Zeta function universality

Last updated
Any non-vanishing holomorphic function f defined on the strip can be approximated by the z-function. Voronin universality theorem.png
Any non-vanishing holomorphic function f defined on the strip can be approximated by the ζ-function.

In mathematics, the universality of zeta functions is the remarkable ability of the Riemann zeta function and other similar functions (such as the Dirichlet L-functions) to approximate arbitrary non-vanishing holomorphic functions arbitrarily well.

Contents

The universality of the Riemann zeta function was first proven by Sergei Mikhailovitch Voronin  [ ru ] in 1975 [1] and is sometimes known as Voronin's universality theorem.

The Riemann zeta function on the strip 1/2 < Re(s) < 1; 103 < Im(s) < 109. Voronin-3.png
The Riemann zeta function on the strip 1/2 < Re(s) < 1; 103 < Im(s) < 109.

Formal statement

A mathematically precise statement of universality for the Riemann zeta function ζ(s) follows.

Let U be a compact subset of the strip

such that the complement of U is connected. Let f : UC be a continuous function on U which is holomorphic on the interior of U and does not have any zeros in U. Then for any ε> 0 there exists a t ≥ 0 such that

 

 

 

 

(1)

for all .

Even more: the lower density of the set of values t which do the job is positive, as is expressed by the following inequality about a limit inferior.

where λ denotes the Lebesgue measure on the real numbers.

Discussion

The condition that the complement of U be connected essentially means that U doesn't contain any holes.

The intuitive meaning of the first statement is as follows: it is possible to move U by some vertical displacement it so that the function f on U is approximated by the zeta function on the displaced copy of U, to an accuracy of ε.

The function f is not allowed to have any zeros on U. This is an important restriction; if we start with a holomorphic function with an isolated zero, then any "nearby" holomorphic function will also have a zero. According to the Riemann hypothesis, the Riemann zeta function does not have any zeros in the considered strip, and so it couldn't possibly approximate such a function. The function f(s) = 0 which is identically zero on U can be approximated by ζ: we can first pick the "nearby" function g(s) = ε/2 (which is holomorphic and doesn't have zeros) and find a vertical displacement such that ζ approximates g to accuracy ε/2, and therefore f to accuracy ε.

The accompanying figure shows the zeta function on a representative part of the relevant strip. The color of the point s encodes the value ζ(s) as follows: the hue represents the argument of ζ(s), with red denoting positive real values, and then counterclockwise through yellow, green cyan, blue and purple. Strong colors denote values close to 0 (black = 0), weak colors denote values far away from 0 (white = ∞). The picture shows three zeros of the zeta function, at about 1/2 + 103.7i, 1/2 + 105.5i and 1/2 + 107.2i. Voronin's theorem essentially states that this strip contains all possible "analytic" color patterns that don't use black or white.

The rough meaning of the statement on the lower density is as follows: if a function f and an ε > 0 is given, there is a positive probability that a randomly picked vertical displacement it will yield an approximation of f to accuracy ε.

The interior of U may be empty, in which case there is no requirement of f being holomorphic. For example, if we take U to be a line segment, then a continuous function f : UCis a curve in the complex plane, and we see that the zeta function encodes every possible curve (i.e., any figure that can be drawn without lifting the pencil) to arbitrary precision on the considered strip.

The theorem as stated applies only to regions U that are contained in the strip. However, if we allow translations and scalings, we can also find encoded in the zeta functions approximate versions of all non-vanishing holomorphic functions defined on other regions. In particular, since the zeta function itself is holomorphic, versions of itself are encoded within it at different scales, the hallmark of a fractal. [2]

The surprising nature of the theorem may be summarized in this way: the Riemann zeta function contains "all possible behaviors" within it, and is thus "chaotic" in a sense, yet it is a perfectly smooth analytic function with a straightforward definition.

Proof sketch

A sketch of the proof presented in (Voronin and Karatsuba, 1992) [3] follows. We consider only the case where U is a disk centered at 3/4:

and we will argue that every non-zero holomorphic function defined on U can be approximated by the ζ-function on a vertical translation of this set.

