Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Riemann Zeta-Function: Theory and Applications
The Riemann Zeta-Function: Theory and Applications
The Riemann Zeta-Function: Theory and Applications
Ebook952 pages9 hours

The Riemann Zeta-Function: Theory and Applications

Rating: 0 out of 5 stars

()

Read preview

About this ebook

"A thorough and easily accessible account."—MathSciNet, Mathematical Reviews on the Web, American Mathematical Society. This extensive survey presents a comprehensive and coherent account of Riemann zeta-function theory and applications. Starting with elementary theory, it examines exponential integrals and exponential sums, the Voronoi summation formula, the approximate functional equation, the fourth power moment, the zero-free region, mean value estimates over short intervals, higher power moments, and omega results. Additional topics include zeros on the critical line, zero-density estimates, the distribution of primes, the Dirichlet divisor problem and various other divisor problems, and Atkinson's formula for the mean square. End-of-chapter notes supply the history of each chapter's topic and allude to related results not covered by the book. 1985 edition.
LanguageEnglish
Release dateJul 12, 2012
ISBN9780486140049
The Riemann Zeta-Function: Theory and Applications

Related to The Riemann Zeta-Function

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for The Riemann Zeta-Function

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Riemann Zeta-Function - Aleksandar Ivic

    ERRATA

    p. V 1. 9 should have a point after results.

    p. 13 1. 1 should be log |z| and f(0) = 1 1. -14 is should be if

    p. 17 1. 1, absolute value signs are not needed in |rn|−1.

    p. 17 1. -2 better: Then, if T is not an ordinate of a zero,

    p. 18 1. 11 should be (A.33), not (A.35).

    p. 20 1. 1 should be (1.38), not (1.48).

    p. 27 better put: (1.71) (for n n0(ε)) follows

    (not >).

    p. 69 in (2.40) should have |S

    p. 71 at the end of (2.44) bracket is missing: should be ... + 2)).

    p. 74 in (2.56) should be (absolute value signs missing) |g(r)(x)|

    p. 77 1. -12 should be r .

    p. 80 1. 1 should have u > 1 (not <).

    p. 87 1. 5 should be xa and not x–2a

    p. 88 1. 17 should be: obtains, for N x,

    p. 89 1. 15 should be convergent and not covergent

    and the next line by

    ,where Aq, Bq ans Cq are uniformly bounded.

    p. 91 1. -8 replace a, b > 0, by 0 < a < b,

    (replace + by =)

    p. 121 1. -8 replace than by then

    p. 134 one absolute value sign missing in (5.17), should be |F’(t)|.

    p. 140 1. -11 should be which at and not which as.

    p. 172 Γ(w) missing in (7.3), should be Γ(w)dw =

    p. 175 1. 1 should be N < N’ ≤ 2N

    p. 226 1. 12 should be 14.6 not 13.6

    ≤ σ ≤ 2

    p. 239 1. -1 ...p−σ)² should be ...p

    p. 257 1. -9 should have du U exp

    p. 263 in (10.40) should have (2x)⁴/, four lines below should be M >> T¹/⁴

    p. 265 1. 5 should have ξ(m) instead of ζ(m)

    p. 267 1. -6 should have T¹/(²k+6) replaced by T¹/(⁶k+6)

    p. 270 (11.7) should have

    and then all as before.

    p. 294 1. 7 should have: If we now use R

    p. 295 1. -1 and 1. -2, S should be S1

    p. 334 1. -2 should be: holds, then we obtain

    p. 334 1.-1 should be: Δ(y) = −yU−1

    (

    p. 350 1. 9 should be: A < 1 not A > 1

    p. 352 in (13.9) delete twice +ε

    p. 363 1. 1 should be the first sum, 1. 10 (13.27) should be (13.26)

    p. 383 1. -13 should have : Hafner (1981a) and not only Hafner (1981)

    p. 386 1. -11 replace γ0 < 1 by–1 < γ0 < 1

    p. 387, 389, 391 in heading on top of the page Summary should be Summatory

    p. 419 1. 1 in the middle expression it should be eit(a(n)−k) not eit(a(n)k)

    p. 432 1. -10 should be: g(n, z)n−1

    p. 437 1. -18 Pintz (in press) should be Pintz (1985)

    p. 438 1. -13 (1982) should be (1972)

    p. 440 1. -4 Renyi’s problem should be Rényi’s problem

    p. 445 1.-12 |s|v should be |s−v|

    p. 448 1. -5 should be e(y)dy

    p. 449 1. 10 should be − logΓ(1 − u) + logΓ(u), 1. -6 (A.34) should be (A.33)

