Grunsky matrix

Last updated

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.

Contents

Historically the inequalities for the disk 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. A detailed exposition using these methods can be found in Hayman (1994). The Grunsky operators and their Fredholm determinants are also 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.

Grunsky Matrix

If f(z) is a holomorphic univalent function on the unit disk, normalized so that f(0) = 0 and f′(0) = 1, the function

is a non-vanishing univalent function on |z| > 1 having a simple pole at ∞ with residue 1:

The same inversion formula applied to g gives back f and establishes a one-one correspondence between these two classes of function.

The Grunsky matrix (cnm) of g is defined by the equation

It is a symmetric matrix. Its entries are called the Grunsky coefficients of g.

Note that

so that the coefficients can be expressed directly in terms of f. Indeed, if

then for m, n > 0

and d0n = dn0 is given by

with

Grunsky inequalities

If f is a holomorphic function on the unit disk with Grunsky matrix (cnm), the Grunsky inequalities state that

for any finite sequence of complex numbers λ1, ..., λN.

Faber polynomials

The Grunsky coefficients of a normalized univalent function in |z| > 1

are polynomials in the coefficients bi which can be computed recursively in terms of the Faber polynomials Φn, a monic polynomial of degree n depending on g.

Taking the derivative in z of the defining relation of the Grunsky coefficients and multiplying by z gives

The Faber polynomials are defined by the relation

Dividing this relation by z and integrating between z and ∞ gives

This gives the recurrence relations for n > 0

with

Thus

so that for n ≥ 1

The latter property uniquely determines the Faber polynomial of g.

Milin's area theorem

Let g(z) be a univalent function on |z| > 1 normalized so that

and let f(z) be a non-constant holomorphic function on C.

If

is the Laurent expansion on z > 1, then

Proof

If Ω is a bounded open region with smooth boundary ∂Ω and h is a differentiable function on Ω extending to a continuous function on the closure, then, by Stokes' theorem applied to the differential 1-form

For r > 1, let Ωr be the complement of the image of |z|> r under g(z), a bounded domain. Then, by the above identity with h = f′, the area of fr) is given by

Hence

Since the area is non-negative

The result follows by letting r decrease to 1.

Milin's proof of Grunsky inequalities

If

then

Applying Milin's area theorem,

(Equality holds here if and only if the complement of the image of g has Lebesgue measure zero.)

So a fortiori

Hence the symmetric matrix

regarded as an operator on CN with its standard inner product, satisfies

So by the Cauchy–Schwarz inequality

With

this gives the Grunsky inequality:

Criterion for univalence

Let g(z) be a holomorphic function on z > 1 with

Then g is univalent if and only if the Grunsky coefficients of g satisfy the Grunsky inequalities for all N.

In fact the conditions have already been shown to be necessary. To see sufficiency, note that

makes sense when |z| and |ζ| are large and hence the coefficients cmn are defined. If the Grunsky inequalities are satisfied then it is easy to see that the |cmn| are uniformly bounded and hence the expansion on the left hand side converges for |z| > 1 and |ζ| > 1. Exponentiating both sides, this implies that g is univalent.

Pairs of univalent functions

Let and be univalent holomorphic functions on |z| < 1 and |ζ| > 1, such that their images are disjoint in C. Suppose that these functions are normalized so that

and

with a ≠ 0 and

The Grunsky matrix (cmn) of this pair of functions is defined for all non-zero m and n by the formulas:

with

so that (cmn) is a symmetric matrix.

In 1972 the American mathematician James Hummel extended the Grunsky inequalities to this matrix, proving that for any sequence of complex numbers λ±1, ..., λ±N

The proof proceeds by computing the area of the image of the complement of the images of |z| < r < 1 under F and |ζ| > R > 1 under g under a suitable Laurent polynomial h(w).

Let and denote the Faber polynomials of g and and set

Then:

The area equals

where C1 is the image of the circle |ζ| = R under g and C2 is the image of the circle |z| = r under F.

Hence

Since the area is positive, the right hand side must also be positive. Letting r increase to 1 and R decrease to 1, it follows that

with equality if and only if the complement of the images has Lebesgue measure zero.