Passing to the logarithm, it is enough to show that for every holomorphic function g : UC and every ε > 0 there exists a real number t such that

We will first approximate g(s) with the logarithm of certain finite products reminiscent of the Euler product for the ζ-function:

where P denotes the set of all primes.

If is a sequence of real numbers, one for each prime p, and M is a finite set of primes, we set

We consider the specific sequence

and claim that g(s) can be approximated by a function of the form for a suitable set M of primes. The proof of this claim utilizes the Bergman space, falsely named Hardy space in (Voronin and Karatsuba, 1992), [3] in H of holomorphic functions defined on U, a Hilbert space. We set

where pk denotes the k-th prime number. It can then be shown that the series

is conditionally convergent in H, i.e. for every element v of H there exists a rearrangement of the series which converges in H to v. This argument uses a theorem that generalizes the Riemann series theorem to a Hilbert space setting. Because of a relationship between the norm in H and the maximum absolute value of a function, we can then approximate our given function g(s) with an initial segment of this rearranged series, as required.

By a version of the Kronecker theorem, applied to the real numbers (which are linearly independent over the rationals) we can find real values of t so that is approximated by . Further, for some of these values t, approximates , finishing the proof.

The theorem is stated without proof in § 11.11 of (Titchmarsh and Heath-Brown, 1986), [4] the second edition of a 1951 monograph by Titchmarsh; and a weaker result is given in Thm. 11.9. Although Voronin's theorem is not proved there, two corollaries are derived from it:

1) Let    be fixed. Then the curve
is dense in
2) Let    be any continuous function, and let    be real constants.
Then cannot satisfy the differential-difference equation
unless    vanishes identically.

Effective universality

Some recent work has focused on effective universality. Under the conditions stated at the beginning of this article, there exist values of t that satisfy inequality (1). An effective universality theorem places an upper bound on the smallest such t.

For example, in 2003, Garunkštis proved that if is analytic in with , then for any ε in , there exists a number in such that

.

For example, if , then the bound for t is .

Bounds can also be obtained on the measure of these t values, in terms of ε:

.

For example, if , then the right-hand side is . See. [5] :p. 210

Universality of other zeta functions

Work has been done showing that universality extends to Selberg zeta functions. [6]

The Dirichlet L-functions show not only universality, but a certain kind of joint universality that allow any set of functions to be approximated by the same value(s) of t in different L-functions, where each function to be approximated is paired with a different L-function. [7] [8] :Section 4

A similar universality property has been shown for the Lerch zeta function , at least when the parameter α is a transcendental number. [8] :Section 5 Sections of the Lerch zeta function have also been shown to have a form of joint universality. [8] :Section 6

Related Research Articles

In mathematics, the prime number theorem (PNT) describes the asymptotic distribution of the prime numbers among the positive integers. It formalizes the intuitive idea that primes become less common as they become larger by precisely quantifying the rate at which this occurs. The theorem was proved independently by Jacques Hadamard and Charles Jean de la Vallée Poussin in 1896 using ideas introduced by Bernhard Riemann.

<span class="mw-page-title-main">Riemann zeta function</span> Analytic function in mathematics

The Riemann zeta function or Euler–Riemann zeta function, denoted by the Greek letter ζ (zeta), is a mathematical function of a complex variable defined as

<span class="mw-page-title-main">Cauchy's integral formula</span> Provides integral formulas for all derivatives of a holomorphic function

In mathematics, Cauchy's integral formula, named after Augustin-Louis Cauchy, is a central statement in complex analysis. It expresses the fact that a holomorphic function defined on a disk is completely determined by its values on the boundary of the disk, and it provides integral formulas for all derivatives of a holomorphic function. Cauchy's formula shows that, in complex analysis, "differentiation is equivalent to integration": complex differentiation, like integration, behaves well under uniform limits – a result that does not hold in real analysis.

The Liouville Lambda function, denoted by λ(n) and named after Joseph Liouville, is an important arithmetic function. Its value is +1 if n is the product of an even number of prime numbers, and −1 if it is the product of an odd number of primes.