    (factor 2 is missing)

    p. 487 1. -7 Γ(a + b + 1) should be Γ(a + b)

    p. 492 (A.35) should have −1/(2s)

    p. 497 ref. R.C. Baker replace (in press) by: 46(1985), 73-79.

    p. 498 ref B.C. Berndt and R.J. Evans (1983) (in press) should be 92(1983), 67-96

    p. 499 ref. P. Deligne replace 53 by 43, (1975) by (1974)

    p. 500 ref. S.W. Graham replace (in press) by: Austin, 1985, pp. 96-126.

    p. 502 1. 7 replace on von by of von

    p. 503 first ref. Karacuba replace 119 by 112 last ref. Kolesnik replace (in press) by:

    45(1985), 115-143.

    p. 508 delete the ref. D. Suryanarayana which is twice printed, namely (1973b)

    CHAPTER ONE

    ELEMENTARY THEORY

    1.1 DEFINITION OF ζ(s) AND ELEMENTARY PROPERTIES

    The Riemann zeta-function ζ(s) is defined as

    (1.1)

    , and the series in (1.1) converges absolutely and uniformly in the half-plane Re s ≥1 + ε, since |ns| ≤ nconverges. Here, ns for complex s is defined as es log n, where log n is the natural logarithm of n. The zeta-function seems to have been studied first by L. Euler (1707–1783), who considered only real values of s. The present notation and the notion of ζ(s) as a function of the complex variable s are due to B. Riemann (1826–1866), who made a number of startling discoveries about ζ(s). Riemann wrote s = σ + it (σ, t real) for the complex variable s, and this tradition still persists, although some authors prefer the more logical notation s = σ. + Uniform convergence for σ = Re s ≥ 1 + ε of the series in (1.1) implies by a well-known theorem of Weierstrass that ζ(s) is regular for σ > 1, and its derivatives in this region may be obtained by termwise differentiation of the series in (1.1).

    Historically, the zeta-function arose from the need for an analytic tool capable of dealing with problems involving prime numbers. It was observed already by Euler that

    (1.2)

    for real s > 1, where the product is taken over all primes. The identity (1.2) plays a fundamental role in the analytic theory of primes and it in fact holds for Res > 1, being a corollary of the following:

    THEOREM 1.1. Let f(n) be a real or complex-valued multiplicative function [i.e., f(mn) = f(m)f(n) if (m, n. Then

    (1.3)

    Proof of Theorem 1.1. The product in (1.3) is called an Euler product, and each of its factors is an absolutely convergent series. Hence we may multiply a finite number of these series to obtain

    denotes the summation over all n having no prime factor larger than x. Since

    where

    we obtain

    converges. Taking f(n) = n−s in (1.3) we obtain (1.2) for σ > 1.

    From (1.2) we have, for σ > 1,

    pp−σ converges for σ > 1. This shows that ζ(s) ≠ 0 for σ > 1. The regions in the complex plane where ζ(s) ≠ 0 are known as zero-free regions for ζ(s). A little later in Theorem 1.5 we shall see that σ ≥ 1 is a zero-free region for ζ(s), and it is of great importance to extend the zero-free region to the left of σ = 1 as much as possible. This problem will be reconsidered in Chapter 6, where we shall obtain the best known zero-free region for ζ(s).

    Euler’s identity (1.2) can be recast into another, often more convenient form —that is, take the logarithm of both sides in (1.2) and differentiate to obtain

    (1.4)

    for s p(log p)/(p¹+ε − 1) for σ ≥ 1 + ε and hence it converges uniformly for σ ≥ 1 + e. Changing the order of summation in (1.4) we obtain, for s = σ > 1,

    (1.5)

    where A(n) is known as the von Mangoldt function, and may be defined as

    (1.6)

    However, the series in (1.5) converges uniformly and absolutely for Re s = σ ≥ 1 + ε; thus by analytic continuation (1.5) holds for Re s = σ > 1.