As in the case of a single function g, this implies the required inequality.

Unitarity

The matrix

of a single function g or a pair of functions F, g is unitary if and only if the complement of the image of g or the union of the images of F and g has Lebesgue measure zero. So, roughly speaking, in the case of one function the image is a slit region in the complex plane; and in the case of two functions the two regions are separated by a closed Jordan curve.

In fact the infinite matrix A acting on the Hilbert space of square summable sequences satisfies

But if J denotes complex conjugation of a sequence, then

since A is symmetric. Hence

so that A is unitary.

Equivalent forms of Grunsky inequalities

Goluzin inequalities

If g(z) is a normalized univalent function in |z| > 1, z1, ..., zN are distinct points with |zn| > 1 and α1, ..., αN are complex numbers, the Goluzin inequalities, proved in 1947 by the Russian mathematician Gennadi Mikhailovich Goluzin (1906-1953), state that

To deduce them from the Grunsky inequalities, let

for k > 0.

Conversely the Grunsky inequalities follow from the Goluzin inequalities by taking

where

with r > 1, tending to ∞.

Bergman–Schiffer inequalities

Bergman & Schiffer (1951) gave another derivation of the Grunsky inequalities using reproducing kernels and singular integral operators in geometric function theory; a more recent related approach can be found in Baranov & Hedenmalm (2008).

Let f(z) be a normalized univalent function in |z| < 1, let z1, ..., zN be distinct points with |zn| < 1 and let α1, ..., αN be complex numbers. The Bergman-Schiffer inequalities state that

To deduce these inequalities from the Grunsky inequalities, set

for k > 0.

Conversely the Grunsky inequalities follow from the Bergman-Schiffer inequalities by taking

where

with r < 1, tending to 0.

Applications

The Grunsky inequalities imply many inequalities for univalent functions. They were also used by Schiffer and Charzynski in 1960 to give a completely elementary proof of the Bieberbach conjecture for the fourth coefficient; a far more complicated proof had previously been found by Schiffer and Garabedian in 1955. In 1968 Pedersen and Ozawa independently used the Grunsky inequalities to prove the conjecture for the sixth coefficient. [1] [2]

In the proof of Schiffer and Charzynski, if

is a normalized univalent function in |z| < 1, then

is an odd univalent function in |z| > 1.

Combining Gronwall's area theorem for f with the Grunsky inequalities for the first 2 x 2 minor of the Grunsky matrix of g leads to a bound for |a4| in terms of a simple function of a2 and a free complex parameter. The free parameter can be chosen so that the bound becomes a function of half the modulus of a2 and it can then be checked directly that this function is no greater than 4 on the range [0,1].

As Milin showed, the Grunsky inequalities can be exponentiated. The simplest case proceeds by writing

with an(w) holomorphic in |w| < 1.

The Grunsky inequalities, with λn = wn imply that

On the other hand, if

as formal power series, then the first of the Lebedev–Milin inequalities (1965) states that [3] [4]

Equivalently the inequality states that if g(z) is a polynomial with g(0) = 0, then

where A is the area of g(D),

To prove the inequality, note that the coefficients are determined by the recursive formula

so that by the Cauchy–Schwarz inequality

The quantities cn obtained by imposing equality here:

satisfy and hence, reversing the steps,

In particular defining bn(w) by the identity

the following inequality must hold for |w| < 1

Beurling transform

The Beurling transform (also called the Beurling-Ahlfors transform and the Hilbert transform in the complex plane) provides one of the most direct methods of proving the Grunsky inequalities, following Bergman & Schiffer (1951) and Baranov & Hedenmalm (2008).

The Beurling transform is defined on L2(C) as the operation of multiplication by on Fourier transforms. It thus defines a unitary operator. It can also be defined directly as a principal value integral [5]

For any bounded open region Ω in C it defines a bounded operator TΩ from the conjugate of the Bergman space of Ω onto the Bergman space of Ω: a square integrable holomorphic function is extended to 0 off Ω to produce a function in L2(C) to which T is applied and the result restricted to Ω, where it is holomorphic. If f is a holomorphic univalent map from the unit disk D onto Ω then the Bergman space of Ω and its conjugate can be identified with that of D and TΩ becomes the singular integral operator with kernel

It defines a contraction. On the other hand, it can be checked that TD = 0 by computing directly on powers using Stokes theorem to transfer the integral to the boundary.

