows from well-known facts about trigonometric polynomial (see, for example, [5], V. 2, Ch. X, Theorem 1.7) that the number of zeros of $T_{j}(x)$ in $(0,2\pi]$ (counted with multiplicities) is equal to $R_j\leq 2(m_j +2[(d_j-1)/2])$. Note that

Therefore we can find an $x_2\in(0,2\pi]$ such that $T_j(x_2)<0$. Since there is an $x_1\in (0,2\pi]$ such that $T_{j}(x_1)>0$, there is an $x_3\in(0,2\pi]$ such that $T_{j}(x_3)=0$. Hence $R_j>0$. Applying (2.13), we obtain that $A_{n}^{j}$ is the union of finitely many pairwise-disjoint open intervals on the circle $(0,2\pi]$ and

Since

we have

Equation (2.14)

Combining (2.7), (2.8), (2.12), (2.14) and (2.2), we obtain

To continue estimating $\operatorname{mes}(\widetilde{B}_{n})$ we need the inequality

Equation (2.15)

(with $q=\lambda_n$ and $k = \tau_n$). This bound follows easily from the equality

for all $q>1$ and $k=0, 1, 2,\dots$, as is easily checked. Also, using the fact that

Equation (2.16)

we have

We now claim that

Equation (2.17)

Two cases may occur.

1) Let $\ln \lambda_n \leq e^{-1/15}$. Then $\rho_n\geq 1/ \ln \lambda_n\geq e^{1/15}$. Hence,

Therefore,

2) Let $\ln \lambda_n > e^{-1/15}$. Then $2/\ln \lambda_n < 2 e^{1/15}=2.1378\dots$ . Therefore,

Thus (2.17) is proved.

It is not hard to see that $f(x)=x^2 q^{-x}$ is decreasing when $x\geq 2/ \ln q$, where $q>1$ is a real number. It follows from the inequality

that

Because

and also $(a+b)^2\leq 2 (a^2 + b^2)$, $a, b\in \mathbb{R}$, it follows that

Since $\ln x\leq x$ when $x>0$, we have

Thus,

We claim further that

Equation (2.18)

Two cases may occur.

1) Let $\lambda_n\geq e^{1/2}$. Then $\lambda_{n}^{-42}\leq e^{-21}$. It follows from $\rho_n \geq 1$ that

2) Let $1<\lambda_n < e^{1/2}$. Then $\rho_n\geq 1/\ln \lambda_n \geq 2$. Therefore,

Thus (2.18) is proved.

Consequently,

Equation (2.19)

This bound implies, in particular, that $\widetilde{B}_{n}\neq (0,2\pi]$. Thus $\widetilde{B}_{n}$ is the union of finitely many pairwise-disjoint open intervals on the circle $(0,2\pi]$.

We claim, in addition, that

Equation (2.20)

We have (see (2.14), (2.2), (2.16)) that

It follows from

that

Two cases may occur.

1) Let $\lambda_n\geq \sqrt{e}$. Then $\lambda_{n}^{-44}\leq e^{-22}$. It follows from $\rho_n \geq 1$ that

2) Let $1< \lambda_n < \sqrt{e}$. Then $\rho_n \geq 1/\ln \lambda_n > 2$ and so

Thus (2.20) is proved.

Put

It is obvious that the length of $I_s$ is $2\pi/d_n$ and the intervals are pairwise disjoint. Consider

the complement of the subset $\Lambda_n$ (see (2.9)). We will find an upper bound for $|\Lambda_{n}^{c}|$. It is not hard to see that $|\Lambda_{n}^{c}|$ is at most the number of intervals $I_s$, $s=1,\dots, d_n$, intersecting $\widetilde{B}_{n}$ non-trivially. Such intervals are of two types: those contained in $\widetilde{B}_{n}$ and those containing a boundary point of $\widetilde{B}_{n}$. Denote by $N_1$ and $N_2$ the numbers of intervals of the first and second types, respectively. Then (see (2.19))

Therefore,

It is clear (see (2.20)) that

Hence,

Thus,

and (2.11) is proved.