    The importance of the zeta-function comes from the fact that the variable s may possess an analytic continuation outside the region σ > 1, where the series converges. In fact in most applications of zeta-function theory, it is precisely the information about ζ(s) when σ ≤ 1 which is of crucial importance. The simplest facts about the analytic character of ζ(s) are contained in

    THEOREM 1.2. The function ζ(s), defined by (1.1) for σ > 1, admits of analytic continuation over the whole complex plane having as its only singularity a simple pole with residue 1 at s = 1.

    Proof of Theorem 1.2. For x > 1, we have

    If σ > 1 and x → ∞, it follows that

    (1.7)

    Since |[t] − t| ≤ 1 it is seen that the integral in (1.7) is uniformly convergent for σ δ and any fixed δ > 0. Thus this integral represents an analytic function for σ > 0, and (1.7) provides the analytic continuation of ζ(s) over the half-plane σ > 0, its only pole for σ > 0 being s = 1 with residue 1. More generally, the Euler-Maclaurin summation formula [see (A.24)] gives, for a > 1 and n ≥ 1 fixed,

    (1.8)

    Using the fact that P2n+1(t) = O(1) we obtain by analytic continuation that (1.8) holds for σ + 2n > 0. Since n is arbitrary, the conclusion of the theorem follows.

    Another useful representation of ζ(s) may be obtained if one observes that, for σ > 1,

    (1.9)

    so that

    (1.10)

    since the alternating series in (1.9) converges for σ > 0. Thus (1.10) also provides the analytic continuation of ζ(s) for σ > 0, and shows in particular that ζ(σ) < 0 for 0 < σ < 1.

    Theorem 1.2 shows that the Laurent expansion of ζ(s) in the neighborhood of its pole s = 1 is

    (1.11)

    It is possible to evaluate explicitly the coefficients γk in (1.11), as shown by

    THEOREM 1.3. If γ0, γ1, γ2, . . . are defined by (1.11), then

    (1.12)

    and, in particular,

    (1.13)

    is Euler’s constant.

    Proof of Theorem 1.3. Let k, r ≥ 0 be integers and let

    By the Stieltjes integral representation we have

    Further let Sr(xn≤xn−1(log(x/n))r. Then

    say. We want to prove that γk = (−1)kck/k! for 0 ≤ k r. Observe that

    since the γk’s are defined by (1.11). On the other hand, using the expression for Sr(x) it is seen that

    where for 0 ≤ k ≤ r we have

    , we obtain γk = (−1)kck/k! for 0 ≤ k r, and since r may be arbitrary (1.12) follows for k = 0, 1 , 2, . . . .

    A formula such as (1.8) may be used for numerical evaluation of ζ(s), but for s = 2k already Euler established a simple, explicit formula which we state here as

    THEOREM 1.4. If k = 1,2, ... and Bj denotes the j th Bernoulli number, then

    (1.14)

    Proof of Theorem 1.4. into the well-known identity

    one obtains by logarithmic differentiation

    (1.15)

    Recall that the Bernoulli numbers Bk are defined by

    so that B0 , and B2k+1 = 0 for k: ≥ 1. Hence, for |u| < 2π, (1.15) gives

    and equating coefficients of u²k we obtain (1.14). In particular,

    (1.16)

    No one has yet succeeded in obtaining a formula as simple as (1.14) for ζ(2k + 1). In fact, besides the result of R. Apéry that ζ(3) is irrational, almost nothing is known about the arithmetical structure of ζ(2k + 1) for k > 1.

    We shall now use Theorem 1.2 to prove the classical result of J. Hadamard and C. J. de la Vallée-Poussin that ζ(s) does not vanish on the line σ = 1. This is

    THEOREM 1.5. For t real, ζ(1 + it) ≠ 0.

    Proof of Theorem 1.5. The proof is by contradiction. Since s = 1 is a pole of ζ(s), assume that 1 + it1 (t1 ≠ 0) is a zero of order m. Then 1 + it1 is a first-order pole with residue m ≥ 1 for the function ζ’(s)/ζ(s). Thus for σ > 1, but sufficiently close to 1, we have

    (1.17)

    Using Theorem 1.2 we have

    (1.18)

    and also

    (1.19)

    In (1.19) the left-hand side is bounded if ζ(1 + 2it1) ≠ 0, hence in that case we may set k = 0, while if ζ(1 + 2it1) = 0 we have k ≥ 1. In view of Λ(n) ≥ 0, (1.5) gives

    (1.20)

    But from (1.17)–(1.19) we see that

    (1.21)

    if σ is sufficiently close to 1, since 4m ≥ 4 and k ≥ 0. Thus (1.21) contradicts (1.20) and proves the theorem.

