Dedekind eta function

Last updated
Dedekind e-function in the upper half-plane Dedekind Eta.jpg
Dedekind η-function in the upper half-plane

In mathematics, the Dedekind eta function, named after Richard Dedekind, is a modular form of weight 1/2 and is a function defined on the upper half-plane of complex numbers, where the imaginary part is positive. It also occurs in bosonic string theory.

Contents

Definition

For any complex number τ with Im(τ) > 0, let q = e2πiτ; then the eta function is defined by,

Raising the eta equation to the 24th power and multiplying by (2π)12 gives

where Δ is the modular discriminant. The presence of 24 can be understood by connection with other occurrences, such as in the 24-dimensional Leech lattice.

The eta function is holomorphic on the upper half-plane but cannot be continued analytically beyond it.

Modulus of Euler phi on the unit disc, colored so that black = 0, red = 4 Q-Eulero.jpeg
Modulus of Euler phi on the unit disc, colored so that black = 0, red = 4
The real part of the modular discriminant as a function of q. Discriminant real part.jpeg
The real part of the modular discriminant as a function of q.

The eta function satisfies the functional equations [1]

In the second equation the branch of the square root is chosen such that = 1 when τ = i.

More generally, suppose a, b, c, d are integers with adbc = 1, so that

is a transformation belonging to the modular group. We may assume that either c > 0, or c = 0 and d = 1. Then

where

Here s(h,k) is the Dedekind sum

Because of these functional equations the eta function is a modular form of weight 1/2 and level 1 for a certain character of order 24 of the metaplectic double cover of the modular group, and can be used to define other modular forms. In particular the modular discriminant of Weierstrass with

can be defined as

and is a modular form of weight 12. Some authors omit the factor of (2π)12, so that the series expansion has integral coefficients.

The Jacobi triple product implies that the eta is (up to a factor) a Jacobi theta function for special values of the arguments: [2]

where χ(n) is "the" Dirichlet character modulo 12 with χ(±1) = 1 and χ(±5) = −1. Explicitly,[ citation needed ]

The Euler function

has a power series by the Euler identity:

Note that by using Euler Pentagonal number theorem for , the eta function can be expressed as

This can be proved by using in Euler Pentagonal number theorem with the definition of eta function.

Because the eta function is easy to compute numerically from either power series, it is often helpful in computation to express other functions in terms of it when possible, and products and quotients of eta functions, called eta quotients, can be used to express a great variety of modular forms.

The picture on this page shows the modulus of the Euler function: the additional factor of q1/24 between this and eta makes almost no visual difference whatsoever. Thus, this picture can be taken as a picture of eta as a function of q.

Combinatorial identities

The theory of the algebraic characters of the affine Lie algebras gives rise to a large class of previously unknown identities for the eta function. These identities follow from the Weyl–Kac character formula, and more specifically from the so-called "denominator identities". The characters themselves allow the construction of generalizations of the Jacobi theta function which transform under the modular group; this is what leads to the identities. An example of one such new identity [3] is

where q = e2πiτ is the q-analog or "deformation" of the highest weight of a module.

Special values

From the above connection with the Euler function together with the special values of the latter, it can be easily deduced that

Eta quotients

Eta quotients are defined by quotients of the form

where d is a non-negative integer and rd is any integer. Linear combinations of eta quotients at imaginary quadratic arguments may be algebraic, while combinations of eta quotients may even be integral. For example, define,

with the 24th power of the Weber modular function 𝔣(τ). Then,

and so on, values which appear in Ramanujan–Sato series.

Eta quotients may also be a useful tool for describing bases of modular forms, which are notoriously difficult to compute and express directly. In 1993 Basil Gordon and Kim Hughes proved that if an eta quotient ηg of the form given above, namely satisfies

then ηg is a weight k modular form for the congruence subgroup Γ0(N) (up to holomorphicity) where [4]

This result was extended in 2019 such that the converse holds for cases when N is coprime to 6, and it remains open that the original theorem is sharp for all integers N. [5] This also extends to state that any modular eta quotient for any level n congruence subgroup must also be a modular form for the group Γ(N). While these theorems characterize modular eta quotients, the condition of holomorphicity must be checked separately using a theorem that emerged from the work of Gérard Ligozat [6] and Yves Martin: [7]

If ηg is an eta quotient satisfying the above conditions for the integer N and c and d are coprime integers, then the order of vanishing at the cusp c/d relative to Γ0(N) is

