1 Introduction
In simplistic terms, a signal might be discrete, such as letters or digits sequences, or it might be analogue, i.e. continuous functions, reading such as a temperature or a pressure. It is a remarkable fact that under additional assumptions, analogue signals and discrete signals are equivalent: an analogue signal f can be recovered exactly from its samples {f(xj)}ωj=−ω, i.e. from a digital signal. This is the essence of the sampling theorems. These theorems are fundamental, in particular, in information theory and communication, particularly since the advent of modern digital computers.
An analogue signal is, in the classical sense, a function such that it is continuous with respect to a real variable (for example, time) and has finite energy, i.e. it is a square-integrable on R. The following classical Shannon sampling theorem (see, e.g. Higgins, 1996; p. 51) states that if f∈L2(R) is bandlimited to [−σ,σ], σ>0, i.e. if the Fourier transform of f is supported on [−σ,σ], then
and this series converge in L2(R)-norm and uniformly on R. This theorem was the starting point for many further developments in sampling theory. Series (1.1) is called a cardinal series of f. The sequence {πk/σ}∞k=−∞ is called a sampling sequence and the corresponding values f(πk/σ) are called sample values with Nyquist sampling rate of σ/π samples per unit time. The first problem is to find conditions under which f can be reconstructed completely by (1.1). In most applications, the assumption that f is bandlimited, or equivalently that the spectrum of f is supported to certain bounded area in R, is completely justified. More precisely, if a signal f is bandlimited to [−σ,σ], 0<σ<∞, or in other words, the quantity σ is the maximal frequency in the spectrum of f, then (1.1) perfectly recovers f by samples, taken every π/σ seconds.
There are several ways by which (1.1) may be generalized (see, e.g. Zayed and Schmeisser, 2014). For example, in 1955, Fogel (Jagerman and Fogel, 1956) was motivated by a problem in aircraft instrument communications (a pilot in the case of pointer-on-scale displays may estimate the pointer position and also the rate information concerning the acceleration of the pointer, corresponding to first- and second-time derivatives) and studied a sampling expansion which involved sample values of a function and its derivatives.
Next, one possible application of sampling theory is in sensor networks. In this case, a large number of sensors is used to monitor some quantity, e.g. temperature. This quantity varies continuously in space and hence, can be viewed as a signal in multidimensional space.
In this paper, we concern with the sampling representation for bandlimited signals f of several variables such that f has finite p-energy for certain 1⩽p<∞ (see Nashed and Sun, 2010 and Nguyen and Unser, 2017).
Note that the reconstruction (1.1) is called stable, if it converges uniformly on R. Such a stability is important not only from the theoretical point of view but, in particular, for applications. If (1.1) is stable, then it is possible to bound the approximation error, which is made by using in (1.1) only finitely many samples. Finally, we note that our sampling series have such a stable reconstruction properties. Namely, they converge absolutely and uniformly on the whole Rn.
2 Definitions, Notation and the Main Result
Let Zn, Rn and Cn be the usual n-dimensional integer lattice, real Euclidean space and complex Euclidean space, respectively. Next, Lp(Rn), 1⩽p⩽∞, denotes the Lebesgue space of complex-values measurable functions on Rn with finite p-energy, i.e. such that |f|p is integrable on Rn. For f∈L1(Rn), we define the Fourier transform of f as usual by
x∈Rn, where ⟨x,t⟩=∑nk=1xktk is the scalar product on Rn. If f∈Lp(Rn) and 1<p⩽∞, then we understand ˆf in a distributional sense of tempered distributions S′(Rn). We recall that the Schwartz space of test functions S(Rn) consists of the complex-valued infinitely differentiable functions φ on Rn satisfying
for all nonnegative integers m and k, also for any nonnegative multi-index α, where Dα is the partial derivative of order α and ‖x‖ denotes the standard Euclidean norm on Rn. The dual space S′(Rn) of S(Rn) is called the space of tempered distributions.
Given a closed subset Ω of Rn, a function ω:Rn→C is called bandlimited to Ω if ˆω vanishes outside Ω. For σ=(σ1,…,σn)∈Rn with σm>0, m=1,…,n, we define the spaces of bandlimited signals by
where
Hence, any f∈BpQnσ is bandlimited to hyperrectangle or n-orthotope Qnσ, i.e. ˆf vanishes outside Qnσ. If we equip BpQnσ with the norm
then BpQnσ is a Banach space and it is called the Bernstein space. By the Paley–Wiener–Schwartz theorem (see Hörmander, 1990; p. 181), any f∈BpQnσ, 1⩽p⩽∞, is infinitely differentiable on Rn and has an (unique) extension onto the complex space Cn to an entire function.