    1.2 THE FUNCTIONAL EQUATION

    The functional equation for the Riemann zeta-function is given by

    THEOREM 1.6. For all complex s

    (1.22)

    This is the symmetric form of the functional equation, which represents one of the fundamental results of zeta-function theory. It was discovered and proved by B. Riemann, and represents one of his most remarkable achievements. The functional equation shows that ζ(− 2n) = 0 for n = 1, 2, . . . , since the gamma-function has poles at nonpositive integers and for s = 0 the pole of ζ(1 − s) cancels the pole of Γ(s/2), and in fact dividing (1.22) by Γ(s/2) and letting s . The zeros s = −2n are often called the trivial zeros of ζ(s), since it will be shown a little later in Theorem 1.7 that ζ(s, and some results concerning their distribution will be proved.

    There are several equivalent ways in which the functional equation (1.22) may be written. Using standard properties of the gamma-function (see Appendix) one may write (1.22) as

    (1.23)

    or equivalently as

    (1.24)

    From Stirling’s formula (A.34) it follows that

    (1.25)

    which is for most purposes a sufficiently sharp approximation. The condition t for σ > 1, and from (1.7) and (1.22) this property holds for other values of s too.

    Proof of Theorem 1.6. There exist in the literature many proofs of the functional equation, and here we shall present two of them. The first is classical, while the second one is partially new and almost elementary. The first proof of (1.22) starts from (1.7) for 0 < σ < 1, which may be rewritten as

    (1.26)

    Actually, it turns out that (1.26) provides the analytic continuation of ζ(s) for σ , and observe that P(xdx = 0. Hence

    If σ > −1, then this expression tends to zero as A, B → ∞, proving that the integral in (1.26) is bounded; hence (1.26) holds for σ > −1. Further, for σ < 0, we have

    hence

    (1.27)

    (x not an integer), we have, by termwise integration of (1.27),

    (1.28)

    This holds for −1 < σ < 0 and reduces easily to (1.24), and by analytic continuation it holds for other values of s too. To justify term-by-term integration observe that the series for ψ(x) is boundedly convergent; hence term-by-term integration in any finite interval is permissible. Thus it will be sufficient to show

    (1.29)

    for −1 < σ < 0. An integration by parts gives

    whence (1.29) easily follows.

    Our second proof of the functional equation starts from the elementary identity

    Logarithmic differentiation gives

    or

    Letting n → ∞ we obtain the identity

    (1.30)

    which is the starting point for the proof of the functional equation. Consider now, for 0 < σ < 1,

    gives now, for 0 < σ < 1,

    (1.31)

    where

    (1.32)

    If we now use the fact that f(x) is self-reciprocal with respect to sine transforms, that is,

    (1.33)

    then (1.22) easily follows from (1.31). Namely, we have

    (1.34)

    and the last equality also holds by analytic continuation outside the strip 0 < σ < 1. Here the inversion of the order of integration is justified, since it is readily seen that

    Using the first identity in (A.30), we obtain then from (1.34)

    which is exactly (1.23). Finally, to see that (1.33) holds we use (1.15) and (1.32) to obtain

    1.3 THE HADAMARD PRODUCT FORMULA

    From the discussion made so far it follows that the only zeros of ζ(s) lying outside the strip 0 < σ < 1 are the so-called trivial zeros s = − 2n (n = 1, 2, . . .). The strip 0 < σ the critical line in zeta-function theory. To show that ζ(s) has indeed many zeros in the critical strip we must first prepare the groundwork by developing a suitable product formula, due to J. Hadamard, from complex function theory. We begin with

    LEMMA 1.1. (Jensen’s formula) Let f(s) be a function of the complex variable s = reiθ (r, θ real) which is regular in |s| ≤ R with no zeros on |s| = R and which satisfies f(0) = 1. Then

    (1.35)

    where n(r) is the number of zeros of f(s) inside the circle |s| = r.