These theorems provide an effective means of creating holomorphic modular eta quotients, however this may not be sufficient to construct a basis for a vector space of modular forms and cusp forms. A useful theorem for limiting the number of modular eta quotients to consider states that a holomorphic weight k modular eta quotient on Γ0(N) must satisfy

where ordp(N) denotes the largest integer m such that pm divides N. [8] These results lead to several characterizations of spaces of modular forms that can be spanned by modular eta quotients. [8] Using the graded ring structure on the ring of modular forms, we can compute bases of vector spaces of modular forms composed of -linear combinations of eta-quotients. For example, if we assume N = pq is a semiprime then the following process can be used to compute an eta-quotient basis of Mk0(N)). [5]

  1. Fix a semiprime N = pq which is coprime to 6 (that is, p, q > 3). We know that any modular eta quotient may be found using the above theorems, therefore it is reasonable to algorithmically to compute them.
  2. Compute the dimension D of Mk0(N)). This tells us how many linearly-independent modular eta quotients we will need to compute to form a basis.
  3. Reduce the number of eta quotients to consider. For semiprimes we can reduce the number of partitions using the bound on

    and by noticing that the sum of the orders of vanishing at the cusps of Γ0(N) must equal

    . [5]
  4. Find all partitions of S into 4-tuples (there are 4 cusps of Γ0(N)), and among these consider only the partitions which satisfy Gordon and Hughes' conditions (we can convert orders of vanishing into exponents). Each of these partitions corresponds to a unique eta quotient.
  5. Determine the minimum number of terms in the q-expansion of each eta quotient required to identify elements uniquely (this uses a result known as Sturm's bound). Then use linear algebra to determine a maximal independent set among these eta quotients.
  6. Assuming that we have not already found D linearly independent eta quotients, find an appropriate vector space Mk0(N)) such that k and Mk0(N)) is spanned by (weakly holomorphic) eta quotients, [8] and Mkk0(N)) contains an eta quotient ηg.
  7. Take a modular form f with weight k that is not in the span of our computed eta quotients, and compute fηg as a linear combination of eta-quotients in Mk0(N)) and then divide out by ηg. The result will be an expression of f as a linear combination of eta quotients as desired. Repeat this until a basis is formed.

A collection of over 6300 product identities for the Dedekind Eta Function in a canonical, standardized form is available at the Wayback machine [9] of Michael Somos' website.

See also

Related Research Articles

<span class="mw-page-title-main">Error function</span> Sigmoid shape special function

In mathematics, the error function, often denoted by erf, is a function defined as:

<span class="mw-page-title-main">Weierstrass elliptic function</span> Class of mathematical functions

In mathematics, the Weierstrass elliptic functions are elliptic functions that take a particularly simple form. They are named for Karl Weierstrass. This class of functions are also referred to as ℘-functions and they are usually denoted by the symbol ℘, a uniquely fancy script p. They play an important role in the theory of elliptic functions. A ℘-function together with its derivative can be used to parameterize elliptic curves and they generate the field of elliptic functions with respect to a given period lattice.

<span class="mw-page-title-main">Theta function</span> Special functions of several complex variables

In mathematics, theta functions are special functions of several complex variables. They show up in many topics, including Abelian varieties, moduli spaces, quadratic forms, and solitons. As Grassmann algebras, they appear in quantum field theory.

<i>j</i>-invariant Modular function in mathematics

In mathematics, Felix Klein's j-invariant or j function, regarded as a function of a complex variable τ, is a modular function of weight zero for SL(2, Z) defined on the upper half-plane of complex numbers. It is the unique such function which is holomorphic away from a simple pole at the cusp such that

In number theory, a Heegner number is a square-free positive integer d such that the imaginary quadratic field has class number 1. Equivalently, the ring of algebraic integers of has unique factorization.

In mathematics, the Chowla–Selberg formula is the evaluation of a certain product of values of the gamma function at rational values in terms of values of the Dedekind eta function at imaginary quadratic irrational numbers. The result was essentially found by Lerch and rediscovered by Chowla and Selberg.

Bosonic string theory is the original version of string theory, developed in the late 1960s and named after Satyendra Nath Bose. It is so called because it contains only bosons in the spectrum.

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

In mathematics, the Euler function is given by

<span class="mw-page-title-main">Bring radical</span> Real root of the polynomial x^5+x+a

In algebra, the Bring radical or ultraradical of a real number a is the unique real root of the polynomial

In mathematics, the classical Kronecker limit formula describes the constant term at s = 1 of a real analytic Eisenstein series in terms of the Dedekind eta function. There are many generalizations of it to more complicated Eisenstein series. It is named for Leopold Kronecker.

In mathematics, a mock modular form is the holomorphic part of a harmonic weak Maass form, and a mock theta function is essentially a mock modular form of weight 1/2. The first examples of mock theta functions were described by Srinivasa Ramanujan in his last 1920 letter to G. H. Hardy and in his lost notebook. Sander Zwegers discovered that adding certain non-holomorphic functions to them turns them into harmonic weak Maass forms.

