Weil–Brezin Map

From HandWiki

In mathematics, the Weil–Brezin map, named after André Weil[1] and Jonathan Brezin,[2] is a unitary transformation that maps a Schwartz function on the real line to a smooth function on the Heisenberg manifold. The Weil–Brezin map gives a geometric interpretation of the Fourier transform, the Plancherel theorem and the Poisson summation formula.[3][4][5] The image of Gaussian functions under the Weil–Brezin map are nil-theta functions, which are related to theta functions. The Weil–Brezin map is sometimes referred to as the Zak transform,[6] which is widely applied in the field of physics and signal processing; however, the Weil–Brezin Map is defined via Heisenberg group geometrically, whereas there is no direct geometric or group theoretic interpretation from the Zak transform.

Heisenberg manifold

The (continuous) Heisenberg group [math]\displaystyle{ N }[/math] is the 3-dimensional Lie group that can be represented by triples of real numbers with multiplication rule

[math]\displaystyle{ \langle x,y,t\rangle \langle a,b,c\rangle = \langle x+a, y+b, t+c+xb\rangle. }[/math]

The discrete Heisenberg group [math]\displaystyle{ \Gamma }[/math] is the discrete subgroup of [math]\displaystyle{ N }[/math] whose elements are represented by the triples of integers. Considering [math]\displaystyle{ \Gamma }[/math] acts on [math]\displaystyle{ N }[/math] on the left, the quotient manifold [math]\displaystyle{ \Gamma\backslash N }[/math] is called the Heisenberg manifold. The Heisenberg group acts on the Heisenberg manifold on the right. The Haar measure [math]\displaystyle{ \mu = dx \wedge dy \wedge dt }[/math] on the Heisenberg group induces a right-translation-invariant measure on the Heisenberg manifold. The space of complex-valued square-integrable functions on the Heisenberg manifold has a right-translation-invariant orthogonal decomposition:

[math]\displaystyle{ L^2(\Gamma\backslash N) = \oplus_{n \in \mathbb{Z}} H_n }[/math]

where

[math]\displaystyle{ H_n =\{f\in L^2(\Gamma\backslash N) \mid f(\Gamma \langle x, y, t+s\rangle) = \exp(2\pi i ns) f(\Gamma \langle x,y,t\rangle)\} }[/math].

Definition

The Weil–Brezin map [math]\displaystyle{ W: L^2(\mathbb R) \to H_1 }[/math] is the unitary transformation given by

[math]\displaystyle{ W(\psi) (\Gamma \langle x, y, t \rangle) = \sum_{l\in \mathbb Z} \psi(x + l) e^{2 \pi i l y} e^{2\pi i t} }[/math]

for every Schwartz function [math]\displaystyle{ \psi }[/math], where convergence is pointwise.

The inverse of the Weil–Brezin map [math]\displaystyle{ W^{-1}: H_1 \to L^2(\mathbb R) }[/math] is given by

[math]\displaystyle{ (W^{-1}f) (x) = \int_0^{1} f(\Gamma \langle x, y, 0\rangle) dy }[/math]

for every smooth function [math]\displaystyle{ f }[/math] on the Heisenberg manifold that is in [math]\displaystyle{ H_1 }[/math].

Fundamental unitary representation of the Heisenberg group

For each real number [math]\displaystyle{ \lambda\ne 0 }[/math], the fundamental unitary representation [math]\displaystyle{ U_{\lambda} }[/math] of the Heisenberg group is an irreducible unitary representation of [math]\displaystyle{ N }[/math] on [math]\displaystyle{ L^2(\mathbb{R}) }[/math] defined by

[math]\displaystyle{ (U_{\lambda}(\langle a, b, c \rangle) \psi) (x) = e^{2 \pi i \lambda (c + bx)} \psi(x +a) }[/math].

By Stone–von Neumann theorem, this is the unique irreducible representation up to unitary equivalence satisfying the canonical commutation relation

[math]\displaystyle{ U_{\lambda}(\langle a, 0, 0 \rangle) U_{\lambda}(\langle 0, b, 0 \rangle) =e^{2\pi i \lambda ab} U_{\lambda}(\langle 0, b, 0 \rangle) U_{\lambda}(\langle a, 0, 0 \rangle) }[/math].