    Proof of Lemma 1.1. Since Re logz = log|z|, the left-hand side of (1.35) is

    If s1, s2, . . . . , sn are the zeros of f(s) in |s| < R and |s1| = r1, . . . , |sn| = rn, f(0) = 1, then

    which is an alternative way of writing Jensen’s formula.

    A function f(s), regular over the whole complex plane is called an integral function of finite order is

    for some finite constant A as s → ∞. The order of f(s) is the lower bound of those A for which the above inequality holds. The study of integral functions of finite order was developed at the end of 19th century by J. Hadamard, who showed that these functions can be written as an infinite product containing factors of the form s s0 corresponding to the zero s0 of the function in question. The integral function useful in zeta-function theory is

    (1.36)

    which is an integral function of order 1. To see this, note that from (1.7) we obtain

    , and so by Stirling’s formula (A.34)

    (1.37)

    . Using the functional equation (1.22) we have ξ(1 − s) = ξ(stoo. Further, by Stirling’s formula for real s , which shows that the order of ζ(s) is exactly 1.

    An immediate consequence of Jensen’s formula (1.35) and (1.37) is the bound

    (1.38)

    where N(T) is the number of zeros of ζ(s) in the region 0 < σ < 1,0 < t T. From the definition of ξ(s) in (1.36) it follows that the zeros of ξ(s) are the nontrivial zeros of ζ(s), since in (1.36) the trivial zeros s = −2n (n = 1, 2, . . .) of ζ(s) are canceled by the poles of Γ(s/2), while sΓ(s/2) has no zeros and the zero of s − 1 cancels the pole of ζ(s). Taking R = 4T, T T0, we have

    on using (1.35) and (1.37). The bound T log T is actually the correct order of magnitude for N(T), as will be shown by the Riemann–von Mangoldt asymptotic formula (1.44).

    Let f(s) from now on denote an integral function of order 1, where we have in mind the eventual application to f(s) = ξ(s). Observe first that using (1.35) we have

    (1.39)

    for any ε > 0 if f(s) is an integral function of order 1. If r1, r2, . . . are the moduli of zeros ρ1, ρ2, ... of f(s), then using (1.39) we have

    for any fixed ε > 0. Hence the product

    is either finite or it converges absolutely for all s, and therefore represents an integral function with zeros (of the appropriate multiplicities) at ρ1, ρ2, .... If we set

    then F(s) is an integral function without zeros, and we wish to show that F(s) = eA+Bs with some suitable constants A and B. To achieve this we shall consider g(s) = logF(s) and show that g(s) is a first-degree polynomial in s. We shall need the fact that F(s) is an integral function of order at most 1, so that for |s| = R,

    and g(s) can be defined to be single valued [since F(s) has no zeros], and consequently it is also an integral function. If we write

    then without loss of generality we may assume g(0) = 0, so that if s = Reiθ, then

    which is a Fourier series in θ. Thus

    hence

    for n ≥ 2 if R is large enough, and a similar argument shows that bn = 0 for n ≥ 2. Therefore g(s) = A + Bs, and it remains to show that F(s) has order at most 1, which will follow from

    (1.40)

    converges, so that there exist arbitrarily large values of R for which

    for all n. Let

    say, where in P1, |ρn| < R/2; in P2, R/2 ≤ |ρn | ≤ 2R; and in P3, |ρn| > 2R. For the factors of P1, we have, on |s| = R,

    so that using (1.39) we obtain

    For the factors of P2,

    where C is a positive constant. Hence by (1.39)

    while finally for P3,

    .Therefore

    converges. Hence

    and (1.40) follows. The net result is that

    (1.41)

    for suitable constants A, B if f(s) is an integral function of order 1 with zeros ρ1, ρ2, . . . .