It follows that the operator with kernel

acts as a contraction on the conjugate of the Bergman space of D. Hence, if

then

Grunsky operator and Fredholm determinant

If Ω is a bounded domain in C with smooth boundary, the operator TΩ can be regarded as a bounded antilinear contractive operator on the Bergman space H = A2(Ω). It is given by the formula

for u in the Hilbert space H= A2(Ω). TΩ is called the Grunsky operator of Ω (or f). Its realization on D using a univalent function f mapping D onto Ω and the fact that TD = 0 shows that it is given by restriction of the kernel

and is therefore a Hilbert–Schmidt operator.

The antilinear operator T = TΩ satisfies the self-adjointness relation

for u, v in H.

Thus A = T2 is a compact self-adjont linear operator on H with

so that A is a positive operator. By the spectral theorem for compact self-adjoint operators, there is an orthonormal basis un of H consisting of eigenvectors of A:

where μn is non-negative by the positivity of A. Hence

with λn ≥ 0. Since T commutes with A, it leaves its eigenspaces invariant. The positivity relation shows that it acts trivially on the zero eigenspace. The other non-zero eigenspaces are all finite-dimensional and mutually orthogonal. Thus an orthonormal basis can be chosen on each eigenspace so that:

(Note that by antilinearity of T.)

The non-zero λn (or sometimes their reciprocals) are called the Fredholm eigenvalues of Ω:

If Ω is a bounded domain that is not a disk, Ahlfors showed that

The Fredholm determinant for the domain Ω is defined by [6] [7]

Note that this makes sense because A = T2 is a trace class operator.

Schiffer & Hawley (1962) showed that if and f fixes 0, then [8] [9]

Here the norms are in the Bergman spaces of D and its complement Dc and g is a univalent map from Dc onto Ωc fixing ∞.

A similar formula applies in the case of a pair of univalent functions (see below).

Singular integral operators on a closed curve

Let Ω be a bounded simply connected domain in C with smooth boundary C = ∂Ω. Thus there is a univalent holomorphic map f from the unit disk D onto Ω extending to a smooth map between the boundaries S1 and C.

Notes

  1. Duren 1983 , pp. 131–133
  2. Koepf 2007
  3. Duren 1983 , pp. 143–144
  4. Apart from the elementary proof of this result presented here, there are several other analytic proofs in the literature. Nikolski (2002 , p. 220), following de Branges, notes that it is a consequence of standard inequalities connected with reproducing kernels. Widom (1988) observed that it was an immediate consequence of Szegő's limit formula (1951). Indeed if f is the real-valued trigonometric polynomial on the circle given as twice the real part of a polynomial g(z) vanishing at 0 on the unit disk, Szegő's limit formula states that the Toeplitz determinants of ef increase to eA where A is the area of g(D). The first determinant is by definition just the constant term in ef = |eg|2.
  5. Ahlfors 1966
  6. Schiffer 1959 , p. 261
  7. Schiffer & Hawley 1962 , p. 246
  8. Schiffer & Hawley 1962 , pp. 245–246
  9. Takhtajan & Teo 2006

Related Research Articles

In number theory, an arithmetic, arithmetical, or number-theoretic function is for most authors any function f(n) whose domain is the positive integers and whose range is a subset of the complex numbers. Hardy & Wright include in their definition the requirement that an arithmetical function "expresses some arithmetical property of n".

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.

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">Jensen's inequality</span> Theorem of convex functions

In mathematics, Jensen's inequality, named after the Danish mathematician Johan Jensen, relates the value of a convex function of an integral to the integral of the convex function. It was proved by Jensen in 1906, building on an earlier proof of the same inequality for doubly-differentiable functions by Otto Hölder in 1889. Given its generality, the inequality appears in many forms depending on the context, some of which are presented below. In its simplest form the inequality states that the convex transformation of a mean is less than or equal to the mean applied after convex transformation; it is a simple corollary that the opposite is true of concave transformations.