We have (see (2.10))

Let $x\in (0,2\pi]$. Then (see (2.10), (1.1), (2.5))

Hence $\|\delta_n\|_{\infty}\leq 6$. It is clear that $\delta_n (x)$ has the form (see (2.10))

Thus we have constructed trigonometric polynomials $\delta_{n}(x)$, $n=1, 2, \dots$, satisfying the conditions in (2.3).

We now claim that these trigonometric polynomials satisfy (2.4). Let $n$ be a positive integer. If $n=1$, then (2.4) holds. Suppose that $n\geq 2$. Put

Since

the function $\omega (x)$ is well defined. Let $x\in (0,2\pi]$. Two cases may occur.

1) Let $\omega (x)\geq n- \tau_n$. Then

2) Let $\omega (x)< n- \tau_n$. Then

We will estimate the last summand. Suppose that $\omega(x)+\tau_n +1 \leq j \leq n$. It follows from $\tau_j \leq \tau_n$ that $\omega(x)+\tau_j +1 \leq j \leq n$. We claim that

Equation (2.21)

Indeed, since

and $1\leq \omega(x)+1\leq j-\tau_j$, the inclusion (2.21) holds (see (2.6), (2.8)). Therefore (see (2.8), (2.9))

for any $s\in\Lambda_j$. We have (see (2.10) and also (1.1), (2.5))

Hence,

Therefore,

and inequality (2.4) is proved. This completes the proof of Theorem 1. $\Box$

Corollary 1.  Let $\lambda>1$, $0\leq \varepsilon<1$, $0<\theta < 1-\varepsilon$ be real numbers and

Then there are complex trigonometric polynomials $\delta_{n}(x)$, $n=1,2,\dots$, such that

where $C(\lambda,\varepsilon,\theta)>0$ depends only on $\lambda$, $\varepsilon$ and $\theta$.

§ 3. Some properties of the space of quasi-continuous functions

Let $f\in L^{1}(0,2\pi)$ have the Fourier series $f\sim \sum_{j=0}^{\infty}\sigma_{j}(f)$, where $\sigma_{0}(f)=a_0 (f)/2$, and

We put

Equation (3.1)

where $\{r_j (\omega)\}_{j=0}^{\infty}$ is the Rademacher system (see, for example, [6]). The norm in (3.1) is called the $\mathrm{QC}$-norm. The closure of the set of trigonometric polynomials with respect to the $\mathrm{QC}$-norm is called the space of quasi-continuous functions. This norm and space were introduced by Kashin and Temlyakov in [7], [8], where they obtained many interesting results on the properties of this norm. In particular, the following theorem was proved in [8] (see [8], Theorem 2.1). If $n$ is a positive integer and $f(x)=\sum_{j=0}^{n} t_j (x)$, where $t_j\in E_j$, $j=0,\dots,n$, then

Equation (3.2)

In connection with the problems of approximation theory (see the details in [8]), it is interesting to relate the $\mathrm{C}$- and $\mathrm{QC}$-norms. It follows from the inequality (3.2) and a result of Grigor'ev [2] that

where $c_1> 0$ is an absolute constant. Oskolkov proved (see [8]) that

where $c_2>0$ is an absolute constant. Oskolkov's example was generalized in [9]. The following question was open for a long time: do there exist subspaces $L_n\subset E_n$, $n=1,2,\dots$, such that $\dim L_n \geq \alpha \dim E_n$, $n=1,2,\dots$, and

for any $n$ and $t\in L_1\oplus \dots \oplus L_n$, where $\alpha\in (0,1)$, $A>0$, $B>0$ are absolute constants? It was shown in [10] that the answer is negative. In the present paper, we strengthen Theorem 2 in [10] (see Corollary 2 below). Our result is obtained by specifying functions of the variable $\varepsilon$ considered in Theorems 1 and 2 in [10]. This is accomplished by a significant improvement of the argument in several places; in particular, Lemmas 1, 2 and 4 of [10] are refined. Note also that our Corollary 2 is stronger than results in [11]. The following theorem is central in this section.