Remark 1.
Recall that the identity theorem for analytic functions shows that the extension of f∈BpQnσ from Rn to Cn is unique. Therefore, we can identify further each f∈BpQnσ with bandlimited function defined on Rn and in other cases consider the same f as entire function defined on the whole Cn. Note that terms “bandlimited signal on Rn”and “bandlimited function on Rn” are equivalent.
Below we shall use the technique of complex variable functions on Cn. In particular, we shall need certain facts in entire functions. Therefore, we use below only function theory terms, i.e. “bandlimited function” term.
Next, to simplify the notation, we shall write Bpσ in the case of functions of one variable Bpσ instead of BpQ1σ.
If we define the sinc-function to be
z∈C, then (1.1) can be rewritten as
This expression holds for each f∈Bpσ, 1⩽p<∞. Next, (2.1) converges absolutely and uniformly on compact subsets of C. The Nyquist sampling rate of σ/π samples per unit time in (2.1) is exact, i.e. f∈Bpσ cannot be recovered from its samples taken at a lower rate. This means that, for any ε>1, there exist two fj∈Bpσ, j=1,2 such that f1≢f2, but f1 coincides with f2 on every point of the sampling sequence {(επk/σ}k∈Z.
On the other hand, if we know sample values of f∈Bpσ only the sampling sequence {2πk/σ}k with the sampling rate σ/(2π), then reconstruction of f is also possible if we know, in addition, its derivative values {f′(2πk/σ)}k. More precisely, if f∈Bpσ, 1⩽p<∞, then
(see Jagerman and Fogel, 1956; p. 145 or Butzer et al., 2011; p. 442).
The following n-dimensional sampling theorem is a standard extension of (2.1) in BpQnσ
z∈Cn (Gosselin, 1977; p. 172) (see also Jerri, 2017 for references). Here and below we are taking
for any τ∈C and all a,b∈Cn such that b1≠0,…,bn≠0. These series converge absolutely and uniformly on compact subsets of Cn.
The aim of this paper is to prove a multidimensional version of (2.2). In the case of two variables, in Fang and Li (2006) (see also a tutorial review Jerri, 2017; p. 40) a version of (2.2) involving only the following sample values sequences was given
Namely, in Fang and Li (2006) the following representation was given:
for all f∈BpQ2σ, 1⩽p<∞. We say that such a sampling theorem fails in general. Indeed, let us take any χ∈S(R2), χ≢0, such that suppχ⊂{x∈R2:|xk|⩽σk/2,k=1,2} and define
x∈R2. Since S(R2) is invariant under the Fourier transform, it follows that ˆχ∈S(R2), and consequently Λ is in BpQ2σ for all 1⩽p⩽∞. On the other hand, if f=Λ, then all sequences in (2.3) are null sequences. Hence, (2.4) generates null function, but not Λ. Even more, this example with our function Λ shows that the same is still true if we add to (2.4) an arbitrary number of sample values sequences
k=1,2, j=2,3,… . Therefore, any multidimensional version of (2.2) must necessarily contain also mixed partial derivatives of f.
(2.4)
f(z)=∑u∈Z2[f(2πuσ)+(z1−2πu1σ1)⋅∂f∂x1(2πuσ)+(z2−2πu2σ2)⋅∂f∂x2(2πuσ)]×sinc2(σ12πz1−u1)sinc2(σ22πz2−u2)Our basic idea based on observation that any sampling series are also an interpolation formula. For example, the m-th coefficient in (2.1) equals to the value of the sum, i.e. the value of f at the point with the number m, i.e. at πm/σ.
We now state the main result of this paper.
Theorem 1.
If f∈BpQ2σ, 1⩽p<∞, then
The series (2.5) converge absolutely and uniformly on R2 and on compact subsets of C2.