In mathematics, a Dirichlet series is any series of the form

In mathematics, the Weierstrass functions are special functions of a complex variable that are auxiliary to the Weierstrass elliptic function. They are named for Karl Weierstrass. The relation between the sigma, zeta, and functions is analogous to that between the sine, cotangent, and squared cosecant functions: the logarithmic derivative of the sine is the cotangent, whose derivative is negative the squared cosecant.

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 number theory, Li's criterion is a particular statement about the positivity of a certain sequence that is equivalent to the Riemann hypothesis. The criterion is named after Xian-Jin Li, who presented it in 1997. In 1999, Enrico Bombieri and Jeffrey C. Lagarias provided a generalization, showing that Li's positivity condition applies to any collection of points that lie on the Re(s) = 1/2 axis.

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

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

In mathematics, Lindelöf's theorem is a result in complex analysis named after the Finnish mathematician Ernst Leonard Lindelöf. It states that a holomorphic function on a half-strip in the complex plane that is bounded on the boundary of the strip and does not grow "too fast" in the unbounded direction of the strip must remain bounded on the whole strip. The result is useful in the study of the Riemann zeta function, and is a special case of the Phragmén–Lindelöf principle. Also, see Hadamard three-lines theorem.

In optimization, a self-concordant function is a function for which

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

Least-squares support-vector machines (LS-SVM) for statistics and in statistical modeling, are least-squares versions of support-vector machines (SVM), which are a set of related supervised learning methods that analyze data and recognize patterns, and which are used for classification and regression analysis. In this version one finds the solution by solving a set of linear equations instead of a convex quadratic programming (QP) problem for classical SVMs. Least-squares SVM classifiers were proposed by Johan Suykens and Joos Vandewalle. LS-SVMs are a class of kernel-based learning methods.

In statistics, the generalized Marcum Q-function of order is defined as

In mathematics, Grunsky's theorem, due to the German mathematician Helmut Grunsky, is a result in complex analysis concerning holomorphic univalent functions defined on the unit disk in the complex numbers. The theorem states that a univalent function defined on the unit disc, fixing the point 0, maps every disk |z| < r onto a starlike domain for r ≤ tanh π/4. The largest r for which this is true is called the radius of starlikeness of the function.

In mathematics, singular integral operators of convolution type are the singular integral operators that arise on Rn and Tn through convolution by distributions; equivalently they are the singular integral operators that commute with translations. The classical examples in harmonic analysis are the harmonic conjugation operator on the circle, the Hilbert transform on the circle and the real line, the Beurling transform in the complex plane and the Riesz transforms in Euclidean space. The continuity of these operators on L2 is evident because the Fourier transform converts them into multiplication operators. Continuity on Lp spaces was first established by Marcel Riesz. The classical techniques include the use of Poisson integrals, interpolation theory and the Hardy–Littlewood maximal function. For more general operators, fundamental new techniques, introduced by Alberto Calderón and Antoni Zygmund in 1952, were developed by a number of authors to give general criteria for continuity on Lp spaces. This article explains the theory for the classical operators and sketches the subsequent general theory.

In mathematics, the Neumann–Poincaré operator or Poincaré–Neumann operator, named after Carl Neumann and Henri Poincaré, is a non-self-adjoint compact operator introduced by Poincaré to solve boundary value problems for the Laplacian on bounded domains in Euclidean space. Within the language of potential theory it reduces the partial differential equation to an integral equation on the boundary to which the theory of Fredholm operators can be applied. The theory is particularly simple in two dimensions—the case treated in detail in this article—where it is related to complex function theory, the conjugate Beurling transform or complex Hilbert transform and the Fredholm eigenvalues of bounded planar domains.

Stochastic portfolio theory (SPT) is a mathematical theory for analyzing stock market structure and portfolio behavior introduced by E. Robert Fernholz in 2002. It is descriptive as opposed to normative, and is consistent with the observed behavior of actual markets. Normative assumptions, which serve as a basis for earlier theories like modern portfolio theory (MPT) and the capital asset pricing model (CAPM), are absent from SPT.

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. That is, 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.

References