In physics, the Lemaître–Tolman metric, also known as the Lemaître–Tolman–Bondi metric or the Tolman metric, is a Lorentzian metric based on an exact solution of Einstein's field equations; it describes an isotropic and expanding universe which is not homogeneous, and is thus used in cosmology as an alternative to the standard Friedmann–Lemaître–Robertson–Walker metric to model the expansion of the universe. It has also been used to model a universe which has a fractal distribution of matter to explain the accelerating expansion of the universe. It was first found by Georges Lemaître in 1933 and Richard Tolman in 1934 and later investigated by Hermann Bondi in 1947.

In mathematics, every analytic function can be used for defining a matrix function that maps square matrices with complex entries to square matrices of the same size.

Bilinear time–frequency distributions, or quadratic time–frequency distributions, arise in a sub-field of signal analysis and signal processing called time–frequency signal processing, and, in the statistical analysis of time series data. Such methods are used where one needs to deal with a situation where the frequency composition of a signal may be changing over time; this sub-field used to be called time–frequency signal analysis, and is now more often called time–frequency signal processing due to the progress in using these methods to a wide range of signal-processing problems.

<span class="mw-page-title-main">Rogers–Ramanujan continued fraction</span> Continued fraction closely related to the Rogers–Ramanujan identities

The Rogers–Ramanujan continued fraction is a continued fraction discovered by Rogers (1894) and independently by Srinivasa Ramanujan, and closely related to the Rogers–Ramanujan identities. It can be evaluated explicitly for a broad class of values of its argument.

<span class="mw-page-title-main">Cnoidal wave</span> Nonlinear and exact periodic wave solution of the Korteweg–de Vries equation

In fluid dynamics, a cnoidal wave is a nonlinear and exact periodic wave solution of the Korteweg–de Vries equation. These solutions are in terms of the Jacobi elliptic function cn, which is why they are coined cnoidal waves. They are used to describe surface gravity waves of fairly long wavelength, as compared to the water depth.

<span class="mw-page-title-main">Modular lambda function</span> Symmetric holomorphic function

In mathematics, the modular lambda function λ(τ) is a highly symmetric Holomorphic function on the complex upper half-plane. It is invariant under the fractional linear action of the congruence group Γ(2), and generates the function field of the corresponding quotient, i.e., it is a Hauptmodul for the modular curve X(2). Over any point τ, its value can be described as a cross ratio of the branch points of a ramified double cover of the projective line by the elliptic curve , where the map is defined as the quotient by the [−1] involution.

In mathematics, the Weber modular functions are a family of three functions f, f1, and f2, studied by Heinrich Martin Weber.

In mathematics, a Ramanujan–Sato series generalizes Ramanujan’s pi formulas such as,

In mathematics, the Neville theta functions, named after Eric Harold Neville, are defined as follows:

References

  1. Siegel, C. L. (1954). "A Simple Proof of η(−1/τ) = η(τ)τ/i". Mathematika . 1: 4. doi:10.1112/S0025579300000462.
  2. Bump, Daniel (1998), Automorphic Forms and Representations, Cambridge University Press, ISBN   0-521-55098-X
  3. Fuchs, Jurgen (1992), Affine Lie Algebras and Quantum Groups, Cambridge University Press, ISBN   0-521-48412-X
  4. Gordon, Basil; Hughes, Kim (1993). "Multiplicative properties of η-products. II.". A Tribute to Emil Grosswald: Number Theory and Related Analysis. Contemporary Mathematics. Vol. 143. Providence, RI: American Mathematical Society. p. 415–430.
  5. 1 2 3 Allen, Michael; Anderson, Nicholas; Hamakiotes, Asimina; Oltsik, Ben; Swisher, Holly (2020). "Eta-quotients of prime or semiprime level and elliptic curves". Involve. 13 (5): 879–900. arXiv: 1901.10511 . doi:10.2140/involve.2020.13.879. S2CID   119620241.
  6. Ligozat, G. (1974). Courbes modulaires de genre 1. Publications Mathématiques d'Orsay. Vol. 75. U.E.R. Mathématique, Université Paris XI, Orsay. p. 7411.
  7. Martin, Yves (1996). "Multiplicative η-quotients". Transactions of the American Mathematical Society . 348 (12): 4825–4856. doi: 10.1090/S0002-9947-96-01743-6 .
  8. 1 2 3 Rouse, Jeremy; Webb, John J. (2015). "On spaces of modular forms spanned by eta-quotients". Advances in Mathematics . 272: 200–224. arXiv: 1311.1460 . doi: 10.1016/j.aim.2014.12.002 .
  9. "Dedekind Eta Function Product Identities by Michael Somos". Archived from the original on 2019-07-09.

Further reading