(2.5)
f(z)=∑u∈Z2[f(2πuσ)+(z1−2πu1σ1)⋅∂f∂x1(2πuσ)+(z2−2πu2σ2)⋅∂f∂x2(2πuσ)+(z1−2πu1σ1)(z2−2πu2σ2)⋅∂2f∂x1∂x2(2πuσ)]×sinc2(σ12πz1−u1)sinc2(σ22πz2−u2).3 Preliminaries and Proofs
Set Q2π={x∈R2:|x1|,|x2|⩽π}. We may assume without loss of generality below that σ1=σ2=π, since the operator
is an isometric isomorphism between BpQ2σ and BpQ2π.
Let 1⩽p<q⩽∞. Then there exists 0<M(p;q)<∞ such that
for all f∈BqQ2π (see Triebel, 1983; pp. 21–22). Therefore,
If f∈BpQ2π with 1⩽p<∞, then
(see Nikol’skii, 1975; p. 118). If f∈B∞Q2π, then (see, e.g. Nikol’skii, 1975; p. 117; Jagerman and Fogel, 1956; p. 181 or Hörmander, 1990; p. 181).
for all z=x+iy, x,y∈R2. Note that (3.2) shows that (3.4) also holds true for any f∈BpQ2π, 1⩽p<∞.
In the sequel, we shall use several times the following function in B∞Q2π/2:
Let us start by proving a simple auxiliary statement for functions of one variable. For completeness, we also give its proof.
Proof.
Define
Then (3.5) implies that G is a well-defined entire function on C. Now, for each positive integer m, set
It can easily be checked that
for each m=1,2,… . On the other hand, according to (3.2) and (3.4), we have that there is a finite number M1>0 such that
for all z∈C. Therefore, by the maximum principle for analytic functions, we see that |G| is bounded on Dm. Moreover, (3.7) and (3.8) imply that |G| is bounded by the same constant on each Dm. Thus, |G| is bounded on
Now take any z∈C∖D. Then it is easy to see that
Combining this with (3.8), we conclude that |G| is also bounded on C∖D. By Liouville’s theorem, G is a constant. Therefore, (3.6) implies that there exists c∈C such that F(z)=csin2(πz/2) on C. Finally, using (3.3), since F∈Bpπ, 1⩽p<∞, we see that c=0. Thus, F≡0. □
Proof.
First we claim that there is an entire function H:C2→C such that
for all z∈C2. To that end, we expand F in a series of homogeneous polynomials
where n∈Z2 is a non-negative multi-index and |n|=n1+n2. Note that (3.13) converges uniformly on compact subsets of C2 (see, e.g. Shabat, 1992; p. 36). In particular, the first condition in (3.10) shows that
for all z2∈C. Using this and the identity theorem for entire function ˜F(λ):=F(0,λ) on C, we see that (3.14) is equivalent to the condition cn=0 in (3.13) for each n=(n1,n2) such that n1=0. Hence in (3.13) we have that p0=0 and pm=z1qm, m=1,2,… , where qm is a homogeneous polynomial of degree m−1 or qm≡0. Next, using the second condition in (3.10), we obtain in the same manner that p1≡0 and
m=2,3,… , where rm is a homogeneous polynomial of degree m−2 or rm≡0. The series ∑∞2rm converge uniformly on compact subsets of C2, since ∑∞0pm has this property. Therefore, ∑∞2rm defines on C2 an entire function, say H. Therefore, (3.13) implies our claim (3.12).
We denote by N(sπ) the zero set of sπ. Clearly,
Hence, the function
is well-defined on C2∖N(sπ). We claim that G can be extended to an entire function on C2. Our proof is based on the Riemann removable singularity theorem (see, e.g. Shabat, 1992; p. 175) for G and the analytic set N(sπ). Since N(sπ) is an analytic set of the codimension codim(N(sπ))=1 (see Scheidemann, 2005; p. 72), it follows that it remains to prove that G is locally bounded on N(sπ), i.e. for every z∈N(sπ) there is an open neighbourhood Uz of z such that G is bounded on (C2∖N(sπ))∩Uz.
Fix λ∈N(sπ). By (3.15), we see that the proof of the fact that G is bounded on (C2∖N(sπ))∩Uz can be divided into two following cases:
Note that if ω∈Z2, then
for all z∈C2. Next, sπ(z1,z2)=sπ(z2,z1). Moreover, the functions Fω(z)=F(z+2ω) and ˜F(z)=F(z2,z1) also satisfy (3.10). Therefore, we may assume without loss of generality that we have the only following two cases:
Suppose that λ=(λ1,λ2) with λ1=0 and λ2∉2Z. Substituting (3.12) into (3.16), we get
where
Of course, H0,H1 and H2 are entire functions on C2 such that
since in this case λ2∉2Z. Therefore, there is an ε>0 and a neighbourhood Uλ⊂C2 of λ such that
for all z∈Uλ. Using this and (3.17), we see that G is bounded on (C2∖N(sπ))∩Uz.