<span class="mw-page-title-main">Analytic number theory</span> Exploring properties of the integers with complex analysis

In mathematics, analytic number theory is a branch of number theory that uses methods from mathematical analysis to solve problems about the integers. It is often said to have begun with Peter Gustav Lejeune Dirichlet's 1837 introduction of Dirichlet L-functions to give the first proof of Dirichlet's theorem on arithmetic progressions. It is well known for its results on prime numbers and additive number theory.

In complex analysis, Liouville's theorem, named after Joseph Liouville, states that every bounded entire function must be constant. That is, every holomorphic function for which there exists a positive number such that for all in is constant. Equivalently, non-constant holomorphic functions on have unbounded images.

The theory of functions of several complex variables is the branch of mathematics dealing with complex-valued functions. The name of the field dealing with the properties of function of several complex variables is called several complex variables, that has become a common name for that whole field of study and Mathematics Subject Classification has, as a top-level heading. A function is n-tuples of complex numbers, classically studied on the complex coordinate space .

<span class="mw-page-title-main">Mertens function</span>

In number theory, the Mertens function is defined for all positive integers n as

<span class="mw-page-title-main">Integral test for convergence</span> Test for infinite series of monotonous terms for convergence

In mathematics, the integral test for convergence is a method used to test infinite series of monotonous terms for convergence. It was developed by Colin Maclaurin and Augustin-Louis Cauchy and is sometimes known as the Maclaurin–Cauchy test.

In mathematics, a Paley–Wiener theorem is any theorem that relates decay properties of a function or distribution at infinity with analyticity of its Fourier transform. The theorem is named for Raymond Paley (1907–1933) and Norbert Wiener (1894–1964). The original theorems did not use the language of distributions, and instead applied to square-integrable functions. The first such theorem using distributions was due to Laurent Schwartz. These theorems heavily rely on the triangle inequality.

In complex analysis, the Phragmén–Lindelöf principle, first formulated by Lars Edvard Phragmén (1863–1937) and Ernst Leonard Lindelöf (1870–1946) in 1908, is a technique which employs an auxiliary, parameterized function to prove the boundedness of a holomorphic function on an unbounded domain when an additional condition constraining the growth of on is given. It is a generalization of the maximum modulus principle, which is only applicable to bounded domains.

<span class="mw-page-title-main">Z function</span>

In mathematics, the Z function is a function used for studying the Riemann zeta function along the critical line where the argument is one-half. It is also called the Riemann–Siegel Z function, the Riemann–Siegel zeta function, the Hardy function, the Hardy Z function and the Hardy zeta function. It can be defined in terms of the Riemann–Siegel theta function and the Riemann zeta function by

In mathematics, the explicit formulae for L-functions are relations between sums over the complex number zeroes of an L-function and sums over prime powers, introduced by Riemann (1859) for the Riemann zeta function. Such explicit formulae have been applied also to questions on bounding the discriminant of an algebraic number field, and the conductor of a number field.

<span class="mw-page-title-main">Chebyshev function</span>

In mathematics, the Chebyshev function is either a scalarising function or one of two related functions. The first Chebyshev functionϑ(x) or θ(x) is given by

The Sokhotski–Plemelj theorem is a theorem in complex analysis, which helps in evaluating certain integrals. The real-line version of it is often used in physics, although rarely referred to by name. The theorem is named after Julian Sochocki, who proved it in 1868, and Josip Plemelj, who rediscovered it as a main ingredient of his solution of the Riemann–Hilbert problem in 1908.

<span class="mw-page-title-main">Riemann hypothesis</span> Conjecture on zeros of the zeta function

In mathematics, the Riemann hypothesis is the conjecture that the Riemann zeta function has its zeros only at the negative even integers and complex numbers with real part 1/2. Many consider it to be the most important unsolved problem in pure mathematics. It is of great interest in number theory because it implies results about the distribution of prime numbers. It was proposed by Bernhard Riemann (1859), after whom it is named.