    We shall now apply the product formula (1.41) to the function ξ(s|rn|−1 diverges if r1, r2 ... are the moduli of the zeros ρ1, ρ2, . . . of ξ(s), proving incidentally that ζ(s) has an infinity of complex (i.e., nontrivial) zeros, which must then all lie in the critical strip 0 < σ < 1. Logarithmic differentiation of (1.41) with f = ξ gives

    (1.42)

    which combined with (1.36) gives

    (1.43)

    where the summation is over all nontrivial zeros of ζ(s). Letting s → 0 in (1.43) we obtain

    Now we have Γ’(1) = —γ, Γ(1) = 1, while using the functional equation for ζ(s) and logarithmic differentiation we find that ζ’(0)/ζ(0) = log 2π, which gives

    1.4 THE RIEMANN—VON MANGOLDT FORMULA

    The formula in question was conjectured by B. Riemann in 1859 and proved by H. von Mangoldt more than 30 years later. It furnishes a precise expression for N(T), the number of zeros of ζ(s) in the region 0 < σ < 1, 0 < t T, and represents an important tool in zeta-function theory. We state the result here as

    THEOREM 1.7.

    (1.44)

    Proof of Theorem 1.7. be the rectangle with vertices 2 ± iT, − 1 ± iT. Then

    (1.45)

    where ξ(s) is defined by (1.36). With η(s) = π−s/²Γ(s/2)ζ(s) we therefore may write

    Observe first that

    while η(s) = η(1 − s) and η(σ ± it) are conjugates, so that

    consists of the segments [2, 2 + iT. Therefore

    Using Stirling’s formula (A.35) and

    we obtain

    Thus from the above estimates we have

    and to prove the theorem it remains to show that

    (1.46)

    is clearly bounded. To prove (1.46) we shall need some estimates involving ζ’/ζ. To begin we suppose that t ≥ 2, 1 ≤ σ ≤ 2, so that the gamma term in (1.43) is 0(log t) and, consequently,

    (1.47)

    In this formula we take now s = 2 + iT, we obtain

    (1.48)

    If ρ = β + is a nontrivial zero of ζ(s), then Re ρ−1 = β(β² + γ²)−1 > 0 and

    hence (1.48) gives

    (1.49)

    where the summation is over all nontrivial zeros ρ of ζ(s). The last bound gives immediately

    (1.50)

    that is, each strip T < t T + 1 contains fewer than C log T zeros of ζ(s) for some absolute C > 0. Using again (1.43) with s and 2 + it (where t > 2 is not the ordinate of any ρ) and subtracting, we obtain

    (1.51)

    Here for terms with |γ − t| ≥ 1 we have

    if −1 ≤ σ ≤ 2. Hence by (1.48) the portion of the sum in (1.51) for which |γ − t| ≥ 1 is O(log t), and for |γ − t| < 1 we have |2 + it ρ| ≥ 1, and the number of such ρ is O(log t) by (1.50). Thus we obtain

    (1.52)

    Now the proof of (1.46) easily follows, since by (1.52)

    since |Δ arg(s −ρand (1.50) holds.

    This finishes the proof of the Riemann—von Mangoldt formula (1.44). Some immediate corollaries are

    (1.53)

    and

    (1.54)

    where 0 < γ1 ≤ γ2 ≤ γ3 ≤ · · · are consecutive ordinates of nontrivial zeros ρ = β + of ζ(s). The bounds in (1.53) follow by partial summation from (1.44), while (1.54) follows from (1.44) and the obvious inequalities

    We conclude this section by remarking that (1.47) can be used to prove that ζ(σ + it) ≠ 0 for

    (1.55)

    where C > 0 is an absolute constant. To see this, take s = σ + it or s = σ + 2it, ρ = β + iγ, and 1 ≤ σ < 2. In each case the sum over ρ in (1.48) is clearly positive, hence (1.47) gives

    for some absolute A > 0. Moreover, we have

    since s = 1 is a simple pole of ζ(s). If we put the above inequalities in (1.20), then (1.55) follows after some simplification with σ = 1 + δ/log t, where δ is a suitable positive constant. The zero-free region (1.55) was the one that was independently obtained by J. Hadamard and C. J. de la Vallée-Poussin around 1900 in their search for a proof of the prime number theorem. (See Chapter 12 for the strongest known version of the prime number theorem.) The estimate (1.55) is primarily of historical interest now, since better zero-free regions are known at present (see Chapter 6), and for this reason (1.55) was not formulated as a theorem.