Theorem 2.  Let $\varepsilon\,{\in}\,(0,1)$ be a real number, $1\,{\leq}\,k_1\,{<}\,\cdots\,{<}\,k_n\,{<}\,\cdots$ a sequence of positive integers, and $L_n$ a subspace of $E_{k_n}$ such that $\dim L_n \geq \varepsilon \dim E_{k_n}$, $n=1,2,\dots$ . Then there are trigonometric polynomials $t_n \in L_n$, $n=1, 2, \dots$, such that

where $a$, $b$, $c$ are positive absolute constants.

Proof.  We need the following lemmas, of which Lemma 2 is well known. Lemma 1 can also be considered as well known, but we include a proof for completeness.

Lemma 1.  Let $1\leq k \leq N$ be positive integers and $L$ a subspace of $\mathbb{R}^{N}$, $\dim L = k$. Then there is a non-zero $x\in L$ such that

Proof.  The case when $N=1$ is trivial. Suppose that $N\geq 2$ and put

Equation (3.3)

Since $B$ is compact, we can replace $\sup$ by $\max$ in (3.3). Let $x_0$ be an element of $B$ such that $\|x_0\|_{l_{2}^{N}}=\rho$. Denote by $s=s(x_0)$ the number of coordinates of $x_0$ having absolute value $1$.

We claim that $s\,{\geq}\, k$. Assume the opposite, that is, $s\,{\leq}\, k-1$. Let $J\,{=}\,\{j_1,\dots, j_s\}$ and $I=\{i_1,\dots, i_{N-s}\}$ be tuples of positive integers such that $j_1<\dots < j_s$, $i_1<\dots < i_{N-s}$, $\{1,\dots, N\}=\{j_1,\dots, j_s\}\sqcup \{i_1,\dots, i_{N-s}\}$,

If $s=0$, then $J$ is empty. Put

It is clear that $L'$ is a subspace of $\mathbb{R}^N$ with $\dim L'=N-s$. We have

Therefore,

Put $U=L\cap L'$. If $\delta >0$ is a real number and $y\in \mathbb{R}^{N}$, define

Put

We claim further that

Equation (3.4)

Indeed, let $y\in M'$. Then $y=x_0+x$, where $x\in M$. Hence $x\in L'$ and so

We have

Note that

Therefore,

Thus,

and (3.4) is proved.

Since $x_0\in L$ and $M\subset L$, we have $M' \subset L$. Hence $M' \subset B$. Note that

Consider the ball $S_{\rho}(0)$ with centre $0$ and radius $\rho$. Then $x_0$ lies on the boundary of the ball. Also, $M'$ is either tangent to the boundary of $S_{\rho}(0)$ at $x_0$ or intersects it transversally at $x_0$. In both cases, there is a $z\in M'$ such that $z\notin S_{\rho}(0)$. Hence $\|z\|_{l_{2}^{N}}>\rho$ and $z\in B$. But this is not the case because of (3.3). This contradiction implies that $s(x_0)\geq k$, as is claimed above.

Thus $\|x_0\|_{l_{1}^{N}}\geq k$, whence $x_0 \neq 0$. Since $x_0\in B$, we have $x_0\in L$ and $\|x_0\|_{l_{\infty}^{N}}\leq 1$. Therefore,

So we can take $x_0$ as the desired vector $x$. The proof of Lemma 1 is complete. $\Box$

Lemma 2  (see [12], p. 40, Theorem 2.5). Let $n$ be a positive integer, $N=4n$, $t\in \mathrm{T}(n)$, and

Then

where $c_1$, $c_2$, $c_3$ are positive absolute constants.

The proof of Lemma 3 given below is based on Lemmas 1 and 2. It was communicated to the author by Y. V. Malykhin. We notice that, when one does not specify $c(\varepsilon)$, the conclusion of the lemma follows from results in [13].