Proof.
Set f=f1−f2. We must prove that f≡0. In our case σ1=σ2=π. Therefore (3.20) and (3.21) are equivalent to
and
u∈Z2, respectively. First, we claim that
for all λ∈C and each k∈Z. Indeed, fix k∈Z and let us define F1(λ):=f(λ,2k) and F2(λ):=f(2k,λ) for λ∈C. Then (3.22) implies that
for all m∈Z. Since F1,F2∈Bpπ, Lemma 1 yields our claim (3.24).
Now, using Lemma 2, we see that there is an entire function g on C2 such that
Now we claim that
for each k∈Z and all λ∈C. To that end, fix k∈Z and set
λ∈C. From (3.22) we have that
for all m∈Z. Since
it follows from (3.23) that
m∈Z. Next, according to Bernstein’s inequality (Nikol’skii, 1975; p. 116), if f∈BpQ2π, then all partial derivatives of f also are elements of BpQ2π. Hence, F1,F2∈Bpπ. Therefore, using (3.27) and (3.28), we see that
by Lemma 1. On the other hand, from (3.25) we get
Together with (3.29), this implies that
for all λ∈C such that λ≠2πn, n∈Z. Since h1(λ):=g(2k,λ) and h2(λ):=g(λ,2k), λ∈C, are entire functions on C, we obtain that h1=h2≡0, which yields our claim (3.26).
Now Lemma 2 shows that then there is an entire function h on C2 such that g(z)=sπ(z)h(z). Then (3.25) implies that
Combining (3.4), (3.6) and (3.9) with the definition of sπ, we see that |h| is bounded on the whole C2. Therefore, by Liouville’s theorem, we obtain that h≡ch, ch∈C. Hence, f(z)=chs2π(z). Since f∈BpQ2π with 1⩽p<∞, it follows from (3.3) that ch=0, i.e. f≡0. This proves Theorem 2. □
Set
where C0(R2) is the usual space of continuous functions on R2 that vanish at infinity. BQ2σ is a Banach space with respect to the sup-norm inherited from C0(R2). Next, (3.3) implies that
for all 1⩽p<∞.
Let us look again at the proof of Theorem 2. Then we can see that together with (3.20) and (3.21), we used, in addition, only the following two properties of fj, j=1,2: a) ^fj is supported in Q2π; b) fj satisfies (3.3). Hence, the following corollary is true:
Next, for each 1<r<∞ and any w>0, the following estimate is true (see Splettstösser, 1982; p. 811):
for all t∈R. In particular, this implies that the series in (3.31) converge on the whole R and its sum is a bounded function on R.
Proof of Theorem 1.
To begin, we recall the following Nikol’skii’s inequality (Nikol’skii, 1975; p. 123): for any 1⩽p<∞ and each 0<ϱ<∞, there exists a finite positive number c(σ;p;ϱ) such that
for all f∈BpQ2σ. A similar inequality holds also for all partial derivatives of f, since by Bernstein’s inequality (Nikol’skii, 1975; p. 116), any partial derivative of f also is an element of BpQ2σ.
Suppose again that σ1=σ2=π and let lp2 be the usual space of sequences {cu∈C:∑u∈Z2|cu|p<∞}. First, we claim that the series (2.5) converge absolutely and uniformly on R2. Indeed, if z=x∈R2, then (2.5) can be divided into four series of the following type
with certain α∈lp2 and some ν=(ν1,ν2), where ν1,ν2∈{0,1}. Therefore, it suffices to show that for any ε>0 there is a positive integer N such that
for all x∈R2. Take any p1⩾p such that 1<p1<∞. Then Hölder’s inequality implies that
with 1<q1<∞ such that 1/p1+1/q1=1. Since α∈lp2⊂lp12, it follows that, for any ε1>0, there exists a positive integer N1 such that
On the other hand, it is clear that
for j=1,2, mj∈Z and all xj∈R. Hence, by (3.31), we see that the second series in (3.34) converge and
for all x∈R2 and each N⩾0. Combining this with (3.33), (3.34) and (3.35), we get that (3.32) converges absolutely and uniformly on R2.