<span class="mw-page-title-main">Anatoly Karatsuba</span> Russian mathematician

Anatoly Alexeyevich Karatsuba was a Russian mathematician working in the field of analytic number theory, p-adic numbers and Dirichlet series.

<span class="mw-page-title-main">Riemann sphere</span> Model of the extended complex plane plus a point at infinity

In mathematics, the Riemann sphere, named after Bernhard Riemann, is a model of the extended complex plane: the complex plane plus one point at infinity. This extended plane represents the extended complex numbers, that is, the complex numbers plus a value for infinity. With the Riemann model, the point is near to very large numbers, just as the point is near to very small numbers.

<span class="mw-page-title-main">Grunsky matrix</span>

In complex analysis and geometric function theory, the Grunsky matrices, or Grunsky operators, are infinite matrices introduced in 1939 by Helmut Grunsky. The matrices correspond to either a single holomorphic function on the unit disk or a pair of holomorphic functions on the unit disk and its complement. The Grunsky inequalities express boundedness properties of these matrices, which in general are contraction operators or in important special cases unitary operators. As Grunsky showed, these inequalities hold if and only if the holomorphic function is univalent. The inequalities are equivalent to the inequalities of Goluzin, discovered in 1947. Roughly speaking, the Grunsky inequalities give information on the coefficients of the logarithm of a univalent function; later generalizations by Milin, starting from the Lebedev–Milin inequality, succeeded in exponentiating the inequalities to obtain inequalities for the coefficients of the univalent function itself. The Grunsky matrix and its associated inequalities were originally formulated in a more general setting of univalent functions between a region bounded by finitely many sufficiently smooth Jordan curves and its complement: the results of Grunsky, Goluzin and Milin generalize to that case.

In mathematical physics, the Wu–Sprung potential, named after Hua Wu and Donald Sprung, is a potential function in one dimension inside a Hamiltonian with the potential defined by solving a non-linear integral equation defined by the Bohr–Sommerfeld quantization conditions involving the spectral staircase, the energies and the potential .

References

  1. Voronin, S.M. (1975) "Theorem on the Universality of the Riemann Zeta Function." Izv. Akad. Nauk SSSR, Ser. Matem. 39 pp.475-486. Reprinted in Math. USSR Izv. 9, 443-445, 1975
  2. Woon, S.C. (1994-06-11). "Riemann zeta function is a fractal". arXiv: chao-dyn/9406003 .
  3. 1 2 Karatsuba, A. A.; Voronin, S. M. (July 1992). The Riemann Zeta-Function . Walter de Gruyter. p.  396. ISBN   3-11-013170-6.
  4. Titchmarsh, Edward Charles; Heath-Brown, David Rodney ("Roger") (1986). The Theory of the Riemann Zeta-function (2nd ed.). Oxford: Oxford U. P. pp. 308–309. ISBN   0-19-853369-1.
  5. Ramūnas Garunkštis; Antanas Laurinčikas; Kohji Matsumoto; Jörn Steuding; Rasa Steuding (2010). "Effective uniform approximation by the Riemann zeta-function". Publicacions Matemàtiques. 54 (1): 209–219. doi:10.5565/publmat_54110_12. JSTOR   43736941.
  6. Paulius Drungilas; Ramūnas Garunkštis; Audrius Kačėnas (2013). "Universality of the Selberg zeta-function for the modular group". Forum Mathematicum. 25 (3). doi:10.1515/form.2011.127. ISSN   1435-5337. S2CID   54965707.
  7. B. Bagchi (1982). "A Universality theorem for Dirichlet L-functions". Mathematische Zeitschrift. 181 (3): 319–334. doi:10.1007/BF01161980. S2CID   120930513.
  8. 1 2 3 Kohji Matsumoto (2013). "A survey on the theory of universality for zeta and L-functions". Plowing and Starring Through High Wave Forms. Proceedings of the 7th China–Japan Seminar. The 7th China–Japan Seminar on Number Theory. Vol. 11. Fukuoka, Japan: World Scientific. pp. 95–144. arXiv: 1407.4216 . Bibcode:2014arXiv1407.4216M. ISBN   978-981-4644-92-1.

Further reading