Lemma 3.  Let $\varepsilon\in (0,1)$ be a real number, $n$ a positive integer, and $L$ a subspace of $\mathrm{T}(n)$ such that $\dim L \geq \varepsilon (2n+1)$. Then there is a trigonometric polynomial $t\in L$ such that

where $c_4 >0$ is an absolute constant.

Proof.  Put $N=4n$. Since $\varepsilon(2n+1)>0$, the subspace $L$ is not trivial, $\dim L= k$, $k\in \mathbb{N}$ and $k\geq \varepsilon(2n+1)$. Consider the map

It is not hard to show that $\varphi$ is linear and injective. Therefore $U=\varphi (L)$ is a subspace of $\mathbb{R}^{N}$ and $\dim U = k$. By Lemma 1, there is a non-zero $x\in U$ such that $\|x\|_{l_{1}^{N}}\geq k \|x\|_{l_{\infty}^{N}}$. Then $x=\varphi (t)$, $t\in L$ and $t\neq 0$ (because otherwise $x=0$). Applying Lemma 2, we have

Since

we obtain

Therefore,

where $c_4=c_1/ (2 c_3) >0$ is an absolute constant. Since $t\neq 0$, we have $\|t\|_{\infty} > 0$. Put $t_0=t/\|t\|_{\infty}$. Then $t_0\in L$, $\|t_0\|_{\infty}=1$ and $\|t_0\|_1 \geq c_4\cdot \varepsilon$. We can take $t_0$ as the desired trigonometric polynomial $t$. Lemma 3 is proved. $\Box$

Lemma 4.  Let $\varepsilon\in (0,1)$ a real number, $n$ a positive integer, and $L$ a subspace of $E_n$ such that $\dim L \geq \varepsilon 2^n=\varepsilon \dim E_n$. Then there is a trigonometric polynomial $t\in L$ such that

where $c_5 >0$ is an absolute constant.

Proof.  Put $m=2^n - 1$. Then $m$ is a positive integer, $L$ is a subspace of $\mathrm{T}(m)$ and

Since $\varepsilon/2 \in (0,1)$, it follows from Lemma 3 that there is a trigonometric polynomial $t\in L$ such that $\|t\|_{\infty}\leq 1$ and

where $c_5 = c_4 / 2 >0$ is an absolute constant. Lemma 4 is proved. $\Box$

Lemma 5.  Let $0\,{<}\,\varepsilon \,{<}\,0.03$ be a real number, $r\,{=}\,1-0.5 \varepsilon$, $m\,{\geq}\, \lceil50/\varepsilon^{3/2}\rceil\,{=:}\,\eta_{0}(\varepsilon)$ a positive integer, and $l\in\{1,\dots, 2\lceil rm\rceil\}$. Then

Proof.  Put $\Delta_1 = 0.5 \varepsilon$, $\Delta_2 = 0.2 \varepsilon$, $\Delta_3 = 0.2 \varepsilon$,

Then $r=1-\Delta_1$. Note that

Equation (3.5)

because $0< rm < m$. We introduce the sets $G=\big\{1\leq s \leq 2 \lceil rm\rceil,\, s\neq l \big\}$,

It is clear that $G=G_1 \sqcup G_2$.

Note that

Equation (3.6)

We claim that

Equation (3.7)

Since $20 \nu \leq 120/\varepsilon$, we have

and it suffices to show that

But this inequality holds because $0< \varepsilon < 0.03$. Thus (3.7) is proved.

Consider the sum

We have

We claim that if $0< h< \pi$ is a real number, then

Equation (3.8)

Indeed, it suffices to apply Parseval's identity to the function

We claim further that if $0 \leq x \leq \sqrt{\Delta_2}$, then

Equation (3.9)

Indeed, if $x=0$, then (3.9) is trivial. Suppose that $0< x \leq \sqrt{\Delta_2}$. In view of the well-known inequality $\sin t \geq t- t^3/6$ when $t\geq 0$, it suffices to prove that $x- x^3/6 \geq \gamma x$. The last is equivalent to

Since

it is enough to show that $x^2 \leq 5 \Delta_2$. But this inequality holds because $0 < x \leq \sqrt{\Delta_2}$. Thus (3.9) is proved.

