Average order of an arithmetic function

Last updated

In number theory, an average order of an arithmetic function is some simpler or better-understood function which takes the same values "on average".

Contents

Let be an arithmetic function. We say that an average order of is if

as tends to infinity.

It is conventional to choose an approximating function that is continuous and monotone. But even so an average order is of course not unique.

In cases where the limit

exists, it is said that has a mean value (average value) .

Examples

Calculating mean values using Dirichlet series

In case is of the form

for some arithmetic function , one has,

Generalizations of the previous identity are found here. This identity often provides a practical way to calculate the mean value in terms of the Riemann zeta function. This is illustrated in the following example.

The density of the k-th power free integers in N

For an integer the set of k-th-power-free integers is

We calculate the natural density of these numbers in N, that is, the average value of , denoted by , in terms of the zeta function.

The function is multiplicative, and since it is bounded by 1, its Dirichlet series converges absolutely in the half-plane , and there has Euler product

By the Möbius inversion formula, we get

where stands for the Möbius function. Equivalently,

where

and hence,

By comparing the coefficients, we get

Using (1), we get

We conclude that,

where for this we used the relation

which follows from the Möbius inversion formula.

In particular, the density of the square-free integers is .

Visibility of lattice points

We say that two lattice points are visible from one another if there is no lattice point on the open line segment joining them.

Now, if gcd(a, b) = d > 1, then writing a = da2, b = db2 one observes that the point (a2, b2) is on the line segment which joins (0,0) to (a, b) and hence (a, b) is not visible from the origin. Thus (a, b) is visible from the origin implies that (a, b) = 1. Conversely, it is also easy to see that gcd(a, b) = 1 implies that there is no other integer lattice point in the segment joining (0,0) to (a,b). Thus, (a, b) is visible from (0,0) if and only if gcd(a, b) = 1.

Notice that is the probability of a random point on the square to be visible from the origin.

Thus, one can show that the natural density of the points which are visible from the origin is given by the average,

is also the natural density of the square-free numbers in N. In fact, this is not a coincidence. Consider the k-dimensional lattice, . The natural density of the points which are visible from the origin is , which is also the natural density of the k-th free integers in N.

Divisor functions

Consider the generalization of :

The following are true:

where .

Better average order

This notion is best discussed through an example. From

( is the Euler–Mascheroni constant) and

we have the asymptotic relation

which suggests that the function is a better choice of average order for than simply .

Mean values over Fq[x]

Definition

Let h(x) be a function on the set of monic polynomials over Fq. For we define

This is the mean value (average value) of h on the set of monic polynomials of degree n. We say that g(n) is an average order of h if

as n tends to infinity.

In cases where the limit,

exists, it is said that h has a mean value (average value) c.

Zeta function and Dirichlet series in Fq[X]

Let Fq[X]=A be the ring of polynomials over the finite field Fq.

Let h be a polynomial arithmetic function (i.e. a function on set of monic polynomials over A). Its corresponding Dirichlet series define to be

where for , set if , and otherwise.

The polynomial zeta function is then

Similar to the situation in N, every Dirichlet series of a multiplicative function h has a product representation (Euler product):

Where the product runs over all monic irreducible polynomials P.

For example, the product representation of the zeta function is as for the integers: .

Unlike the classical zeta function, is a simple rational function:

In a similar way, If ƒ and g are two polynomial arithmetic functions, one defines ƒ * g, the Dirichlet convolution of ƒ and g, by

where the sum extends over all monic divisors d of m, or equivalently over all pairs (a, b) of monic polynomials whose product is m. The identity still holds. Thus, like in the elementary theory, the polynomial Dirichlet series and the zeta function has a connection with the notion of mean values in the context of polynomials. The following examples illustrate it.

Examples

The density of the k-th power free polynomials in Fq[X]

Define to be 1 if is k-th power free and 0 otherwise.

We calculate the average value of , which is the density of the k-th power free polynomials in Fq[X], in the same fashion as in the integers.

By multiplicativity of :

Denote the number of k-th power monic polynomials of degree n, we get

Making the substitution we get:

Finally, expand the left-hand side in a geometric series and compare the coefficients on on both sides, to conclude that

Hence,

And since it doesn't depend on n this is also the mean value of .

Polynomial Divisor functions

In Fq[X], we define

We will compute for .

First, notice that

where and .

Therefore,

Substitute we get,

, and by Cauchy product we get,

Finally we get that,

Notice that

Thus, if we set then the above result reads

which resembles the analogous result for the integers:

Number of divisors

Let be the number of monic divisors of f and let be the sum of over all monics of degree n.

where .

Expanding the right-hand side into power series we get,

Substitute the above equation becomes:

which resembles closely the analogous result for integers , where is Euler constant.