(3.34)
|Iν;N(x)|⩽(∑m∈Z2|m1|,|m2|⩾N|αm|p1)1/p1(∑m∈Z2|m1|,|m2|⩾N|x12−m1|q1ν1|x22−m2|q1ν2×|sinc(x12−m1)|2q1|sinc(x22−m2)|2q1)1/q1(3.36)
(∑m∈Z2|m1|,|m2|⩾N|x12−m1|q1ν1|x22−m2|q1ν2|sinc(x12−m1)|2q1|sinc(x22−m2)|2q1)1/q1⩽(∑m1∈Z,|m1|⩾N|sinc(x12−m1)|(2−ν1)q1)1/q1)⩽((2−ν1)q1(2−ν1)q1−1)1/q1Next, for any compact subset K of C2, there is a strip Eτ={z∈C2:|ℑz1|,|ℑz2|⩽τ}, 0<τ<∞, such that K⊂Eτ. If f∈BpQ2π, then (3.2) shows that f∈B∞Q2π. Therefore, using (3.4), we obtain that (2.5) also converges absolutely and uniformly on K.
Let us denote by F the sum of (2.5). Any sum of partial finite subseries of (2.5) belongs to BpQ2π for any 1⩽p<∞. By (3.30), it is also an element of BQ2π. Since BQ2π is a Banach space and (2.5) converges absolutely and uniformly on R2, i.e. in BQ2π-norm, it follows that F∈BQ2π. Using the fact, that f∈BpQ2π⊂BQ2π for each 1⩽p<∞, it is easy to check that the functions f1:=f and f2=F satisfy the conditions of Corollary 1. Hence, f≡F and this proves our theorem. □
4 The Truncation Error
The sampling formula (2.5) requires us to know values of a signal f at infinitely many points {2πu/σ}u∈Z2. In practice, only finitely many samples are available. For N∈Z2 with certain positive integers N1 and N2, let us define the partial sum of (2.5) by
Then the truncation error of f is defined by
x∈R2. The best known EN(x) estimates are local, i.e. these estimates are valid only on certain compacts subsets of R2. We shall indicate some uniform bounds of Ef;N(x), i.e. the bounds of
Such an estimate is simpler in the case if we apply to functions f∈BpQ2σ certain additional conditions of constructiveness relating to the decay of f at infinity in (3.3) (see, e.g. Lin, 2019 and Wang et al., 2018). Note that the spectral function, i.e. the Fourier transform ˆf, of many important signals f are smooth enough. Hence, these signals f have rapid decay in the time domain for large time. In light of this, it is natural to study the truncation error for functions f∈BpQ2σ, 1⩽p<∞ that satisfy the following simple decay condition
for all x∈R2 such that , |x1|⩾N1⩾1, |x2|⩾N2⩾1. Here cf is a positive number that depends on f. Note that the function
satisfies (4.3) with cf=1/(σ1σ2).
(4.1)
fN−1(z)=∑u∈Z2|u1|⩽N1−1,|u2|⩽N2−1[f(2πuσ)+(z1−2πu1σ1)∂f∂z1(2πuσ)+(z2−2πu2σ2)∂f∂z2(2πuσ)+(z1−2πu1σ1)(z2−2πu2σ2)∂2f∂z1∂z2(2πuσ)]×sinc2(σ12πz1−u1)sinc2(σ22πz2−u2).Theorem 3.
Let f∈BpQ2σ, 1⩽p<∞, and let N1 and N2 be two positive integers. Assume that f satisfies (4.3). Then, for any 1<ω<∞, we have that
The estimate (4.4) can be substantially simplified. Indeed, the following corollary holds:
Of course, (4.5) is less exact than (4.4).
Proof of Theorem 3.
For ν=(ν1,ν2) with ν1,ν2∈{0;1}, set
u∈Z2. The truncation error EN;f(x) can be divided into four series of the type
Next, for any ω such that 1<ω<∞, Hölder’s inequality implies that
with 1<s<∞ such that 1/ω+1/s=1.