The fundamental representation [math]\displaystyle{ U=U_1 }[/math] of [math]\displaystyle{ N }[/math] on [math]\displaystyle{ L^2(\mathbb{R}) }[/math] and the right translation [math]\displaystyle{ R }[/math] of [math]\displaystyle{ N }[/math] on [math]\displaystyle{ H_1 \subset L^2(\Gamma \backslash N) }[/math] are intertwined by the Weil–Brezin map

[math]\displaystyle{ W U(\langle a, b, c \rangle) = R(\langle a, b, c \rangle) W }[/math].

In other words, the fundamental representation [math]\displaystyle{ U }[/math] on [math]\displaystyle{ L^2(\mathbb{R}) }[/math] is unitarily equivalent to the right translation [math]\displaystyle{ R }[/math] on [math]\displaystyle{ H_1 }[/math] through the Wei-Brezin map.

Relation to Fourier transform

Let [math]\displaystyle{ J: N \to N }[/math] be the automorphism on the Heisenberg group given by

[math]\displaystyle{ J(\langle x, y, t \rangle) =\langle y, -x, t-xy \rangle }[/math].

It naturally induces a unitary operator [math]\displaystyle{ J^* : H_1 \to H_1 }[/math], then the Fourier transform

[math]\displaystyle{ \mathcal F = W^{-1} J^{*} W }[/math]

as a unitary operator on [math]\displaystyle{ L^2(\mathbb{R}) }[/math].

Plancherel theorem

The norm-preserving property of [math]\displaystyle{ W }[/math] and [math]\displaystyle{ J^* }[/math], which is easily seen, yields the norm-preserving property of the Fourier transform, which is referred to as the Plancherel theorem.

Poisson summation formula

For any Schwartz function [math]\displaystyle{ \psi }[/math],

[math]\displaystyle{ \sum_l\psi(l) = W(\psi)(\Gamma\langle 0, 0, 0) \rangle) = (J^*W(\psi))(\Gamma\langle 0, 0, 0) \rangle) = W (\hat{\psi})(\Gamma\langle 0, 0, 0) \rangle)=\sum_l \hat{\psi}(l) }[/math].

This is just the Poisson summation formula.

Relation to the finite Fourier transform

For each [math]\displaystyle{ n\ne 0 }[/math], the subspace [math]\displaystyle{ H_n\subset L^2(\Gamma\backslash N) }[/math] can further be decomposed into right-translation-invariant orthogonal subspaces

[math]\displaystyle{ H_n = \oplus_{m=0}^{|n|-1} H_{n,m} }[/math]

where

[math]\displaystyle{ H_{n,m} =\{f\in H_n \mid f(\Gamma \langle x, y+ {1\over n}, t\rangle) = e^{2\pi i m/n}f(\Gamma \langle x, y, t\rangle)\} }[/math].

The left translation [math]\displaystyle{ L(\langle 0, 1/n, 0\rangle) }[/math] is well-defined on [math]\displaystyle{ H_n }[/math], and [math]\displaystyle{ H_{n,0}, ... , H_{n, |n|-1} }[/math] are its eigenspaces.

The left translation [math]\displaystyle{ L(\langle m/n, 0, 0\rangle) }[/math] is well-defined on [math]\displaystyle{ H_n }[/math], and the map

[math]\displaystyle{ L(\langle m/n, 0, 0\rangle) : H_{n,0} \to H_{n,m} }[/math]

is a unitary transformation.

For each [math]\displaystyle{ n\ne 0 }[/math], and [math]\displaystyle{ m = 0, ..., |n|-1 }[/math], define the map [math]\displaystyle{ W_{n, m}: L^2(\mathbb R) \to H_{n, m} }[/math] by

[math]\displaystyle{ W_{n,m}(\psi) (\Gamma \langle x, y, t \rangle) = \sum_{l\in \mathbb Z} \psi(x + l + {m \over n}) e^{2 \pi i (nl+m) y} e^{2\pi i n t} }[/math]

for every Schwartz function [math]\displaystyle{ \psi }[/math], where convergence is pointwise.