We now claim that

Equation (3.10)

It suffices to prove that $r m \geq \pi \nu / (2 \sqrt{\Delta_2})$. Since $rm >m/2\geq \eta_{0}(\varepsilon)/2$, it is enough to show that $\eta_{0}(\varepsilon)\geq \pi \nu/\sqrt{\Delta_2}$. It is enough to show, in turn, that

Applying (3.6), we obtain

Thus the inequality (3.10) is proved.

Therefore,

when $s\in \{1,\dots, \nu\}$ and hence (see (3.9))

Then

Applying (3.8) with $h=\pi m / (2 \lceil rm\rceil)$ (it is clear that $0< h < \pi$) and using the fact that $m\geq \eta_0 (\varepsilon)\geq 50/\varepsilon^{3/2}$, we have

Next, we will find a bound for $S_2$. Applying (1.1), (3.5) and $\nu\geq 5/\varepsilon$, we deduce (see also the definition of $G_2$) that

We finally have

Since $0< \varepsilon < 0.03$, we obtain

Therefore,

Lemma 5 is proved. $\Box$

Lemma 6.  Let $\varepsilon$, $0<\varepsilon <0.03$, be a real number, $r=1-0.5 \varepsilon$, $m\geq \lceil50/\varepsilon^{3/2}\rceil$ a positive integer, $\alpha_s, \beta_s \in \mathbb{R}$, $s=1,\dots, 2 \lceil rm \rceil $, and

Then

where $A>0$ is an absolute constant.

Proof.  We first claim that if the real trigonometric polynomial

satisfies $\|T\|_{\infty} \leq 1$, then

Equation (3.11)

where $d(\varepsilon)$ is as in Lemma 5. Let the absolute value of $\gamma_l$ be the largest for the $\gamma_s$, $s=1,\dots, 2 \lceil rm\rceil$. If $|\gamma_l| = 0$, then (3.11) is trivial. Suppose that $|\gamma_l| > 0$. Put

Then $T(x)= \gamma_l \cdot T_{1}(x)$. Hence,

Equation (3.12)

We have

Applying Lemma 5, we obtain

Therefore,

Thus (see (3.12))

and (3.11) is proved.

We continue with the proof of the lemma. Put

By hypothesis, $\|P\|_{\infty}\leq 1$. Consider the points

Then $\cos (3m x_\nu)=\pm 1$, $\sin (3m x_\nu) = 0$, $\nu = 1,\dots, 6 m$, and so

Since $\deg(K_{m-1}) = m-1$, it follows from the celebrated theorem of Marcinkiewicz (see [5], V. II, Russian p. 54, Theorem 7.28) that

where $A_2 > 0$ is an absolute constant. Therefore,

It follows from the argument above that $|\alpha_s / A_2|\leq 1/ (1-d(\varepsilon))$, $s=1,\dots$, $2 \lceil rm \rceil$, that is,

Put

Then $\cos (3m \widetilde{x}_\nu)=0$, $\sin (3m \widetilde{x}_\nu)=\pm 1$, $\nu = 1, \dots, 6m$, and so

Again applying Marcinkiewicz' theorem, we obtain

Therefore, $\|P_2 / A_2\|_{\infty} \leq 1$ and so $|\beta_s / A_2|\leq 1/(1-d(\varepsilon))$, $s=1,\dots, 2 \lceil rm\rceil$, that is,

By Lemma 5, we have

where $A=50A_2 / 9 > 0$ is an absolute constant. Lemma 6 is proved. $\Box$

Lemma 7.  Let $0<\varepsilon <0.03$ be a real number, $r=1-0.5 \varepsilon$, $m\geq \lceil50/\varepsilon^{3/2}\rceil$ a positive integer, $\alpha_s, \beta_s \in \mathbb{R}$, $s=1,\dots, 2 \lceil rm \rceil$, and

Equation (3.13)

Then $\alpha_s=0$, $\beta_s = 0$, $s=1,\dots, 2 \lceil rm \rceil$.