(4.7)
|Iν(x)|⩽(∑u∈Z2|u1|⩾N1,|u2|⩾N2|αν(u)|ω)1/ω(∑u∈Z2|u1|⩾N1,|u2|⩾N2|x1−2πσ1u1|sν1|x2−2πσ2u2|sν2×|sinc(σ12πx1−u1)|2s|sinc(σ22πx2−u2)|2s)1/sWe claim that
for all x∈R2. Indeed,
j=1,2, x∈R2. Therefore, by a similar argument as is used in the proof of (3.36), we obtain (4.8).
(4.8)
(∑u∈Z2|u1|⩾N1,|u2|⩾N2|x1−2πσ1u1|sν1|x2−2πσ2u2|sν2|sinc(σ12πx1−u1)|2s|×sinc(σ22πx2−u2)|2s)1/s⩽2ν1+ν2σν11σν22(s2(2−ν1)(2−ν2)[(2−ν1)s−1)][(2−ν2)s−1])1/sNow we estimate the first series in (4.6). Here we have four cases. Start with ν1=ν2=0. Then
u∈Z2. Hence, using (4.3), we get
Since
it follows from (4.9) that
Thus,
Next, assume that ν1=1 and ν2=0. Then
We shall estimate these α1,0. For this purpose let us define the function
x∈R2. According to (4.3), it is in B∞Q2σ and
Given any nonnegative multi-index ν∈Z2 with ν1,ν2∈{0;1}, we get by Bernstein’s inequality in B∞Q2σ (see Nikol’skii, 1975; p. 116) that
for all x∈R2. Now ν1=1 and ν2=0. Hence, (4.13) implies that
Hence, using (4.3), we get that
for all x∈R2. Therefore, we have that
Since
it follows from (4.15) that
Finally, with (4.9), (4.10) and (4.11) in mind, we obtain that
A similar argument shows that
Assume now that ν1=ν2=1. Then
Now (4.13) implies that
for all x∈R2. Therefore, using (4.3) and (4.14), we have that
x∈R2. Since
we obtain from (4.18) that
Now, combining (4.7), (4.8), (4.11), (4.16), (4.16) and (4.19) with the estimate
x∈R2, we have that
Finally, since
(4.20) implies (4.4). This proves our theorem. □
(4.18)
|∂2f∂x1∂x2(x)|⩽cfσ1σ2|x1||x2|+|f(x)||x1||x2|+1|x2||∂f∂x1(x)|+1|x1||∂f∂x2(x)|⩽cf|x1||x2|(σ1σ2+σ1|x2|+σ2|x1|+3|x1||x2|),
∑u∈Z2u1⩾N1,u2⩾N2|α1,1(u)|ω=∑u∈Z2u1⩾N1,u2⩾N2|∂2f∂x1∂x1(2πuσ)|ω⩽(cfσ1σ2)ω(1+12πN1+12πN2+34π2N1N2)ω×∑u∈Z2u1⩾N1,u2⩾N21|2πu1/σ1|ω|2πu2/σ2|ω⩽(cfσ21σ224π2)ω(1+12πN1+12πN2+34π2N1N2)ω(ω−1)2(N1N2)ω−1.
Therefore,
(4.19)
(∑u∈Z2|u1|⩾N1,|u2|⩾N2|α1,1(u)|ω)1/ω⩽cfσ21σ224π2(1+12πN1+12πN2+34π2N1N2)×(2(ω−1)2)1/ω(N1N2)1−1/ω.(4.20)
eN;f⩽supx∈R2|Ef;N(x)|⩽cfσ1σ24π2[(4s2(2s−1)2)1/s+2(2s2(s−1)(2s−1))1/s×(2+12πN1+12πN2)+4(s2(s−1)2)1/s×(1+12πN1+12πN2+34π2N1N2)](2(ω−1)2)1/ω(N1N2)1−1/ω.5 A Numerical Analysis of the Truncation Error
In this section, we comment the estimates (4.4) and (4.5) for the uniformly truncated error eN;f. We also give two numerical examples of eN;f by tables.
The estimates (4.4) and (4.5) depend on the signal f magnitude cf, the highest frequencies σ=(σ1,σ2) of f, the numbers N1 and N2 of samples that are used in the partial sum (4.1), and a number ω which we can choose free. Now we will briefly comment on their action on ef;N.
Magnitude {\boldsymbol{c}_{\boldsymbol{f}}}.