[math]\displaystyle{ W_{n,m} =L(\langle m/n, 0, 0\rangle) \circ W_{n,0}. }[/math]

The inverse map [math]\displaystyle{ W_{n,m}^{-1}: H_{n,m} \to L^2(\mathbb R) }[/math] is given by

[math]\displaystyle{ (W_{n,m}^{-1}f) (x) = \int_0^{1} e^{-2\pi i m y} f(\Gamma \langle x - {m \over n}, y, 0\rangle) dy }[/math]

for every smooth function [math]\displaystyle{ f }[/math] on the Heisenberg manifold that is in [math]\displaystyle{ H_{n,m} }[/math].

Similarly, the fundamental unitary representation [math]\displaystyle{ U_n }[/math] of the Heisenberg group is unitarily equivalent to the right translation on [math]\displaystyle{ H_{n,m} }[/math] through [math]\displaystyle{ W_{n,m} }[/math]:

[math]\displaystyle{ W_{n,m} U_n(\langle a, b, c \rangle) = R(\langle a, b, c \rangle) W_{n,m} }[/math].

For any [math]\displaystyle{ m, m' }[/math],

[math]\displaystyle{ (W_{n,m'}^{-1}J^* W_{n,m} \psi ) (x) = e^{2\pi i m'm / n} \hat{\psi}(nx) }[/math].

For each [math]\displaystyle{ n\gt 0 }[/math], let [math]\displaystyle{ \phi_n(x) =(2n)^{1/4} e^{- \pi n x^2 } }[/math]. Consider the finite dimensional subspace [math]\displaystyle{ K_n }[/math] of [math]\displaystyle{ H_n }[/math] generated by [math]\displaystyle{ \{\boldsymbol{e}_{n,0}, ..., \boldsymbol{e}_{n,n-1}\} }[/math] where

[math]\displaystyle{ \boldsymbol{e}_{n,m} = W_{n,m} ( \phi_n ) \in H_{n,m}. }[/math]

Then the left translations [math]\displaystyle{ L(\langle 1/n, 0, 0\rangle) }[/math] and [math]\displaystyle{ L(\langle 0, 1/n, 0\rangle) }[/math] act on [math]\displaystyle{ K_n }[/math] and give rise to the irreducible representation of the finite Heisenberg group. The map [math]\displaystyle{ J^* }[/math] acts on [math]\displaystyle{ K_n }[/math] and gives rise to the finite Fourier transform

[math]\displaystyle{ J^* \boldsymbol{e}_{n,m} = {1 \over \sqrt{n}}\sum_{m'} e^{2\pi i m'm / n} \boldsymbol{e}_{n,m'}. }[/math]

Nil-theta functions

Nil-theta functions are functions on the Heisenberg manifold that are analogous to the theta functions on the complex plane. The image of Gaussian functions under the Weil–Brezin Map are nil-theta functions. There is a model[7] of the finite Fourier transform defined with nil-theta functions, and the nice property of the model is that the finite Fourier transform is compatible with the algebra structure of the space of nil-theta functions.

Definition of nil-theta functions

Let [math]\displaystyle{ \mathfrak{n} }[/math] be the complexified Lie algebra of the Heisenberg group [math]\displaystyle{ N }[/math]. A basis of [math]\displaystyle{ \mathfrak{n} }[/math] is given by the left-invariant vector fields [math]\displaystyle{ X, Y, T }[/math] on [math]\displaystyle{ N }[/math]:

[math]\displaystyle{ X(x,y,t) = {\partial \over \partial x}, }[/math]
[math]\displaystyle{ Y(x,y,t) = {\partial \over \partial y} + x {\partial \over \partial t}, }[/math]
[math]\displaystyle{ T(x,y,t) = {\partial \over \partial t}. }[/math]

These vector fields are well-defined on the Heisenberg manifold [math]\displaystyle{ \Gamma \backslash N }[/math].

Introduce the notation [math]\displaystyle{ V_{-i} = X-i Y }[/math]. For each [math]\displaystyle{ n\gt 0 }[/math], the vector field [math]\displaystyle{ V_{-i} }[/math] on the Heisenberg manifold can be thought of as a differential operator on [math]\displaystyle{ C^{\infty} (\Gamma \backslash N) \cap H_{n,m} }[/math] with the kernel generated by [math]\displaystyle{ \boldsymbol{e}_{n,m} }[/math].

