j-invariant

Source: Wikipedia, the free encyclopedia.
Klein's j-invariant in the complex plane

In

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

Rational functions of j are modular, and in fact give all modular functions. Classically, the j-invariant was studied as a parameterization of elliptic curves over , but it also has surprising connections to the symmetries of the Monster group (this connection is referred to as monstrous moonshine).

Definition

Real part of the j-invariant as a function of the square of the nome on the unit disk
Phase of the j-invariant as a function of the square of the nome on the unit disk

The j-invariant can be defined as a function on the

Im
(τ) > 0},

with the third definition implying can be expressed as a cube, also since 1728.

The given functions are the

modular discriminant
, Dedekind eta function , and modular invariants,

where , are Fourier series,

and , are Eisenstein series,

and (the square of the nome). The j-invariant can then be directly expressed in terms of the Eisenstein series as,

with no numerical factor other than 1728. This implies a third way to define the modular discriminant,[1]

For example, using the definitions above and , then the Dedekind eta function has the exact value,

implying the

transcendental numbers
,

but yielding the algebraic number (in fact, an integer),

In general, this can be motivated by viewing each τ as representing an isomorphism class of elliptic curves. Every elliptic curve E over C is a complex torus, and thus can be identified with a rank 2 lattice; that is, a two-dimensional lattice of C. This lattice can be rotated and scaled (operations that preserve the isomorphism class), so that it is generated by 1 and τH. This lattice corresponds to the elliptic curve (see

Weierstrass elliptic functions
).

Note that j is defined everywhere in H as the modular discriminant is non-zero. This is due to the corresponding cubic polynomial having distinct roots.

The fundamental region

The usual choice of a fundamental domain (gray) for the modular group acting on the upper half plane.

It can be shown that Δ is a

projective special linear group
PSL(2, Z).

By a suitable choice of transformation belonging to this group,

we may reduce τ to a value giving the same value for j, and lying in the

fundamental region
for j, which consists of values for τ satisfying the conditions

The function j(τ) when restricted to this region still takes on every value in the complex numbers C exactly once. In other words, for every c in C, there is a unique τ in the fundamental region such that c = j(τ). Thus, j has the property of mapping the fundamental region to the entire complex plane.

Additionally two values τ,τ' ∈H produce the same elliptic curve iff τ = T(τ') for some T ∈ PSL(2, Z). This means j provides a bijection from the set of elliptic curves over C to the complex plane.[2]

As a Riemann surface, the fundamental region has genus 0, and every (level one) modular function is a rational function in j; and, conversely, every rational function in j is a modular function. In other words, the field of modular functions is C(j).

Class field theory and j

The j-invariant has many remarkable properties:

  • If τ is any point of the upper half plane whose corresponding elliptic curve has
    singular moduli
    .
  • The field extension Q[j(τ), τ]/Q(τ) is abelian, that is, it has an abelian Galois group.
  • Let Λ be the lattice in C generated by {1, τ}. It is easy to see that all of the elements of Q(τ) which fix Λ under multiplication form a ring with units, called an
    unramified extensions
    of Q(τ).

These classical results are the starting point for the theory of complex multiplication.

Transcendence properties

In 1937 Theodor Schneider proved the aforementioned result that if τ is a quadratic irrational number in the upper half plane then j(τ) is an algebraic integer. In addition he proved that if τ is an algebraic number but not imaginary quadratic then j(τ) is transcendental.

The j function has numerous other transcendental properties. Kurt Mahler conjectured a particular transcendence result that is often referred to as Mahler's conjecture, though it was proved as a corollary of results by Yu. V. Nesterenko and Patrice Phillipon in the 1990s. Mahler's conjecture was that if τ was in the upper half plane then e and j(τ) were never both simultaneously algebraic. Stronger results are now known, for example if e is algebraic then the following three numbers are algebraically independent, and thus at least two of them transcendental:

The q-expansion and moonshine

Several remarkable properties of j have to do with its

q-expansion (Fourier series expansion), written as a Laurent series
in terms of q = e, which begins:

Note that j has a

simple pole
at the cusp, so its q-expansion has no terms below q−1.

All the Fourier coefficients are integers, which results in several

Ramanujan's constant
:

.

