Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 The Riemann problem  





2 The Hilbert problem  





3 RiemannHilbert problems  





4 Matrix RiemannHilbert problems  



4.1  Example: Scalar RiemannHilbert factorization problem  







5 Generalizations  



5.1  DBAR problem  





5.2  Generalized analytic functions  







6 Applications to integrability theory  



6.1  Use for asymptotics  







7 See also  





8 Notes  





9 References  














RiemannHilbert problem






Deutsch
Español

 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


Inmathematics, Riemann–Hilbert problems, named after Bernhard Riemann and David Hilbert, are a class of problems that arise in the study of differential equations in the complex plane. Several existence theorems for Riemann–Hilbert problems have been produced by Mark Krein, Israel Gohberg and others.[1]

The Riemann problem

[edit]

Suppose that is a smooth, simple, closed contour in the complex plane.[2] Divide the plane into two parts denoted by (the inside) and (the outside), determined by the index of the contour with respect to a point. The classical problem, considered in Riemann's PhD dissertation, was that of finding a function

analytic inside , such that the boundary values of along satisfy the equation

for , where , and are given real-valued functions.[3][4] For example, in the special case where and is a circle, the problem reduces to deriving the Poisson formula.[5]

By the Riemann mapping theorem, it suffices to consider the case when is the circle group .[6] In this case, one may seek along with its Schwarz reflection

For , one has and so

Hence the problem reduces to finding a pair of analytic functions and on the inside and outside, respectively, of the unit disk, so that on the unit circle

and, moreover, so that the condition at infinity holds:

The Hilbert problem

[edit]

Hilbert's generalization of the problem attempted to find a pair of analytic functions and on the inside and outside, respectively, of the curve , such that for one has

where , and are given complex-valued functions (no longer just complex conjugates).[7]

Riemann–Hilbert problems

[edit]

In the Riemann problem as well as Hilbert's generalization, the contour was simple. A full Riemann–Hilbert problem allows that the contour may be composed of a union of several oriented smooth curves, with no intersections. The "+" and『−』sides of the "contour" may then be determined according to the index of a point with respect to . The Riemann–Hilbert problem is to find a pair of analytic functions and on the "+" and『−』side of , respectively, such that for one has

where , and are given complex-valued functions.

Matrix Riemann–Hilbert problems

[edit]

Given an oriented contour (technically: an oriented union of smooth curves without points of infinite self-intersection in the complex plane), a Riemann–Hilbert factorization problem is the following.

Given a matrix function defined on the contour , find a holomorphic matrix function defined on the complement of , such that the following two conditions are satisfied[8]

  1. If and denote the non-tangential limits of as we approach , then , at all points of non-intersection in .
  2. tends to the identity matrix as along any direction outside .

In the simplest case is smooth and integrable. In more complicated cases it could have singularities. The limits and could be classical and continuous or they could be taken in the -sense. At end-points or intersection points of the contour , the jump condition is not defined; constraints on the growth of near those points have to be posed to ensure uniqueness (see the scalar problem below).

Example: Scalar Riemann–Hilbert factorization problem

[edit]

Suppose and . Assuming is bounded, what is the solution ?

To solve this, let's take the logarithm of equation .

Since tends to , as.

A standard fact about the Cauchy transform is that where and are the limits of the Cauchy transform from above and below ; therefore, we get

when . Because the solution of a Riemann–Hilbert factorization problem is unique (an easy application of Liouville's theorem (complex analysis)), the Sokhotski–Plemelj theorem gives the solution. We get

and therefore

which has a branch cut at contour .

Check:

therefore,

CAVEAT 1: If the problem is not scalar one cannot easily take logarithms. In general explicit solutions are very rare.

CAVEAT 2: The boundedness (or at least a constraint on the blow-up) of near the special points and is crucial. Otherwise any function of the form

is also a solution. In general, conditions on growth are necessary at special points (the end-points of the jump contour or intersection point) to ensure that the problem is well-posed.

Generalizations

[edit]

DBAR problem

[edit]

Suppose is some simply connected domain of the complex plane. Then the scalar equation

is a generalization of a Riemann-Hilbert problem, called the DBAR problem (or problem). It is the complex form of the nonhomogeneous Cauchy-Riemann equations. To show this, let

with , , and all real-valued functions of real variables and . Then, using

the DBAR problem yields

As such, if is holomorphic for , then the Cauchy-Riemann equations must be satisfied.[9]

In case as and , the solution of the DBAR problem is[10]

integrated over the entire complex plane; denoted by , and where the wedge product is defined as

Generalized analytic functions

[edit]

If a function is holomorphic in some complex region , then

in. For generalized analytic functions, this equation is replaced by

in a region , where is the complex conjugate of and and are functions of and . [11]

Generalized analytic functions have applications in differential geometry, in solving certain type of multidimensional nonlinear partial differential equations and multidimensional inverse scattering.[12]

Applications to integrability theory

[edit]

Riemann–Hilbert problems have applications to several related classes of problems.

A. Integrable models
The inverse scattering or inverse spectral problem associated to the Cauchy problems for 1+1 dimensional partial differential equations on the line, or to periodic problems, or even to initial-boundary value problems (Fokas (2002)), can be stated as a Riemann–Hilbert problem. Likewise the inverse monodromy problem for Painlevé equations can be stated as a Riemann–Hilbert problem.
B. Orthogonal polynomials, Random matrices
Given a weight on a contour, the corresponding orthogonal polynomials can be computed via the solution of a Riemann–Hilbert factorization problem (Fokas, Its & Kitaev (1992)). Furthermore, the distribution of eigenvalues of random matrices in several classical ensembles is reduced to computations involving orthogonal polynomials (see e.g. Deift (2000)).
C. Combinatorial probability
The most celebrated example is the theorem of Baik, Deift & Johansson (1999) on the distribution of the length of the longest increasing subsequence of a random permutation. Together with the study of B above, it is one of the original rigorous investigations of so-called "integrable probability". But the connection between the theory of integrability and various classical ensembles of random matrices goes back to the work of Dyson (see e.g. Dyson (1976)).
D. Connection to Donaldson-Thomas theory
The work of Bridgeland Bridgeland (2019) studies a class of Riemann-Hilbert problems coming from Donaldson-Thomas theory and makes connections with Gromov-Witten theory and exact WKB.