Proof.  If $\alpha_s=0$, $\beta_s = 0$, $s=1,\dots, 2 \lceil rm \rceil$, then (3.13) holds. Suppose that $(\alpha_1,\beta_1,\dots,\alpha_{2 \lceil rm \rceil}, \beta_{2 \lceil rm \rceil})$ is a non-zero tuple of real numbers such that (3.13) holds. Suppose for definiteness that $\alpha_l \neq 0$. Put

Let $\lambda > 0$ be a real number. Then $\lambda T(x) \equiv 0$. By Lemma 6, we have $|\lambda \alpha_l|\leq A/\varepsilon$. But this is not the case when $\lambda$ is sufficiently large (for example, when $\lambda = (2A)/ (\varepsilon |\alpha_l|)$). We have a contradiction. Thus (3.13) holds only when $\alpha_s = 0$, $\beta_s=0$, $s=1,\dots, 2 \lceil rm\rceil$. Lemma 7 is proved. $\Box$

We continue with the proof of the theorem. Suppose that $0 < \varepsilon < 0.03$. Put

Equation (3.14)

where $A$ is the absolute constant in Lemma 6 and $c_H \geq 1$ is the absolute constant in (2.1). Then

Equation (3.15)

We have

Hence $\tau \geq 30$ is a positive integer.

We claim that

Equation (3.16)

It is not hard to see that $2^{x/2}\geq x^2$ when $x\geq 16$. Hence,

Therefore,

The last inequality is equivalent to $\tau> 2 \log_{2} (5440 / \varepsilon)$, which obviously holds. Thus (3.16) is proved.

Note that if $ j< n$ are positive integers, then

Equation (3.17)

We define

It is clear that $m_j\in \mathbb{N}$, $j\geq\tau+ \eta_0+4$.

We write

Equation (3.18)

where $c_5$ is the absolute constant in Lemma 4. Thus $\widetilde{a}$, $\widetilde{b}$ are $\widetilde{c}$ are positive absolute constants. We put

Equation (3.19)

Hence $a$, $b$ are $c$ are positive absolute constants.

We will construct trigonometric polynomials $t_{n}(x)$, $n=1, 2, \dots$, such that

Equation (3.20)

Equation (3.21)

We use induction to construct trigonometric polynomials $t_{n}(x)$, $n=1, 2, \dots$, satisfying the conditions in (3.20) and then show that (3.21) also holds. Suppose that $j\in\{1,\dots, \tau + \eta_0 +4 \}$. By Lemma 4, there is a trigonometric polynomial $t\in L_j$ such that $\|t\|_{\infty} \leq 1$, $\|t\|_1 \geq c_5 \varepsilon\geq 0.4c_5\varepsilon$. We define $t_j = t$. Thus $t_{j}(x)$, $j=1,\dots, \tau+\eta_0 +4$ have been constructed. We now proceed by induction. Suppose that $n\geq \tau + \eta_0 +5$ is a positive integer and $t_1 (x),\dots,t_{n-1} (x)$ are given. We shall now construct $t_n (x)$. Write

Put

Equation (3.22)

Equation (3.23)

Equation (3.24)

Equation (3.25)

We claim that

Equation (3.26)

Indeed, if $\widetilde{B}_n = \varnothing$, then

and (3.26) is proved.

Let $\widetilde{B}_n \neq \varnothing$. Then the set

is not empty and

We will first find an upper bound for $\operatorname{mes}(B_n)$. Put

It is obvious that

We have (see (3.22), (3.23))

Applying Chebyshev's inequality and (2.1), and then using the orthogonality of the polynomials $t_j$ and the bound $\|t_j\|_{\infty}\leq 1$, $j=1,\dots,n-1$, we obtain

Equation (3.27)

Let $j\in W$. As in the proof of Theorem 1, one can show that $A_{n}^{j}$ is the union of finitely many pairwise-disjoint open intervals on the circle $(0,2\pi]$ and

Equation (3.28)

Taking (3.23), (3.24), (3.27) and (3.28) into account, we obtain

Applying (3.17) and (2.15) with $q= 2$ and $k=\tau$, we have

Equation (3.29)

This bound implies, in particular, that $\operatorname{mes}(\widetilde{B}_{n}) < 2\pi$ and so $\widetilde{B}_{n} \neq (0, 2\pi]$. Indeed (see (3.15), (3.16)),

Thus $\widetilde{B}_{n}$ is the union of finitely many pairwise-disjoint open intervals on the circle $(0,2\pi]$. We have (see (3.28), (3.17))

Equation (3.30)

Put

It is obvious that the length of $I_s$ is $\pi / \lceil r m_n\rceil$ and the intervals are pairwise disjoint. Consider

the complement of the subset $\Lambda_n$ (see (3.25)). We will find an upper bound for $|\Lambda_{n}^{c}|$. It is not hard to see that $|\Lambda_{n}^{c}|$ is at most the number of $I_s$, $s=1,\dots,2 \lceil r m_n\rceil$, intersecting $\widetilde{B}_{n}$ non-trivially. Such intervals are of two types: those contained in $\widetilde{B}_{n}$ and those containing a boundary point of $\widetilde{B}_{n}$. Denote by $N_1$ and $N_2$ the numbers of intervals of the first and second types, respectively. Then (see (3.29))

Since $\lceil rm_n\rceil \leq m_n$, we obtain

It is clear (see (3.30)) that

Therefore,

Hence (see (3.15), (3.16)),

and (3.26) is proved.

We define $U$ as the linear hull (over $\mathbb{R}$) of

It is obvious that $U$ is a subspace of $E_{k_n}$. Since $k_n \geq n$, we have

It follows from Lemma 7 that $\dim U\,{=}\,2 |\Lambda_n|\,{\geq}\,(1\,{-}\,0.6 \varepsilon)2^{k_n}$. Hence,

By Lemma 4, there is a $t\in U\cap L_n$ such that $\|t\|_{\infty} \leq 1$, $\|t\|_1 \geq 0.4 c_5 \varepsilon$. Put $t_n = t$. Since $t_n \in U$, we have

Equation (3.31)

where $\alpha_{s}^{(n)}$, $\beta_{s}^{(n)}$, $s\in \Lambda_n$, are real numbers. Since $m_n > \eta_0$ and $\|t_n\|_{\infty}\leq 1$, it follows from Lemma 6 that

Equation (3.32)

Thus we have constructed trigonometric polynomials $t_{n}(x)$, $n=1, 2, \dots$, satisfying the conditions in (3.20).

We now claim that these trigonometric polynomials satisfy (3.21). Let $n$ be a positive integer. If $1 \leq n \leq \tau + \eta_0 + 4$, then

Suppose that $n \geq \tau + \eta_0 + 5$. Put

Since

the function $\omega (x)$ is well defined. Let $x\in (0,2\pi]$. Two cases may occur.

1) $\omega (x) \geq n - \tau - \eta_0 - 4$. Then

2) $\omega (x) < n - \tau - \eta_0 - 4$. Then

We will estimate the last summand. Suppose that $\omega(x) + \tau + \eta_0 +5 \leq j \leq n$. We claim that

Equation (3.33)

Indeed, since

and $1\leq\omega (x)+1 \leq j -\tau $, the inclusion (3.33) is proved (see (3.22), (3.24)). Therefore (see (3.24), (3.25)), we obtain

Equation (3.34)

for any $s\in \Lambda_j$. We have (see (3.31), (3.32), (1.1), (3.34)) that

Hence,

We finally have that in the second case

and (3.21) is proved. Thus we have constructed trigonometric polynomials $t_n (x)$, $n=1, 2, \dots$, satisfying (3.20) and (3.21).