The

asymptotic formula
for the coefficient of qn is given by

,

as can be proved by the

Moonshine

More remarkably, the Fourier coefficients for the positive exponents of q are the dimensions of the graded part of an infinite-dimensional

moonshine theory
.

The study of the Moonshine conjecture led John Horton Conway and Simon P. Norton to look at the genus-zero modular functions. If they are normalized to have the form

then John G. Thompson showed that there are only a finite number of such functions (of some finite level), and Chris J. Cummins later showed that there are exactly 6486 of them, 616 of which have integral coefficients.[6]

Alternate expressions

We have

where x = λ(1 − λ) and λ is the modular lambda function

a ratio of Jacobi theta functions θm, and is the square of the elliptic modulus k(τ).[7] The value of j is unchanged when λ is replaced by any of the six values of the cross-ratio:[8]

The branch points of j are at {0, 1, ∞}, so that j is a

Belyi function.[9]

Expressions in terms of theta functions

Define the nome q = eπ and the Jacobi theta function,

from which one can derive the auxiliary theta functions, defined here. Let,

where ϑij and θn are alternative notations, and a4b4 + c4 = 0. Then we have the for

modular invariants
g2, g3,

and modular discriminant,

with Dedekind eta function η(τ). The j(τ) can then be rapidly computed,

Algebraic definition

So far we have been considering j as a function of a complex variable. However, as an invariant for isomorphism classes of elliptic curves, it can be defined purely algebraically.[10] Let

be a plane elliptic curve over any field. Then we may perform successive transformations to get the above equation into the standard form y2 = 4x3g2xg3 (note that this transformation can only be made when the characteristic of the field is not equal to 2 or 3). The resulting coefficients are:

where g2 = c4 and g3 = c6. We also have the discriminant

The j-invariant for the elliptic curve may now be defined as

In the case that the field over which the curve is defined has characteristic different from 2 or 3, this is equal to

Inverse function

The inverse function of the j-invariant can be expressed in terms of the hypergeometric function 2F1 (see also the article Picard–Fuchs equation). Explicitly, given a number N, to solve the equation j(τ) = N for τ can be done in at least four ways.

Method 1: Solving the

sextic
in λ,

where x = λ(1 − λ), and λ is the modular lambda function so the sextic can be solved as a cubic in x. Then,

for any of the six values of λ, where M is the arithmetic–geometric mean.[note 1]

Method 2: Solving the quartic in γ,

then for any of the four

roots
,

Method 3: Solving the cubic in β,

then for any of the three roots,

Method 4: Solving the quadratic in α,

then,

One root gives τ, and the other gives 1/τ, but since j(τ) = j(−1/τ), it makes no difference which α is chosen. The latter three methods can be found in

elliptic functions
to alternative bases.

The inversion is applied in high-precision calculations of elliptic function periods even as their ratios become unbounded.[

compass and straightedge constructions). The latter result is hardly evident since the modular equation for j of order 2 is cubic.[11]

Pi formulas

The Chudnovsky brothers found in 1987,[12]

a proof of which uses the fact that

For similar formulas, see the Ramanujan–Sato series.

Special values

The j-invariant vanishes at the "corner" of the fundamental domain:

Here are a few more special values[citation needed] given in terms of the alternative notation J(τ) ≡ 1/1728 j(τ), the first five well known:

Failure to classify elliptic curves over other fields

The -invariant is only sensitive to isomorphism classes of elliptic curves over the complex numbers, or more generally, an algebraically closed field. Over other fields there exist examples of elliptic curves whose -invariant is the same, but are non-isomorphic. For example, let be the elliptic curves associated to the polynomials

both having -invariant . Then, the rational points of can be computed as:

since There are no rational solutions with . This can be shown using

Cardano's formula
to show that in that case the solutions to are all irrational. On the other hand, on the set of points

the equation for becomes . Dividing by to eliminate the solution, the quadratic formula gives the rational solutions:

If these curves are considered over , there is an isomorphism sending

References

Notes

  1. ^ The equality holds if the arithmetic–geometric mean of complex numbers (such that ) is defined as follows: Let , , , where the signs are chosen such that for all . If , the sign is chosen such that . Then . When are positive real (with ), this definition coincides with the usual definition of the arithmetic–geometric mean for positive real numbers. See The Arithmetic-Geometric Mean of Gauss by David A. Cox.

Other