If f\in {B_{{Q_{\sigma }^{2}}}^{p}}, 1\leqslant p<\infty , satisfies (4.3), then the function F(x)={x_{1}}{x_{2}}f(x) is in {B_{{Q_{\sigma }^{2}}}^{\infty }}. Hence, we can choose {c_{f}} equal to the maximal amplitude of F in the domain \{x\in {\mathbb{R}^{2}}:|{x_{1}}|\geqslant {N_{1}},|{x_{2}}|\geqslant {N_{2}}\}, i.e.
However, instead of this choice for {c_{f}}, we can associate {c_{f}} with certain function relating to the p -energy of signal f. Indeed, if f satisfies (4.3), then we can take
where {\Sigma _{p;N}}(f) is a part of the p-energy of f supported in \{x\in {\mathbb{R}^{2}}:\hspace{2.5pt}|{x_{1}}|\geqslant {N_{1}},\hspace{2.5pt}|{x_{2}}|\geqslant {N_{2}}\}, i.e.
{{\Sigma _{p;N}}(f)=\bigg({\int _{|{x_{1}}|\geqslant {N_{1}},|{x_{2}}|\geqslant {N_{2}}}}{\big|f(x)\big|^{p}}\hspace{0.1667em}dx\bigg)^{1/p}}.
Therefore, we shall calculate below only the quantity {e_{N;f}}/{c_{f}} instead of the usual truncated error {e_{N;f}}.
Frequencies \boldsymbol{\sigma }\mathbf{=}\boldsymbol{(}{\boldsymbol{\sigma }_{\mathbf{1}}}\mathbf{,}{\boldsymbol{\sigma }_{\mathbf{2}}}\boldsymbol{)}.
We recall that humans can hear a range of frequencies from 20 to 20,000 Hz, i.e. we can hear signals which have 20-20000 full cicles per second. On the other hand, in the case of smartphones (cell phones) the frequency band ranges from approximately 300 Hz to 3400 Hz. Cutting out the other frequencies reduced the amount of information that would have to be transmitted and reduced the Nyquist frequency. Consequently, our smartphones have such low quality, i.e. they are not picking up all the frequencies that make up our voices or that we can hear. For this reason, we shall calculate below the estimates of {e_{N;f}}/{c_{f}} in the case {\sigma _{1}}={\sigma _{2}}=3400 Hz.
The numbers of samples {\boldsymbol{N}_{\mathbf{1}}} and {\boldsymbol{N}_{\mathbf{2}}}.
Obviously, by getting higher values of {N_{1}} and {N_{2}}, we get a better estimate of {e_{N}}(f) and {e_{N}}(f)/{c_{f}}. For example, if {N_{1}} and {N_{2}} are such that |{e_{N}}(f)/{c_{f}}|<0.01, then we can understand that the difference |f-{f_{N}}| is less than 1 percent of the maximal amplitude {c_{f}} of f. Below we calculate the estimates of {e_{N}}(f)/{c_{f}} for different values of {N_{1}} and {N_{2}} and the same values of other parameters σ and ω.
A parameter \boldsymbol{\omega }.
From the proof of Theorem 3 follows that we can choose the value of ω in (1,\infty ) free, since ω is only the technical parameter that was used in the proof of this theorem. However, a preliminary analysis of the action of ω on {e_{N;f}} can be done. Indeed, let us define
\begin{aligned}{}{A_{N,\omega }}(f)& =\bigg[{\bigg(\frac{4{\omega ^{2}}}{{(\omega +1)^{2}}}\bigg)^{1-1/\omega }}+2{\bigg(\frac{2{\omega ^{2}}}{\omega +1}\bigg)^{1-1/\omega }}\bigg(2+\frac{1}{2\pi {N_{1}}}+\frac{1}{2\pi {N_{2}}}\bigg)\\ {} & \hspace{1em}+4{\omega ^{2(1-1/\omega )}}\bigg(1+\frac{1}{2\pi {N_{1}}}+\frac{1}{2\pi {N_{2}}}+\frac{3}{4{\pi ^{2}}{N_{1}}{N_{2}}}\big)\bigg]{\big(2{(\omega -1)^{2}}\big)^{1/\omega }}\end{aligned}
and
Then the right-hand side of (4.4) is equal to
It is obvious that for given {N_{1}} and {N_{2}}, {B_{N}}(f) can be made smaller by taking the values of ω close to infinity. On the other hand, it is easy to check that
Therefore, the choice of optimized values of ω for the best estimates in (4.4) is a distinct and nontrivial problem. Here we give two numerical examples for estimate of {e_{N;f}}.Table 1
Estimatesof {e_{N;f}}/{c_{f}} in (4.4) with {N_{1}}={N_{2}}=n and with frequencies up to {\sigma _{1}}={\sigma _{2}}=3400 Hz.