    1.5 AN APPROXIMATE FUNCTIONAL EQUATION

    Although very important, the functional equation (1.22) for ζ(s) has the shortcoming of not expressing ζ(s) explicitly, but only in terms of ζ(1 − s) and the gamma-function. Many formulas exist in the literature which express ζ(s) [or ζk (s)] as a number of finite sums involving the function ns. These formulas have been nicknamed approximate functional equations, and in many instances they lead to very precise results about ζ(s). Approximate functional equations for ζ(s) will be discussed extensively in Chapter 4. Our aim here is to prove only one of the simplest approximate functional equations for ζ(s) which we present as

    THEOREM 1.8. For 0 < σ0 ≤ σ ≤ 2, x ≥ |t|/π, s = σ + it,

    (1.56)

    where the O-constant depends only on σ0.

    For the proof of this theorem we shall need a simple result concerning exponential sums, which is the subject of Chapter 2. This is

    LEMMA 1.2. Let f(x) be a real-valued function on the interval [a, b], and let f’(x) be continuous and monotonic on [a, b] and |f’(x)| ≤ δ < 1. Then

    (1.57)

    where the O-constant is absolute.

    Proof of Lemma 1.2. We may suppose that f’(x) ≥ 0 on [a, b] by taking the conjugate of the sum in (1.57) if f’(x) < 0. Let ϕn(x) = e(f(n + x)) for 0 < x < 1 and let us extend the definition of ϕn(x) over the whole real line by making it periodic with period 1 and

    Then ϕn(x) may be expanded into a Fourier series of the form

    (1.58)

    and for k ≠ 0 an integration by parts yields

    Therefore setting x = 1 in (1.58) we obtain

    Summation over n gives then

    where

    Recall that the second mean value for real integrals asserts that there is a ξ in the interval [a, b] such that

    (1.59)

    providing that u, v are Riemann integrable on [a, b] and u is monotonic. Thus we apply (1.59) to the real and imaginary part of the integral for Rk to obtain

    Hence

    and the lemma follows.

    Proof of Theorem 1.8. We have, for Re s > 1 and N ≥ 2,

    Therefore

    (1.60)

    and by analytic continuation this is valid for σ (1 + |t|)Nσ. If u ≥ x we set

    and apply Lemma 1.2 with f(x) = ()−1|t|log x, provided that x ≥ |t|/π. From (1.57) we have

    For x N, partial summation gives

    Substituting this estimate in (1.60) we finally have, for σ0 ≤ σ ≤ 2,

    and letting N → ∞ (1.56) follows.

    We are now in a position to say something about the order of |ζ(s)|, which is one of the most important and difficult problems in zeta-function theory. From (1.56) we have, for 1 ≤ σ ≤ 2, uniformly in σ,

    (1.61)

    There is no loss of generality in supposing t > 0, since ζ(σ ± it) are conjugates. Furthermore, for σ , while using (1.23) and (1.25) we have

    (1.62)

    A useful way to compare the orders of |ζ(s)| at two different values of σ is given by the following:

    LEMMA 1.3. For 0 ≤ σ1 ≤ σ0 ≤ 3/2, t t0, we have uniformly in σ0

    (1.63)

    Proof of Lemma 1.3. be the rectangle with vertices σ1 + it ± i log²t, 3 + it ± i log²t. Then f(s) = ζ(s)Γ(s s0 + 2) (s0 = σ0 + it, and therefore by the maximum modulus principle

    we have |Im(s s0 + 2| = log²t, and trivially [e.g., from (1.56)] ζ(st²; hence the maximum of |f(s)| over these sides is o(1) by Stirling’s formula (A.34), while the maximum on the side with σ = 3 is clearly O(1). Also by Stirling’s formula we have

    To obtain an order result for 0 ≤ σ ≤ 1 we use ζ(1 + itlog t, ζ(itt¹/² log t and apply the Phragmén–Lindelöf principle (see Section A.8) to the function

    in the strip 0 ≤ σ ≤ 1. This function is regular in the strip and bounded for σ = 0 and σ = 1 by (1.61) and (1.62). Thus we obtain

    (1.64)

    Therefore it makes sense to define the function μ(σ) for each real σ as the infimum of number c ≥ 0 such that ζ(σ + ittc, or alternatively as

    (1.65)

    From (1.61) and (1.62) we have μ(σ) = 0 for σ for σ ≤ 0, but the exact value of μ(σ) for any 0 < σ ]. From Lemma 1.3 we see immediately that μ(σ) is nonincreasing,

    Enjoying the preview?
    Page 1 of 1