Not much is known about the error term for the integers, while in the polynomials case, there is no error term! This is because of the very simple nature of the zeta function , and that it has NO zeros.

Polynomial von Mangoldt function

The Polynomial von Mangoldt function is defined by:

Where the logarithm is taken on the basis of q.

Proposition. The mean value of is exactly 1.

Proof. Let m be a monic polynomial, and let be the prime decomposition of m.

We have,

Hence,

and we get that,

Now,

Thus,

We got that:

Now,

Hence,

and by dividing by we get that,

Polynomial Euler totient function

Define Euler totient function polynomial analogue, , to be the number of elements in the group . We have,

See also

Related Research Articles

In number theory, a multiplicative function is an arithmetic function f(n) of a positive integer n with the property that f(1) = 1 and whenever a and b are coprime, then

In mathematics, the classic Möbius inversion formula was introduced into number theory in 1832 by August Ferdinand Möbius.

In number theory, 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.

Riemann zeta function Analytic function

The Riemann zeta function or Euler–Riemann zeta function, ζ(s), is a function of a complex variable s that analytically continues the sum of the Dirichlet series

In mathematics, a square-free integer (or squarefree integer) is an integer which is divisible by no perfect square other than 1. That is, its prime factorization has exactly one factor for each prime that appears in it. For example, 10 = 2 ⋅ 5 is square-free, but 18 = 2 ⋅ 3 ⋅ 3 is not, because 18 is divisible by 9 = 32. The smallest positive square-free numbers are

The fundamental theorem of algebra states that every non-constant single-variable polynomial with complex coefficients has at least one complex root. This includes polynomials with real coefficients, since every real number is a complex number with its imaginary part equal to zero.

The Liouville Lambda function, denoted by λ(n) and named after Joseph Liouville, is an important arithmetic function.

Zeta distribution probability distribution on the integers in which the probability of a number is inversely proportion to a fixed power of the number

In probability theory and statistics, the zeta distribution is a discrete probability distribution. If X is a zeta-distributed random variable with parameter s, then the probability that X takes the integer value k is given by the probability mass function

Harmonic number Sum of the first n whole number reciprocals; 1/1 + 1/2 + 1/3 + ... + 1/n

In mathematics, the n-th harmonic number is the sum of the reciprocals of the first n natural numbers:

In number theory, the local zeta functionZ(Vs) is defined as

Hurwitz zeta function mathematics term

In mathematics, the Hurwitz zeta function, named after Adolf Hurwitz, is one of the many zeta functions. It is formally defined for complex arguments s with Re(s) > 1 and q with Re(q) > 0 by

Mertens function

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

In mathematics, the Mahler measureof a polynomial with complex coefficients is defined as

In mathematics, the von Mangoldt function is an arithmetic function named after German mathematician Hans von Mangoldt. It is an example of an important arithmetic function that is neither multiplicative nor additive.

In mathematics, the Selberg class is an axiomatic definition of a class of L-functions. The members of the class are Dirichlet series which obey four axioms that seem to capture the essential properties satisfied by most functions that are commonly called L-functions or zeta functions. Although the exact nature of the class is conjectural, the hope is that the definition of the class will lead to a classification of its contents and an elucidation of its properties, including insight into their relationship to automorphic forms and the Riemann hypothesis. The class was defined by Atle Selberg in, who preferred not to use the word "axiom" that later authors have employed.

Chebyshev function

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

In mathematics and analytic number theory, Vaughan's identity is an identity found by R. C. Vaughan (1977) that can be used to simplify Vinogradov's work on trigonometric sums. It can be used to estimate summatory functions of the form

In mathematics, the Grunsky matrices, or Grunsky operators, are matrices introduced by Grunsky (1939) in complex analysis and geometric function theory. They 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. Historically the inequalities were used in proving special cases of the Bieberbach conjecture up to the sixth coefficient; the exponentiated inequalities of Milin were used by de Branges in the final solution. The Grunsky operators and their Fredholm determinants are related to spectral properties of bounded domains in the complex plane. The operators have further applications in conformal mapping, Teichmüller theory and conformal field theory.

In number theory, the prime omega functions and count the number of prime factors of a natural number Thereby counts each distinct prime factor, whereas the related function counts the total number of prime factors of honoring their multiplicity. For example, if we have a prime factorization of of the form for distinct primes , then the respective prime omega functions are given by and . These prime factor counting functions have many important number theoretic relations.

The purpose of this page is to catalog new, interesting, and useful identities related to number-theoretic divisor sums, i.e., sums of an arithmetic function over the divisors of a natural number , or equivalently the Dirichlet convolution of an arithmetic function with one:

References