\omega \setminus n | 2\cdot {10^{4}} | 5\cdot {10^{4}} | {10^{5}} | 5\cdot {10^{5}} | {10^{6}} | 2\cdot {10^{6}} |
1,25 | 10231,9 | 7092,18 | 5374,85 | 2823,44 | 2139,76 | 1621.64 |
2 | 328,501 | 131,399 | 65,6996 | 13,1399 | 6,56994 | 3.28497 |
4 | 10,6876 | 2,70375 | 0,95592 | 8,5499\cdot {10^{-2}} | 3,0229\cdot {10^{-2}} | 10,6874\cdot {10^{-3}} |
8 | 3,02323 | 0,60824 | 0,18083 | 1,0816\cdot {10^{-2}} | 0,3217\cdot {10^{-2}} | 0,9560\cdot {10^{-3}} |
12 | 2,72504 | 0,50794 | 0,14254 | 0,7455\cdot {10^{-2}} | 0,2092\cdot {10^{-2}} | 0,5871\cdot {10^{-3}} |
16 | 3,05007 | 0,54722 | 0,14919 | 0,7297\cdot {10^{-2}} | 0,1989\cdot {10^{-2}} | 0,5424\cdot {10^{-3}} |
20 | 3,60504 | 0,63215 | 0,16938 | 0,7958\cdot {10^{-2}} | 0,2132\cdot {10^{-2}} | 0,5714\cdot {10^{-3}} |
50 | 11,4879 | 1,90666 | 0,49006 | 2,0906\cdot {10^{-2}} | 0,5373\cdot {10^{-2}} | 1,3812\cdot {10^{-3}} |
Table 2
Estimatesof {e_{N;f}}/{c_{f}} in (4.5) with {N_{1}}={N_{2}}=n and with frequencies up to {\sigma _{1}}={\sigma _{2}}=3400 Hz.
\omega \setminus n | 2\cdot {10^{4}} | 5\cdot {10^{4}} | {10^{5}} | 5\cdot {10^{5}} | {10^{6}} | 2\cdot {10^{6}} |
1,25 | 13856,6 | 9604,67 | 7278,96 | 3823,68 | 2897,80 | 1153,64 |
2 | 496,926 | 198,771 | 99,3853 | 19,8771 | 9,93853 | 0,99385 |
4 | 20,4711 | 5,17884 | 1,83099 | 16,377\cdot {10^{-2}} | 5,7912\cdot {10^{-2}} | 20,471\cdot {10^{-3}} |
8 | 7,05162 | 1,41872 | 0,42179 | 2,5229\cdot {10^{-2}} | 0,7501\cdot {10^{-2}} | 2,2299\cdot {10^{-3}} |
12 | 6,88186 | 1,28277 | 0,35996 | 1,8829\cdot {10^{-2}} | 0,5284\cdot {10^{-2}} | 1,4827\cdot {10^{-3}} |
16 | 8,03339 | 1.44132 | 0.39294 | 1,9220\cdot {10^{-2}} | 0,5239\cdot {10^{-2}} | 1,4286\cdot {10^{-3}} |
20 | 9,74328 | 1,70852 | 0,45778 | 2,1510\cdot {10^{-2}} | 0,5763\cdot {10^{-2}} | 1,5442\cdot {10^{-3}} |
50 | 33,0651 | 5,48788 | 1,41054 | 6,0174\cdot {10^{-2}} | 1,5466\cdot {10^{-2}} | 3,9752\cdot {10^{-3}} |
Those tables give us certain knowledge in the case of smartphones, where the frequency band ranges up to 3400 Hz. For example, assume that we want to know what quantity N=({N_{1}},{N_{2}}) of sample values of a signal f and its derivatives in (4.1) guarantees us that |f-{f_{N}}| is less than 1 percent of the maximal amplitude {c_{f}}. In Table 1 we see that that is enough to take {N_{1}},{N_{2}}\geqslant 5\cdot {10^{5}} if the parameter ω ranges from 12 to 20.