We call

[math]\displaystyle{ \ker(V_{-i}: C^{\infty} (\Gamma \backslash N) \cap H_n \to H_n) = \left\{ \begin{array}{lr} K_n, & n\gt 0 \\ \mathbb{C}, & n=0 \end{array} \right. }[/math]

the space of nil-theta functions of degree [math]\displaystyle{ n }[/math].

Algebra structure of nil-theta functions

The nil-theta functions with pointwise multiplication on [math]\displaystyle{ \Gamma \backslash N }[/math] form a graded algebra [math]\displaystyle{ \oplus_{n\ge 0} K_n }[/math] (here [math]\displaystyle{ K_0 = \mathbb{C} }[/math]).

Auslander and Tolimieri showed that this graded algebra is isomorphic to

[math]\displaystyle{ \mathbb{C}[x_1, x_2^2, x_3^3]/(x_3^6 + x_1^4 x_2^2 + x_2^6) }[/math],

and that the finite Fourier transform (see the preceding section #Relation to the finite Fourier transform) is an automorphism of the graded algebra.

Relation to Jacobi theta functions

Let [math]\displaystyle{ \vartheta(z; \tau) = \sum_{l =-\infty}^\infty \exp (\pi i l^2 \tau + 2 \pi i l z) }[/math] be the Jacobi theta function. Then

[math]\displaystyle{ \vartheta(n(x+iy); ni) = (2n)^{-1/4} e^{\pi n y^2} \boldsymbol{e}_{n,0}(\Gamma\langle y, x, 0 \rangle) }[/math].

Higher order theta functions with characteristics

An entire function [math]\displaystyle{ f }[/math] on [math]\displaystyle{ \mathbb{C} }[/math] is called a theta function of order [math]\displaystyle{ n }[/math], period [math]\displaystyle{ \tau }[/math] ([math]\displaystyle{ \mathrm{Im}(\tau)\gt 0 }[/math]) and characteristic [math]\displaystyle{ [^a_b] }[/math] if it satisfies the following equations:

  1. [math]\displaystyle{ f(z+1) = \exp(\pi i a ) f(z) }[/math],
  2. [math]\displaystyle{ f(z+\tau) =\exp(\pi i b) \exp(-\pi i n (2z +\tau)) f(z) }[/math].

The space of theta functions of order [math]\displaystyle{ n }[/math], period [math]\displaystyle{ \tau }[/math] and characteristic [math]\displaystyle{ [^a_b] }[/math] is denoted by [math]\displaystyle{ \Theta_n[^a_b](\tau, A) }[/math].

[math]\displaystyle{ \dim \Theta_n[^a_b](\tau, A) =n }[/math].

A basis of [math]\displaystyle{ \Theta_n[^0_0](i, A) }[/math] is

[math]\displaystyle{ \theta_{n,m}(z) = \sum_{l\in \mathbb{Z}} \exp [ -\pi n (l+{m \over n} )^2 + 2 \pi i (l n+ m) z ) ] }[/math].

These higher order theta functions are related to the nil-theta functions by

[math]\displaystyle{ \theta_{n,m}(x+ iy) = (2n)^{-1/4} e^{\pi n y^2} \boldsymbol{e}_{n,m} (\Gamma \langle y, x, 0 \rangle ) }[/math].

See also

References

  1. Weil, André. "Sur certains groupes d'opérateurs unitaires." Acta mathematica 111.1 (1964): 143-211.
  2. Brezin, Jonathan. "Harmonic analysis on nilmanifolds." Transactions of the American Mathematical Society 150.2 (1970): 611-618.
  3. Auslander, Louis, and Richard Tolimieri. Abelian harmonic analysis, theta functions and function algebras on a nilmanifold. Springer, 1975.
  4. Auslander, Louis. "Lecture notes on nil-theta functions." Conference Board of the Mathematical Sciences, 1977.
  5. Zhang, D. "Integer Linear Canonical Transforms, Their Discretization, and Poisson Summation Formulae"
  6. "Zak Transform". http://mathworld.wolfram.com/ZakTransform.html. 
  7. Auslander, L., and R. Tolimieri. "Algebraic structures for⨁Σ _ {𝑛≥ 1} 𝐿2 (𝑍/𝑛) compatible with the finite Fourier transform." Transactions of the American Mathematical Society 244 (1978): 263-272.