It is easy to see that $\log_{2} x \leq x$ when $x \geq 1$. Since $5440/ \varepsilon > 5440$, we have $\log_{2}(5440/\varepsilon) > 1$. Hence (see (3.14)),

Let $n\in \mathbb{N}$. We obtain (see (3.18))

Thus we have constructed a sequence $\{t_n(x)\}_{n=1}^{\infty}$ of trigonometric polynomials such that

It is evident that the constants $\widetilde{a}$, $\widetilde{b}$ and $\widetilde{c}$ can be replaced by $a$, $b$ and $c$, respectively (see (3.19)).

Now suppose that $0.03 \leq \varepsilon < 1$. Then

It follows from the argument above that there is a sequence $\{t_n (x)\}_{n=1}^{\infty}$ of trigonometric polynomials such that

for any positive integer $n$. Theorem 2 is proved. $\Box$

Theorem 3.  Let $\varepsilon\,{\in}\,(0,1)$ be a real number, $1\,{\leq}\,k_1\,{<}\,\cdots\,{<}\, k_n\,{<}\,\cdots$ a sequence of positive integers, and $L_n$ a subspace of $E_{k_n}$ such that $\dim L_n \geq \varepsilon \dim E_{k_n}$, $n=1,2,\dots$ . Then

Equation (3.35)

where $a$, $b$, $d$ and $\gamma$ are positive absolute constants and, moreover, $a$ and $b$ are the constants in Theorem 2.

Proof.  By Theorem 2, there is a sequence $\{ t_{n}(x)\}_{n=1}^{\infty}$ of trigonometric polynomials such that

where $a$, $b$ and $c$ are positive absolute constants. Let $n$ be a positive integer. Put $t(x)=t_{1}(x)+\dots + t_{n}(x)$. Then $t\in L_1 \oplus \dots \oplus L_n$ and

Since the spectra of the trigonometric polynomials $t_j (x)$, $j=1,\dots, n$, are pairwise disjoint and $\|t_1\|_1\geq c \varepsilon >0$, we obtain that $t(x)$ is a non-zero real trigonometric polynomial. By the inequality (3.2),

where $d=c/(48\pi)$ is a positive absolute constant. Thus

and the first inequality in (3.35) is proved. Further,

where $\gamma=d/ (a+b)$ is a positive absolute constant. Thus the second inequality in (3.35) is also proved. The proof of Theorem 3 is complete. $\Box$

Corollary 2.  Suppose that sequences $\{\tau_n\}_{n=1}^{\infty}$ and $\{\varepsilon_n\}_{n=1}^{\infty}$ of real numbers sasisfy the conditions

i) $\tau_n \geq 1$, $\tau_n \leq \tau_{n+1}$ ($n=1, 2, \dots$) and $\tau_n \to +\infty$ when $n\to +\infty$;

ii) $\varepsilon_n \in (0,1)$, $\varepsilon_n \geq \varepsilon_{n+1}$ ($n=1, 2, \dots$) and, starting at some $N$,

Let $L_n$ be a subspace of $E_n$ such that

Then

where $\rho$ is a positive absolute constant. In particular,

Proof.  Let $n \geq N$ be a positive integer. Write $\varepsilon = \varepsilon_n$. Then $\varepsilon \in (0,1)$ is a real number and, by condition ii), the inequality $\varepsilon_j \geq \varepsilon$ holds for any positive integer $j$, $1\leq j \leq n$. Put $\widetilde{L}_{j} = L_j$ when $1\leq j \leq n$ and $\widetilde{L}_j = E_j$ when $j>n$.

We claim that $\widetilde{L}_{j}$ is a subspace of $E_j$ and

Equation (3.36)

Indeed, if $1 \leq j \leq n$, then $\widetilde{L}_j = L_j$ is a subspace of $E_j$ and

If $j>n$, then $\widetilde{L}_j = E_j$ is a subspace of $E_j$ and

and so (3.36) is proved.

Putting $k_j= j$, $j=1, 2, \dots$, in Theorem 3 we have

In particular,

when $j=n$. Since $n\geq N$,

Thus,

where $\rho = d/ (a+b)$ is a positive absolute constant. Corollary 2 is proved. $\Box$

The author grateful to the referee for his careful reading of the paper.

Please wait… references are loading.
10.1070/IM8887