The numerical analysis of Riemann–Hilbert problems can provide an effective way for numerically solving integrable PDEs (see e.g. Trogdon & Olver (2016)).

Use for asymptotics

[edit]

In particular, Riemann–Hilbert factorization problems are used to extract asymptotic values for the three problems above (say, as time goes to infinity, or as the dispersion coefficient goes to zero, or as the polynomial degree goes to infinity, or as the size of the permutation goes to infinity). There exists a method for extracting the asymptotic behavior of solutions of Riemann–Hilbert problems, analogous to the method of stationary phase and the method of steepest descent applicable to exponential integrals.

By analogy with the classical asymptotic methods, one "deforms" Riemann–Hilbert problems which are not explicitly solvable to problems that are. The so-called "nonlinear" method of stationary phase is due to Deift & Zhou (1993), expanding on a previous idea by Its (1982) and Manakov (1974) and using technical background results from Beals & Coifman (1984) and Zhou (1989). A crucial ingredient of the Deift–Zhou analysis is the asymptotic analysis of singular integrals on contours. The relevant kernel is the standard Cauchy kernel (see Gakhov (2001); also cf. the scalar example below).

An essential extension of the nonlinear method of stationary phase has been the introduction of the so-called finite gap g-function transformation by Deift, Venakides & Zhou (1997), which has been crucial in most applications. This was inspired by work of Lax, Levermore and Venakides, who reduced the analysis of the small dispersion limit of the KdV equation to the analysis of a maximization problem for a logarithmic potential under some external field: a variational problem of "electrostatic" type (see Lax & Levermore (1983)). The g-function is the logarithmic transform of the maximizing "equilibrium" measure. The analysis of the small dispersion limit of KdV equation has in fact provided the basis for the analysis of most of the work concerning "real" orthogonal polynomials (i.e. with the orthogonality condition defined on the real line) and Hermitian random matrices.

Perhaps the most sophisticated extension of the theory so far is the one applied to the "non self-adjoint" case, i.e. when the underlying Lax operator (the first component of the Lax pair) is not self-adjoint, by Kamvissis, McLaughlin & Miller (2003). In that case, actual "steepest descent contours" are defined and computed. The corresponding variational problem is a max-min problem: one looks for a contour that minimizes the "equilibrium" measure. The study of the variational problem and the proof of existence of a regular solution, under some conditions on the external field, was done in Kamvissis & Rakhmanov (2005); the contour arising is an "S-curve", as defined and studied in the 1980s by Herbert R. Stahl, Andrei A. Gonchar and Evguenii A Rakhmanov.

An alternative asymptotic analysis of Riemann–Hilbert factorization problems is provided in McLaughlin & Miller (2006), especially convenient when jump matrices do not have analytic extensions. Their method is based on the analysis of d-bar problems, rather than the asymptotic analysis of singular integrals on contours. An alternative way of dealing with jump matrices with no analytic extensions was introduced in Varzugin (1996).

Another extension of the theory appears in Kamvissis & Teschl (2012) where the underlying space of the Riemann–Hilbert problem is a compact hyperelliptic Riemann surface. The correct factorization problem is no more holomorphic, but rather meromorphic, by reason of the Riemann–Roch theorem. The related singular kernel is not the usual Cauchy kernel, but rather a more general kernel involving meromorphic differentials defined naturally on the surface (see e.g. the appendix in Kamvissis & Teschl (2012)). The Riemann–Hilbert problem deformation theory is applied to the problem of stability of the infinite periodic Toda lattice under a "short range" perturbation (for example a perturbation of a finite number of particles).

Most Riemann–Hilbert factorization problems studied in the literature are 2-dimensional, i.e., the unknown matrices are of dimension 2. Higher-dimensional problems have been studied by Arno Kuijlaars and collaborators, see e.g. Kuijlaars & López (2015).

See also

[edit]

Notes

[edit]
  • ^ Bitsadze 2001.
  • ^ Pandey 1996.
  • ^ Ablowitz & Fokas 2003, pp. 514.
  • ^ Pandey 1996, §2.2.
  • ^ Noble 1958, §4.2.
  • ^ Ablowitz & Fokas 2003, §7.5.
  • ^ Ablowitz & Fokas 2003, p. 598.
  • ^ Ablowitz & Fokas 2003, p. 600.
  • ^ Vekua 2014.
  • ^ Ablowitz & Fokas 2003, p. 601.
  • References

    [edit]
    Retrieved from "https://en.wikipedia.org/w/index.php?title=Riemann–Hilbert_problem&oldid=1233647929"

    Categories: 
    Complex analysis
    Exactly solvable models
    Integrable systems
    Solitons
    Scattering theory
    Harmonic analysis
    Microlocal analysis
    Ordinary differential equations
    Partial differential equations
    Mathematical problems
    Bernhard Riemann
    David Hilbert
    Hidden categories: 
    Articles with FAST identifiers
    Articles with BNF identifiers
    Articles with BNFdata identifiers
    Articles with GND identifiers
    Articles with J9U identifiers
    Articles with LCCN identifiers
    Articles with SUDOC identifiers
     



    This page was last edited on 10 July 2024, at 05:44 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki