Neumann–Poincaré Operator
   HOME

TheInfoList



OR:

In
mathematics Mathematics is an area of knowledge that includes the topics of numbers, formulas and related structures, shapes and the spaces in which they are contained, and quantities and their changes. These topics are represented in modern mathematics ...
, the Neumann–Poincaré operator or Poincaré–Neumann operator, named after
Carl Neumann Carl Gottfried Neumann (also Karl; 7 May 1832 – 27 March 1925) was a German mathematician. Biography Neumann was born in Königsberg, Prussia, as the son of the mineralogist, physicist and mathematician Franz Ernst Neumann (1798–1895), who w ...
and
Henri Poincaré Jules Henri Poincaré ( S: stress final syllable ; 29 April 1854 – 17 July 1912) was a French mathematician, theoretical physicist, engineer, and philosopher of science. He is often described as a polymath, and in mathematics as "The ...
, is a non-self-adjoint
compact operator In functional analysis, a branch of mathematics, a compact operator is a linear operator T: X \to Y, where X,Y are normed vector spaces, with the property that T maps bounded subsets of X to relatively compact subsets of Y (subsets with compact ...
introduced by Poincaré to solve
boundary value problem In mathematics, in the field of differential equations, a boundary value problem is a differential equation together with a set of additional constraints, called the boundary conditions. A solution to a boundary value problem is a solution to t ...
s for the Laplacian on bounded domains in Euclidean space. Within the language of
potential theory In mathematics and mathematical physics, potential theory is the study of harmonic functions. The term "potential theory" was coined in 19th-century physics when it was realized that two fundamental forces of nature known at the time, namely gravi ...
it reduces the
partial differential equation In mathematics, a partial differential equation (PDE) is an equation which imposes relations between the various partial derivatives of a Multivariable calculus, multivariable function. The function is often thought of as an "unknown" to be sol ...
to an
integral equation In mathematics, integral equations are equations in which an unknown Function (mathematics), function appears under an integral sign. In mathematical notation, integral equations may thus be expressed as being of the form: f(x_1,x_2,x_3,...,x_n ; ...
on the boundary to which the theory of
Fredholm operator In mathematics, Fredholm operators are certain operators that arise in the Fredholm theory of integral equations. They are named in honour of Erik Ivar Fredholm. By definition, a Fredholm operator is a bounded linear operator ''T'' : ''X ...
s 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 Complex analysis, traditionally known as the theory of functions of a complex variable, is the branch of mathematical analysis that investigates functions of complex numbers. It is helpful in many branches of mathematics, including algebraic ...
, the conjugate
Beurling transform In mathematics, singular integral operators of convolution type are the singular integral operators that arise on R''n'' and T''n'' through convolution by distributions; equivalently they are the singular integral operators that commute with transl ...
or complex
Hilbert transform In mathematics and in signal processing, the Hilbert transform is a specific linear operator that takes a function, of a real variable and produces another function of a real variable . This linear operator is given by convolution with the functi ...
and the Fredholm eigenvalues of bounded planar domains.


Dirichlet and Neumann problems

Green's theorem In vector calculus, Green's theorem relates a line integral around a simple closed curve to a double integral over the plane region bounded by . It is the two-dimensional special case of Stokes' theorem. Theorem Let be a positively orient ...
for a bounded region Ω in the plane with smooth boundary ∂Ω states that :\displaystyle One direct way to prove this is as follows. By subtraction, it is sufficient to prove the theorem for a region bounded by a simple smooth curve. Any such is diffeomorphic to the closed
unit disk In mathematics, the open unit disk (or disc) around ''P'' (where ''P'' is a given point in the plane), is the set of points whose distance from ''P'' is less than 1: :D_1(P) = \.\, The closed unit disk around ''P'' is the set of points whose di ...
. By change of variables it is enough to prove the result there. Separating the ''A'' and ''B'' terms, the right hand side can be written as a double integral starting in the ''x'' or ''y'' direction, to which the
fundamental theorem of calculus The fundamental theorem of calculus is a theorem that links the concept of differentiating a function (calculating its slopes, or rate of change at each time) with the concept of integrating a function (calculating the area under its graph, or ...
can be applied. This converts the integral over the disk into the integral over its boundary. Let Ω be a region bounded by a simple closed curve. Given a smooth function ''f'' on the closure of Ω its normal derivative ∂''n''''f'' at a boundary point is the directional derivative in the direction of the outward pointing normal vector. Applying Green's theorem with ''A'' = ''v''''x'' ''u'' and ''B'' = ''v''''y'' ''u'' gives the first of
Green's identities In mathematics, Green's identities are a set of three identities in vector calculus relating the bulk with the boundary of a region on which differential operators act. They are named after the mathematician George Green, who discovered Green's ...
: :\displaystyle where the Laplacian Δ is given by :\displaystyle \Delta = -\partial^2_x -\partial_y^2. Swapping ''u'' and ''v'' and subtracting gives the second of Green's identities: :\displaystyle If now ''u'' is harmonic in Ω and ''v'' = 1, then this identity implies that :\displaystyle so the integral of the normal derivative of a harmonic function on the boundary of a region always vanishes. A similar argument shows that the average of a harmonic function on the boundary of a disk equals its value at the centre. Translating the disk can be taken to be centred at 0. Green's identity can be applied to an annulus formed of the boundary of the disk and a small circle centred on 0 with ''v'' = ''z''2: it follows that the average is independent of the circle. It tends to the value at its value at 0 as the radius of the smaller circle decreases. This result also follows easily using Fourier series and the
Poisson integral In mathematics, and specifically in potential theory, the Poisson kernel is an integral kernel, used for solving the two-dimensional Laplace equation, given Dirichlet boundary conditions on the unit disk. The kernel can be understood as the deriv ...
. For continuous functions ''f'' on the whole plane which are smooth in Ω and the complementary region Ω''c'', the first derivative can have a jump across the boundary of Ω. The value of the normal derivative at a boundary point can be computed from inside or outside Ω. The interior normal derivative will be denoted by ∂''n''− and the exterior normal derivative by ∂''n''+. With this terminology the four basic problems of classical potential theory are as follows: *Interior Dirichlet problem: ∆''u'' = 0 in Ω, ''u'' = ''f'' on ∂Ω *Interior Neumann problem: ∆''u'' = 0 in Ω, ∂''n''− ''u'' = ''f'' on ∂Ω *Exterior Dirichlet problem: ∆''u'' = 0 in Ω''c'', ''u'' = ''f'' on ∂Ω, ''u'' continuous at ∞ *Exterior Neumann problem: ∆''u'' = 0 in Ω''c'', ∂''n''+ ''u'' = ''f'' on ∂Ω, ''u'' continuous at ∞ For the exterior problems the inversion map ''z''−1 takes harmonic functions on Ω''c'' into harmonic functions on the image of Ω''c'' under the inversion map. The transform ''v'' of ''u'' is continuous in a small disc , ''z'', ≤ ''r'' and harmonic everywhere in the interior except possibly 0. Let ''w'' be the harmonic function given by the
Poisson integral In mathematics, and specifically in potential theory, the Poisson kernel is an integral kernel, used for solving the two-dimensional Laplace equation, given Dirichlet boundary conditions on the unit disk. The kernel can be understood as the deriv ...
on , ''z'', ≤ ''r'' with the same boundary value ''g'' as ''v'' on , ''z'', = ''r''. Applying the
maximum principle In the mathematical fields of partial differential equations and geometric analysis, the maximum principle is any of a collection of results and techniques of fundamental importance in the study of elliptic and parabolic differential equations. ...
to ''v'' − ''w'' + ε log , ''z'', on δ ≤ , ''z'', ≤ ''r'', it must be negative for δ small. Hence ''v''(''z'') ≤ ''u''(''z'') for ''z'' ≠ 0. The same argument applies with ''v'' and ''w'' swapped, so ''v'' = ''w'' is harmonic in the disk. Thus the singularity at ∞ is removable. By the maximum principle the interior and exterior Dirichlet problems have unique solutions. For the interior Neumann problem, if a solution ''u'' is harmonic in 0 and its interior normal derivative vanishes, then Green's first identity implies the ''u''''x'' = 0 = ''u''''y'', so that ''u'' is constant. This shows the interior Neumann problem has a unique solution up to adding constants. Applying inversion, the same holds for the external Neumann problem. For both Neumann problems, a necessary condition for a solution to exist is :\displaystyle For the interior Neumann problem, this follows by setting ''v'' = 1 in Green's second identity. For the exterior Neumann problem, the same can be done for the intersection of Ω''c'' and a large disk , ''z'', < ''R'', giving :\displaystyle At ∞ ''u'' is the real part of a holomorphic function ''F'' with :\displaystyle The interior normal derivative on , ''z'', = ''R'' is just the radial derivative ∂''r'', so that for , ''z'', = ''R'' :\displaystyle Hence :\displaystyle so the integral over ∂Ω must vanish. The
fundamental solution In mathematics, a fundamental solution for a linear partial differential operator is a formulation in the language of distribution theory of the older idea of a Green's function (although unlike Green's functions, fundamental solutions do not ad ...
of the Laplacian is given by :\displaystyle ''N''(''z'') = − ''E''(''z'') is called the
Newtonian potential In mathematics, the Newtonian potential or Newton potential is an operator in vector calculus that acts as the inverse to the negative Laplacian, on functions that are smooth and decay rapidly enough at infinity. As such, it is a fundamental object ...
in the plane. Using polar coordinates, it is easy to see that ''E'' is in L''p'' on any closed disk for any finite ''p'' ≥ 1. To say that ''E'' is a fundamental solution of the Laplacian means that for any smooth function φ of compact support :\displaystyle The standard proof uses Green's second identity on the annulus ''r'' ≤ , ''z'', ≤ ''R'' where the support of φ is contained in , ''z'', < ''R''. In fact, since ''E'' is harmonic away from 0, :\displaystyle As ''r'' tends to zero, the first term on the right hand side tends to φ(0) and the second to 0, since ''r'' log ''r'' tends to 0 and the normal derivatives of φ are uniformly bounded. (That both sides are equal even before taking limits follows from the fact that the average of a harmonic function on the boundary of a disk equals it value at the centre, while the integral of its normal derivative vanishes.)


Neumann–Poincaré kernel

The properties of the fundamental solution lead to the following formula for recovering a harmonic function ''u'' in Ω from its boundary values: :\displaystyle where ''K'' is the Neumann−Poincaré kernel :\displaystyle To prove this identity, Green's second identity can be applied to Ω with a small disk centred on ''z'' removed. This reduces to showing that the identity holds in the limit for a small disk centred on ''z'' shrinking in size. Translating, it can be assumed that ''z'' = 0 and the identity becomes :\displaystyle which was proved above. A similar formula holds for functions harmonic in Ω''c'': :\displaystyle The signs are reversed because of the direction of the normal derivative. In two dimensions the Neumann–Poincaré kernel ''K''(''z'',''w'') has the remarkable property that it restricts to a smooth function on ∂Ω × ∂Ω. It is ''a priori'' only defined as a smooth function off the diagonal but it admit a (unique) smooth extension to the diagonal. Using vector notation v(''t'') = (''x''(''t''), ''y''(''t'')) to parametrize the boundary curve by arc length, the following classical formulas hold: :\displaystyle Thus the unit tangent vector t(''t'') at ''t'' is the velocity vector :\displaystyle so the oriented unit normal n(''t'') is :\displaystyle The constant relating the acceleration vector to the normal vector is the
curvature In mathematics, curvature is any of several strongly related concepts in geometry. Intuitively, the curvature is the amount by which a curve deviates from being a straight line, or a surface deviates from being a plane. For curves, the canonic ...
of the curve: :\displaystyle Thus the curvature is given by :\displaystyle There are two further formulas of Frenet: :\displaystyle The Neumann–Poincaré kernel is given by the formula :\displaystyle For ''s'' ≠ ''t'', set :\displaystyle The function :\displaystyle is smooth and nowhere vanishing with ''a''(''s'',''s'') = ''L''2 if the length of the curve is 2''L''. Similarly the function :\displaystyle is smooth. In fact writing ''s'' = ''t'' + ''h'', :\displaystyle so that :\displaystyle On the diagonal ''b''(''t'',''t'') = κ ''L''2 / 2. Since ''k'' is proportional to ''b'' / ''a'', it is also smooth. Its diagonal values are given by the formula :\displaystyle Another expression for ''k''(''s'',''t'') is as follows: :\displaystyle where ''z''(''t'') = ''x''(''t'') + i ''y''(''t'') is the boundary curve parametrized by arc length. This follows from the identity :\displaystyle and the
Cauchy–Riemann equations In the field of complex analysis in mathematics, the Cauchy–Riemann equations, named after Augustin Cauchy and Bernhard Riemann, consist of a system of two partial differential equations which, together with certain continuity and differ ...
which can be used to express the normal derivative in terms of the tangential derivative, and :\displaystyle so in the direction normal to the boundary curve ''K'' is discontinuous at the boundary.


Double layer potentials

The
double layer potential In potential theory, an area of mathematics, a double layer potential is a solution of Laplace's equation corresponding to the electrostatic or magnetic potential associated to a dipole distribution on a closed surface ''S'' in three-dimensions. ...
with moment φ in C(∂Ω) is defined on the complement of ∂Ω as :\displaystyle It is a continuous function on the complement. Since the restriction of ''K'' extends to a smooth function on ∂Ω × ∂Ω, ''D''(φ) can also be defined on ∂Ω. However like the Neumann–Poincaré kernel it will have discontinuities at the boundary. These are jump discontinuities. If φ is real, then the double layer potential is just the real part of a Cauchy integral: :\displaystyle The simplest case is when φ is identically 1 on ∂Ω. In this case ''D''(1) equals *1 on Ω, by the vanishing of the integral of the normal derivative on the boundary region bounded by ∂Ω and a small disk centred on ''z''; so the integral over the ∂Ω equals the average of the function 1 on the boundary of small disk and hence equals 1. (This integral and the one for Ω''c'' can also be calculated using
Cauchy's integral theorem In mathematics, the Cauchy integral theorem (also known as the Cauchy–Goursat theorem) in complex analysis, named after Augustin-Louis Cauchy (and Édouard Goursat), is an important statement about line integrals for holomorphic functions in t ...
.) *0 on Ω''c'', because it is the integral of a normal derivative of a harmonic function. *1/2 on ∂Ω, since ::\displaystyle By definition the Neumann–Poincaré operator ''T''''K'' is the operator on L2(∂Ω) given by the kernel ''K''(''z'',''w''). It is a
Hilbert–Schmidt operator In mathematics, a Hilbert–Schmidt operator, named after David Hilbert and Erhard Schmidt, is a bounded operator A \colon H \to H that acts on a Hilbert space H and has finite Hilbert–Schmidt norm \, A\, ^2_ \ \stackrel\ \sum_ \, Ae_i\, ^2_ ...
since the kernel is continuous. It takes values in C(∂Ω) since the kernel is smooth. The third computation above is equivalent to the statement that the constant function 1 is an eigenfunction of ''T''''K'' with eigenvalue 1/2. To establish jump formulas for more general functions, it is necessary to check that the integrals for ''D''(1) are uniformly absolutely convergent, i.e. that there is a uniform finite bound ''C'' such that :\displaystyle for all ''z'' not in the boundary. It is enough to check this for points in a tubular neighbourhood of the boundary. Any such point u lies on a normal through a unique point, v(0) say, on the curve and it is enough to look at the contribution to the integral from points v(''t'') with ''t'' in a small interval around 0. Writing :\displaystyle it follows that :\displaystyle So for ''t'' sufficiently small :\displaystyle for some constant ''C''1. (The first inequality gives an approximate version of
Pythagoras' theorem In mathematics, the Pythagorean theorem or Pythagoras' theorem is a fundamental relation in Euclidean geometry between the three sides of a right triangle. It states that the area of the square whose side is the hypotenuse (the side opposite ...
in the tubular neighbourhood.) Hence :\displaystyle Uniform boundedness follows because the first term has a finite integral independent of λ: :\displaystyle The bound above can be used to prove that if the moment φ vanishes at a boundary point ''z'' then its double layer potential ''D''(φ) is continuous at ''z''. More generally if φ''n'' tends uniformly to φ, then ''D''(φ''n'')(''z''''n'') converges to ''D''(φ)(''z''). In fact suppose that , φ(''w''), ≤ ε if , ''w'' - ''z'', ≤ δ. Taking ''z''''n'' tending to ''z'' :, D(\varphi_n)(z_n) - D(\varphi)(z), \le \int_ , K(z_n,w)-K(z,w), , \varphi(w), \,, dw, +\int_ (, K(z_n,w), +, K(z,w), )\cdot, \varphi(w), \, , dw, ::::::::::::: + \int_ , K(z_n,w), \cdot , \varphi_n(w) -\varphi(w), \, , dw, . The first integrand tends uniformly to 0 so the integral tends to 0. The second integral is bounded above by 2ε''C''. The third integral is bounded by C times the supremum norm of φ''n'' − φ. Hence ''D''(φ)(''z''''n'') tends to ''D''(φ)(''z''). JUMP FORMULAS. If φ is a continuous function on ∂Ω, the restrictions of its double layer potential ''u'' = ''D''φ to Ω and Ω''c'' extend uniquely to continuous functions on their closures. Let ''u'' and ''u''+ be the resulting continuous functions on ∂Ω. Then :\displaystyle In particular :\displaystyle In fact the expressions for ''u''± are continuous, so it is enough to show that the if ''z''''n'' tends to a boundary point ''z'' with ''z''''n'' in Ω or Ω''c'' then ''u''(''z''''n'') tends to the expression for ''u''±(''z''). If ''z''''n'' lie in Ω or Ω''c'' then :\displaystyle where ψ(''w'') = φ(''w'') − φ(''z''). The right hand side tends to zero since ψ vanishes at ''z''.


Single layer potentials

The
single layer potential In mathematics, the Newtonian potential or Newton potential is an operator in vector calculus that acts as the inverse to the negative Laplacian, on functions that are smooth and decay rapidly enough at infinity. As such, it is a fundamental obje ...
with moment φ in C(∂Ω) is defined on C as :\displaystyle where ''N'' is the
Newtonian potential In mathematics, the Newtonian potential or Newton potential is an operator in vector calculus that acts as the inverse to the negative Laplacian, on functions that are smooth and decay rapidly enough at infinity. As such, it is a fundamental object ...
:\displaystyle The single layer potential is harmonic off ∂Ω. Since :\displaystyle and the first integrand tends uniformly to 0 as , ''z'', tends to infinity, the single layer potential is harmonic at infinity if and only if ∫ φ = 0. The single layer potential is continuous on C. In fact continuity off ∂Ω is clear. If ''z''''n'' tends to ''z'' with ''z'' in ∂Ω, then :, S(\varphi)(z)-S(\varphi)(z_n), \le\int_ , \log , z-w, -\log , z_n-w, , \,, \varphi(w), \, , dw, + \, \varphi\, _\infty \int_ (, \log, z-w, , + , \log, z_n-w, , )\, , dw, . The first integrand tends uniformly to 0 on , ''w'' - ''z'', ≥ ε. For ''n'' sufficiently large the last integral is bounded by :\displaystyle which tends to 0 as ε tends to 0, by the
Cauchy–Schwarz inequality The Cauchy–Schwarz inequality (also called Cauchy–Bunyakovsky–Schwarz inequality) is considered one of the most important and widely used inequalities in mathematics. The inequality for sums was published by . The corresponding inequality fo ...
since the integrand is square integrable. The same argument shows that ''S'' = ''T''''N'' defines a bounded operator on C(∂Ω): :\displaystyle for φ in C(∂Ω). Although the single layer potentials are continuous, their first derivatives have a jump discontinuity across ∂Ω. On the tubular neighbourhood of ∂Ω, the normal derivative is defined by :\displaystyle It follows that :\displaystyle so it is given by the adjoint kernel of ''K'': :\displaystyle The kernel ''K''* extends naturally to a smooth function on ∂Ω × ∂Ω and the operator ''T''''K''* is the adjoint of ''T''''K'' on L2(∂Ω). JUMP FORMULAS. If φ is a continuous function on ∂Ω, the normal derivatives of the single layer potential ''u'' = ''S''(φ) on Ω and Ω''c'' near ∂Ω extend continuously to the closure of both regions, defining continuous functions ∂''n''- ''u'' and ∂''n''+ ''u'' on ∂Ω. Then :\displaystyle In particular :\displaystyle In fact let ''v'' = ''D''(φ) be the double layer potential with moment φ. On ∂Ω set :\displaystyle and on the complement of ∂Ω in a tubular neighbourhood set :\displaystyle Then ''f'' is continuous on the tubular neighbourhood. In fact, by definition is continuous on ∂Ω and its complement, so it suffices to that ''f''(''z''''n'') tends to ''f''(''z'') whenever ''z''''n'' is a sequence of points in the complement tending to a boundary point ''z''. In this case :\displaystyle The integrand tends uniformly to 0 for , ''w'' − ''z'', ≥ δ, so the first integral tends to 0. To show the second integral is small for δ small, it suffices to show that the integrand is uniformly bounded. This follows because, if ζ''n'' is the point on ∂Ω with normal containing ''z''''n'', then :\displaystyle The first term the last product uniformly bounded because of the smoothness of the Gauss map n(''t''). The second is uniformly bounded because of the approximate version of Pythagoras' theorem: :\displaystyle Continuity of ''f'' implies that on ∂Ω :\displaystyle which gives the jump formulas.


Derivatives of layer potentials

If the moment φ is smooth, the derivatives of the single and double layer potentials on Ω and Ω''c'' extend continuously to their closures. As usual the
gradient In vector calculus, the gradient of a scalar-valued differentiable function of several variables is the vector field (or vector-valued function) \nabla f whose value at a point p is the "direction and rate of fastest increase". If the gradi ...
of a function ''f'' defined on an open set in R2 is defined by :\displaystyle Set :\displaystyle If the moment φ is smooth, then :\displaystyle In fact :\displaystyle so that :\displaystyle Moreover :\displaystyle The second relation can be rewritten by substituting in from the first relation: :\displaystyle Regularity of layer potentials. As a consequence of these relations, successive derivatives can all be expressed in terms of single and double layer potentials of smooth moments on the boundary. Since the layer potentials on Ω and Ω''c'' have continuous limits on the boundary it follows that they define smooth functions on the closures of Ω and Ω''c''. Continuity of normal derivatives of double layer potentials. Just as the single layer potentials are continuous at the boundary with a jump in the normal derivative, so the double layer potentials have a jump across the boundary while their normal derivatives are continuous. In fact from the formula above :\displaystyle If ''s''''n'' tends to ''s'' and λ''n'' tends to 0, the first term tends to ''T''''K''(''v''(s)) since the moments tend uniformly to a moment vanishing at ''t'' = ''s''; the second term is continuous because it is a single layer potential.


Solution of Dirichlet and Neumann problems

The following properties of ''T'' = ''T''''K'' are required to solve the boundary value problem: * 1/2 is not a generalized eigenvalue of ''T''''K'' or ''T''''K''*; it has multiplicity one. * −1/2 is not an eigenvalue of ''T''''K'' or ''T''''K''*. In fact since ''a'' I + ''T'' is a Fredholm operator of index 0, it and its adjoint have kernels of equal dimension. The same applies to any power of this operator. So it suffices to verify each of the statements for either ''T'' or ''T''*. To check that ''T'' has no generalized eigenvectors with eigenvalue 1/2 it suffices to show that :\displaystyle has no solutions. The definition of the double layer potential shows that it vanishes at ∞, so that it is harmonic at ∞. The equation above shows that if ''u'' = ''D''(φ) then ''u''+ = 1. On the other hand, applying the inversion map gives a contradiction; for it would produce a harmonic map in bounded region vanishing at an interior point with boundary value 1, which contradicts the fact that 1 is the only harmonic map with boundary value 1. If the eigenvalue 1/2 has multiplicity greater than 1, there is a moment φ such that ''T''*φ = φ/2 and ∫ φ = 0. It follows that if ''u'' = ''S''(φ) then ∂''n''− ''u'' = 0. By uniqueness ''u'' is constant on Ω. Since ''u'' is continuous on R2 ∪ ∞ and is harmonic at ∞ (since ∫ φ = 0) and constant on ∂Ω, it must be zero. Hence φ = ∂''n''+ ''u'' − ∂''n''− ''u'' = 0. Thus the eigenspace is one-dimensional and the eigenfunction ψ can be normalized so that ''S''(ψ) = 1 on ∂Ω. In general if :\displaystyle then :\displaystyle since :\int f=(f,1)=((T_K^* +)\varphi,1)=(\varphi,(T_K + )1)=(\varphi,1)=\int \varphi. If φ satisfies :\displaystyle it follows that ∫ φ = 0 and so ''u'' = ''S''(φ) is harmonic at infinity. By the jump formulas, ∂''n''-''u'' = 0. By uniqueness ''u'' is constant on Ω. By continuity it is constant on ∂Ω. Since it is harmonic on Ω''c'' and vanishes at infinity, it must vanish identically. As above this forces φ = 0. These results on the eigenvalues of ''T''''K'' lead to the following conclusions about the four boundary value problems: *there is always a unique solution to the interior and exterior Dirichlet problems; *there is a solution to the interior and exterior Neumann problems if and only if ∫ ''f'' = 0; the solution is unique up to a constant for the interior Neumann problem and unique for the exterior problem; *the solution is smooth on the closure of the domain if the boundary data is smooth. The solution is obtained as follows: *Interior Dirichlet problem. Let φ be the unique solution of ''T''''K''φ + φ/2 = ''f''. Then ''u'' = ''D''(φ) gives the solution of the Dirichlet problem in Ω by the jump formula. *Exterior Dirichlet problem. Since 1 is not in the range of ''T''''K'' − ½''I'', ''f'' can be written uniquely as ''f'' = ''T''''K''φ − φ/2 + λ where φ is unique up to a constant. Then ''u'' = ''D''(φ) + λ''S''(ψ) gives the solution of the Dirichlet problem in Ω''c'' by the jump formula. *Interior Neumann problem. The condition (''f'',1) = 0 implies that ''f'' = ''T''''K''*φ − φ/2 can be solved. Then ''u'' = ''S''(φ) gives the solution of the Neumann problem in Ω by the jump formula. *Exterior Neumann problem. Let φ be the unique solution of ''T''''K''*φ + φ/2 = ''f''. Then ''u'' = ''S''(φ) gives the solution of the Neumann problem in Ω by the jump formula. The smoothness of the solution follows from the regularity of single and double layer potentials.


Calderón projector

There is another consequence of the laws governing the derivatives, which completes the symmetry of the jump relations, is that normal derivative of the double layer potential has no jump across the boundary, i.e. it has a continuous extension to a tubular neighbourhood of the boundary given by :\displaystyle ''H'' is called a hypersingular operator. Although it takes smooth functions to smooth functions, it is not a bounded operator on L2(∂Ω). In fact it is a
pseudodifferential operator In mathematical analysis a pseudo-differential operator is an extension of the concept of differential operator. Pseudo-differential operators are used extensively in the theory of partial differential equations and quantum field theory, e.g. in m ...
of order 1, so does define a bounded operator between Sobolev spaces on ∂Ω, decreasing the order by 1. It allows a 2 × 2 matrix of operators to be defined by :\displaystyle The matrix satisfies ''C''2 = ''C'', so is an
idempotent Idempotence (, ) is the property of certain operation (mathematics), operations in mathematics and computer science whereby they can be applied multiple times without changing the result beyond the initial application. The concept of idempotence ...
, called the Calderón projector. This identity is equivalent the following classical relations, the first of which is the symmetrization relation of Plemelj: :\displaystyle The operators ''T'' and ''S'' are pseudodifferential operators of order −1. The relations above follow by considering ''u'' = ''S''(φ). It has boundary value ''S''φ) and normal derivative ''T''* φ − φ/2. Hence in Ω :\displaystyle Taking the boundary values of both sides and their normal derivative yields 2 equations. Two more result by considering ''D''(Ψ); these imply the relations for the Calderón projector.


Fredholm eigenvalues

The non-zero
eigenvalue In linear algebra, an eigenvector () or characteristic vector of a linear transformation is a nonzero vector that changes at most by a scalar factor when that linear transformation is applied to it. The corresponding eigenvalue, often denoted b ...
s of the Neumann–Poincaré operator ''T''''K'' are called the Fredholm eigenvalues of the region Ω. Since ''T''''K'' is a
compact operator In functional analysis, a branch of mathematics, a compact operator is a linear operator T: X \to Y, where X,Y are normed vector spaces, with the property that T maps bounded subsets of X to relatively compact subsets of Y (subsets with compact ...
, indeed a
Hilbert–Schmidt operator In mathematics, a Hilbert–Schmidt operator, named after David Hilbert and Erhard Schmidt, is a bounded operator A \colon H \to H that acts on a Hilbert space H and has finite Hilbert–Schmidt norm \, A\, ^2_ \ \stackrel\ \sum_ \, Ae_i\, ^2_ ...
, all non-zero elements in its spectrum are eigenvalues of finite multiplicity by the general theory of
Fredholm operator In mathematics, Fredholm operators are certain operators that arise in the Fredholm theory of integral equations. They are named in honour of Erik Ivar Fredholm. By definition, a Fredholm operator is a bounded linear operator ''T'' : ''X ...
s. The solution of the boundary value requires knowledge of the spectrum at ± 1/2, namely that the constant function gives an eigenfunction with eigenvalue 1/2 and multiplicity one; that there are no corresponding generalized eigenfunctions with eigenvalue 1/2; and that -1/2 is not an eigenvalue. proved that all non-zero eigenvalues are real and contained in the interval (-1/2,1/2]. proved that the other non-zero eigenvalues have an important symmetry property, namely that if λ is an eigenvalue with 0 < , λ, < 1/2, then so is –λ, with the same multiplicity. Plemelj also showed that ''T'' = ''T''''K'' is a symmetrizable compact operator, so that, even though it is not self-adjoint, it shares many of the properties of self-adjoint operators. In particular there are no generalized eigenfunctions for non-zero eigenvalues and there is a
variational principle In science and especially in mathematical studies, a variational principle is one that enables a problem to be solved using calculus of variations, which concerns finding functions that optimize the values of quantities that depend on those func ...
similar to the
minimax principle Minimax (sometimes MinMax, MM or saddle point) is a decision rule used in artificial intelligence, decision theory, game theory, statistics, and philosophy for ''mini''mizing the possible loss for a worst case (''max''imum loss) scenario. When de ...
for determining the non-zero eigenvalues. If λ ≠ 1/2 is an eigenvalue of ''T''''K''* then λ is real, with λ ≠ ± 1/2. Let φ be a corresponding eigenfunction and, following Plemelj, set ''u'' = ''S''(φ). Then the jump formulas imply that :\displaystyle and hence that :\displaystyle Since ∫ φ = 0, ''u'' is harmonic at ∞. So by Green's theorem :\displaystyle If both the integrals vanish then ''u'' is constant on Ω and Ω''c''. Since it is continuous and vanishes at ∞, it must therefore be identically 0, contradicting φ = ∂''n''+ - ∂''n''−. So both integrals are strictly positive and hence λ must lie in (−½,½). Let φ be an eigenfunction of ''T''''K''* with real eigenvalue λ satisfying 0 < , λ, < 1/2. If ''u'' = ''S''(φ), then on ∂Ω :\displaystyle This process can be reversed. Let ''u'' be a continuous function on R2 ∪ ∞ which is harmonic on Ω and Ω''c'' ∪ ∞ and such that the derivatives of ''u'' on Ω and Ω''c'' extend continuously to their closures. Suppose that :\displaystyle Let ψ be the restriction of ''u'' to ∂Ω. Then :\displaystyle The jump formulas for the boundary values and normal derivatives give :\displaystyle and :\displaystyle It follows that :\displaystyle so that ψ and φ are eigenfunctions of ''T'' and ''T''* with eigenvalue λ. Let ''u'' be a real harmonic function on Ω extending to a smooth function on its closure. The
harmonic conjugate In mathematics, a real-valued function u(x,y) defined on a connected open set \Omega \subset \R^2 is said to have a conjugate (function) v(x,y) if and only if they are respectively the real and imaginary parts of a holomorphic function f(z) of th ...
''v'' of ''u'' is the unique real function on Ω such that ''u'' + ''i'' ''v'' is holomorphic. As such it must satisfy the
Cauchy–Riemann equations In the field of complex analysis in mathematics, the Cauchy–Riemann equations, named after Augustin Cauchy and Bernhard Riemann, consist of a system of two partial differential equations which, together with certain continuity and differ ...
: :\displaystyle If ''a'' is a point in Ω, a solution is given by :\displaystyle where the integral is taken over any path in the closure of Ω. It is easily verified that ''v''''x'' and ''v''''y'' exist and are given by the corresponding derivatives of ''u''. Thus ''v'' is a smooth function on the closure of Ω, vanishing at 0. By the Cauchy-Riemann equations, ''f'' = ''u'' + ''i'' ''v'' is smooth on the closure of Ω, holomorphic on Ω and ''f''(a) = 0. Using the inversion map, the same result holds for a harmonic function in Ω''c'' harmonic at ∞. It has a harmonic conjugate ''v'' such that ''f'' = ''u'' + ''i'' ''v'' extends smoothly to the boundary and ''f'' is holomorphic on Ω ∪ ∞. Adjusting ''v'' by a constant it can be assumed that ''f''(∞) = 0. Following , let φ be an eigenfunction of ''T''''K''* with real eigenvalue λ satisfying 0 < , λ, < 1/2. Let ''u'' = ''S''(φ) and let ''v''± be the harmonic conjugates of ''u''± in Ω and Ω''c''. Since on ∂Ω :\displaystyle the Cauchy-Riemann equations give on ∂Ω :\displaystyle Now define :\displaystyle Thus ''U'' is continuous on R2 and :\displaystyle It follows that −λ is an eigenvalue of ''T''. Since −''u'' is the harmonic conjugate of ''v'', the process of taking harmonic conjugates is one-one, so the multiplicity of −λ as an eigenvalue is the same as that of λ. By Green's theorem :\displaystyle Adding the two integrals and using the jump relations for the single layer potential, it follows that :\displaystyle Thus :\displaystyle This shows that the operator ''S'' is self-adjoint and non-negative on L2(∂Ω). The image of ''S'' is dense (or equivalently it has zero kernel). In fact the relation ''SH'' = ¼ ''I'' - ''T''2 =(½ ''I'' – ''T'') (½ ''I'' + ''T'') shows that the closure of the image of ''S'' contains the image of ½ ''I'' – ''T'', which has codimension 1. Its orthogonal complement is given by the kernel of ''T'' – ½ ''I'', i.e. the eigenfunction ψ such that ''T''*ψ = ½ ψ. On the other hand ''ST''=''T''* ''S''. If the closure of the image is not the whole of L2(∂Ω) then necessarily ''S''ψ = 0. Hence ''S''{ψ) is constant. But then ψ = ∂''n''+''S''(ψ) – ∂''n''−''S''(ψ) = 0, a contradiction. Since ''S'' is strictly positive and ''T'' satisfies the Plemelj symmetrization relation ''ST''* = ''TS'', the operator ''T''* is a symmetrizable compact operator. The operator ''S'' defines a new inner product on L2(∂Ω): :\displaystyle{(f,g)_S=(Sf,g).} The operator ''T''* is formally self-adjoint with respect to this inner product and by general theory its restriction is bounded and it defines a self-adjoint Hilbert–Schmidt operator on the Hilbert space completion. Since ''T''* is formally self-adjoint on this inner product space, it follows immediately that any generalized eigenfunction of ''T''* must already be an eigenfunction. By Fredholm theory, the same is true for ''T''. By general theory the kernel of ''T'' and its non-zero eigenspaces span a dense subspace of L2(∂Ω). The
Fredholm determinant In mathematics, the Fredholm determinant is a complex-valued function which generalizes the determinant of a finite dimensional linear operator. It is defined for bounded operators on a Hilbert space which differ from the identity operator by a ...
is defined by :\displaystyle{\Delta(z)=\det (I -zT^2).} It can be expressed in terms of the Fredholm eigenvalues λ''n'' with modulus less than 1/2, counted with multiplicity, as :\displaystyle{\Delta(z)=(1-z/4)\cdot \prod_{n\ge 1} (1-z\lambda_n^2).}


Complex Hilbert transform

Now define the complex Hilbert transform or conjugate Beurling transform ''T''''c'' on L2(C) by :\displaystyle{T_c f(w)=\lim_{\varepsilon\rightarrow 0} - {1\over \pi } \iint_{, z-w, \ge \varepsilon} {\overline{f(z)} \over (z-w)^2}\, dx\,dy.} This is a conjugate-linear isometric involution. It commutes with ∂ so carries A2(Ω) ⊕ A2''c'') onto itself. The compression of ''T''''c'' to A2(Ω) is denoted ''T''Ω. 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 ''T''Ω becomes the conjugate-linear singular integral operator with kernel : K_F(z,w)={F^\prime(z)F^\prime(w)\over (F(z)-F(w))^2}. It defines a
contraction Contraction may refer to: Linguistics * Contraction (grammar), a shortened word * Poetic contraction, omission of letters for poetic reasons * Elision, omission of sounds ** Syncope (phonology), omission of sounds in a word * Synalepha, merged ...
. On the other hand it can be checked that ''T''''D'' = 0 by computing directly on powers ''n'' using Stokes theorem to transfer the integral to the boundary. It follows that the conjugate-linear operator with kernel :\displaystyle{ {F^\prime(z)F^\prime(w)\over (F(z)-F(w))^2} \,-\,{1\over (z-w)^2 acts as a contraction on the Bergman space of ''D''. It is thus a
Hilbert–Schmidt operator In mathematics, a Hilbert–Schmidt operator, named after David Hilbert and Erhard Schmidt, is a bounded operator A \colon H \to H that acts on a Hilbert space H and has finite Hilbert–Schmidt norm \, A\, ^2_ \ \stackrel\ \sum_ \, Ae_i\, ^2_ ...
. The conjugate-linear operator ''T'' = ''T''Ω satisfies the self-adjointness relation :\displaystyle{(Tu,v)=(Tv,u)} for ''u'', ''v'' in A2(Ω). Thus ''A'' = ''T''2 is a compact self-adjoint linear operator on ''H'' with :\displaystyle{(Au,u)=(Tu,Tu)=\, Tu\, ^2\ge 0,} so that ''A'' is a positive operator. By the spectral theorem for compact self-adjoint operators, there is an orthonormal basis ''u''''n'' of ''H'' consisting of eigenvectors of ''A'': : \displaystyle{Au_n=\mu_n u_n,} where μ''n'' is non-negative by the positivity of ''A''. Hence : \displaystyle{\mu_n=\lambda_n^2} with λ''n'' ≥ 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: : \displaystyle{Tu_n=\lambda_n u_n.}, and :\displaystyle{ T(iu_n)=-\lambda_n iu_n} by conjugate-linearity of ''T''.


Connection with Hilbert transform on a closed curve

The Neumann–Poincaré operator is defined on real functions ''f'' as :\displaystyle{Tf(w)={1\over 2\pi}\int_{\partial\Omega}\partial_n (\log, z-w, ) f(z)={1\over 2}\Re (Hf)(w),} where ''H'' is the
Hilbert transform In mathematics and in signal processing, the Hilbert transform is a specific linear operator that takes a function, of a real variable and produces another function of a real variable . This linear operator is given by convolution with the functi ...
on ∂Ω. Let ''J'' denote complex conjugation. Writing ''h'' = ''f'' + ''ig'', :\displaystyle{2Th =\Re(Hf) + i\Re(Hg)={1\over 2} (Hf +JHf +iHg +iJHg)={1\over 2}(H+JHJ)h} so that :\displaystyle{T={1\over 4} (H +JHJ),} The imaginary part of the Hilbert transform can be used to establish the symmetry properties of the eigenvalues of ''T''''K''. Let :\displaystyle{A={1\over 2} (H+JHJ),\,\,\,\, B={1\over 2i} (H-JHJ),} so that :\displaystyle{H=A+iB.} Then :\displaystyle{AB=-BA,\,\,\,\,A^2-B^2 =I.} The Cauchy idempotent ''E'' satisfies ''E''1 = 1 = ''E''*1. Since ''J''1 = 1, it follows that ''E'' and ''E''* leave invariant L20(∂Ω), the functions orthogonal to constant functions. The same is also true of ''A'' = 2 ''T''''K'' and ''B''. Let ''A''1 and ''B''1 be their restrictions. Since 1 is an eigenvector of ''T''''K'' with eigenvalue 1/2 and multiplicity one and ''T''''K'' + ½ ''I'' is invertible, :\displaystyle{A_1^2 -I =B_1^2} is invertible, so that ''B''1 is invertible. The equation ''A''1''B''1 = − ''B''1 ''A''1 implies that if λ is an eigenvalue of ''A''1 then so is −λ and they have the same multiplicity.


Eigenfunctions of complex Hilbert transform

The links between the Neumann–Poincaré operator and
geometric function theory Geometric function theory is the study of geometric properties of analytic functions. A fundamental result in the theory is the Riemann mapping theorem. Topics in geometric function theory The following are some of the most important topics in ge ...
appeared first in . The precise relationship between single and double layer potentials, Fredholm eigenvalues and the complex Hilbert transform is explained in detail in . Briefly given a smooth Jordan curve, the complex derivatives of its single and double layer potentials are −1 and +1 eigenfunctions of the complex Hilbert transform. Let 𝕳 be the direct sum :\displaystyle{\mathfrak{H}=\mathfrak{H}(\overline{\Omega})\oplus \mathfrak{H}(\overline{\Omega^c}),} where the first space consists of functions smooth on the closure of Ω and harmonic on Ω; and the second consists of functions smooth on the closure of Ω''c'', harmonic on Ω''c'' and at ≈. The space 𝕳 is naturally an inner product space with corresponding norm given by :\displaystyle{\, f_-\oplus f_+\, _{\mathfrak{H^2=\iint_\Omega , \nabla f_-, ^2 + \iint_{\Omega^c} , \nabla f_+, ^2.} Each element of 𝕳 can be written uniquely as the restriction of the sum of a double layer and single layer potential, provided that the moments are normalized to have 0 integral on ∂Ω. Thus for ''f'' ⊕ ''f''+ in 𝕳, there are unique φ, ψ in C(∂Ω) with integral 0 such that :\displaystyle{f_-=D(\varphi), _{\Omega} + S(\psi), _\Omega,\,\,\,\,\, f_+=D(\varphi), _{\Omega^c} + S(\psi), _{\Omega^c}.} Under this correspondence :\displaystyle{\varphi=f_-, _{\partial \Omega} - f_+, _{\partial\Omega},\,\,\,\, \psi=\partial_{n}f_-, _{\partial\Omega} -\partial_{n} f_+, _{\partial\Omega}.} The layer potentials can be identified with their images in 𝕳: :\displaystyle{D(\varphi)=D(\varphi), _\Omega \oplus D(\varphi), _{\Omega^c},\,\,\,\, S(\psi)=S(\psi), _\Omega \oplus S(\psi)_{\Omega^c}.} The space of double layer potentials is orthogonal to the space of single layer potentials for the inner product. In fact by Green's theorem :\displaystyle{(S,D)=\iint_{\Omega \cup \Omega^c}\nabla S\cdot \nabla D= -\int_{\partial\Omega} S\partial_n D + \int_{\partial\Omega} S\partial_n D=0.} Define an isometric embedding of 𝕳R in L2(C) by :\displaystyle{U(f_- \oplus f_+) =(\partial_{z} f_-)\chi_\Omega + (\partial_{z} f_-)\chi_{\Omega^c}.} The image lies in A2(Ω) ⊕ A2''c''), the direct sum of the
Bergman space In complex analysis, functional analysis and operator theory, a Bergman space, named after Stefan Bergman, is a function space of holomorphic functions in a domain ''D'' of the complex plane that are sufficiently well-behaved at the boundary that t ...
s of square integrable holomorphic functions on Ω and Ω''c''. Since polynomials in ''z'' are dense in A2(Ω) and polynomials in ''z''−1 without constant term are dense in A2''c''), the image of ''U'' is dense in A2(Ω) ⊕ A2''c''). It can be verified directly that for φ, ψ real :\displaystyle{T_c(U(D(\varphi) + S(\psi)))=U(D(\varphi)-S(\psi)).} In fact for single layer potentials, applying Green's theorem on the domain Ω ∪ Ω''c'' with a small closed disk of radius ε removed around a point ''w'' of the domain, it follows that :\displaystyle{\iint_{\Omega\cup\Omega^c, \,\, , z-w, >\varepsilon} \nabla N(w-z) \cdot \nabla S(\psi)(z) \,dx\, dy=-\int_{, z-w, =\varepsilon} \partial_n N(z-w)\,S(\psi)(z)=-S(\psi)(z),} since the mean of a harmonic function over a circle is its value at the centre. Using the fact that −1 is the fundamental solution for ∂''w'', this can be rewritten as :\displaystyle{{1\over \pi} \,\iint_{\Omega\cup\Omega^c}{\overline{\partial_z S(\psi)(z)}\over z-w}\,dx\,dy= -S(\psi)(w).} Applying ∂''w'' to both sides gives :\displaystyle{T_c(\partial_z S(\psi)) = -\partial_z S(\psi).} Similarly for a double layer potential : \displaystyle{\iint_{\Omega\cup\Omega^c, \,\, , z-w, >\varepsilon} \nabla N(w-z) \cdot \nabla D(\varphi)(z) \,dx\, dy=\int_{, z-w, =\varepsilon} N(z-w)\,\partial_n D(\varphi)(z)=0,} since the mean of the normal derivative of a harmonic function over a circle is zero. As above, using the fact {{overline, π''z''−1 is the fundamental solution for ∂''w'', this can be rewritten in terms of complex derivatives as :\displaystyle{{1\over \pi} \,\iint_{\Omega\cup\Omega^c}{\overline{\partial_z D(\varphi)(z)}\over z-w}\,dx\,dy= D(\varphi)(w).} Applying ∂''w'' to both sides, :\displaystyle{T_c(\partial_z D(\varphi)) = \partial_z D(\varphi).}


Connection with Hilbert transform on a domain

Let L2(∂Ω)0 be the closed subspace of L2(∂Ω) orthogonal to the constant functions. Let ''P''0 the orthogonal projection onto L2(∂Ω)0 and set :\displaystyle{T_{K,0}=P_0 T_K P_0.} With respect to the new inner product on L2(∂Ω)0 :\displaystyle{(f,g)_0=((1/2-T_K)^{-1}Sf,g)} the operator ''T''''K'',0 is formally self-adjoint. Let ''H''0 be the Hilbert space completion. Define a unitary operator ''V'' from ''H''0 onto A2(Ω) by :\displaystyle{V(\psi)=U(D(\varphi)+S(\psi)), _{\Omega},} where :\displaystyle{({1\over 2} I-T_K)\varphi= -S\psi.} Then :\displaystyle{VT_{K,0}V^*=T_\Omega.}


Fredholm eigenfunctions

If φ is an eigenfunction of ''T''''K'' on ∂Ω corresponding to an eigenvalue λ with , λ, < 1/2, then φ is orthogonal to the constants and can be taken real-valued. Let :\displaystyle{\Phi_-=\partial_z D(\varphi), _\Omega \in A^2(\Omega),\,\,\, \Phi_+=\partial_z D(\varphi), _{\Omega^c}\in A^2(\Omega^c).} Since double potentials are harmonic, given as the real part of a holomorphic function, :\displaystyle{\Phi_{\pm}(w)={1\over 2\pi i}\int_{\partial \Omega} {\varphi(z)\over (z-w)^2} \,dz.} Then :\displaystyle{T_\Omega \Phi_-=\lambda \Phi_-,\,\,\, T_{\Omega^c} \Phi_+=\lambda \Phi_+.} Moreover :\displaystyle{(T_c \Phi_-), _{\Omega^c}= (\lambda + {1\over 2})\Phi_+,\,\,\,(T_c \Phi_+), _{\Omega}=(\lambda-{1\over 2})\Phi_-.} If two eigenfunctions φ and ψ are orthogonal for the inner product defined by ''S'', then their transforms Φ± and Ψ± are orthogonal in A2(Ω) and A2''c'').


Eigenfunctions in Hardy space

{{See also, Singular integral operators on closed curves The Hardy space H2(∂Ω) can be defined as the closure of the complex polynomials in ''z'' in L2(∂Ω). The Cauchy transform of ''f'' in H2(∂Ω) :\displaystyle{F(w)={1\over 2\pi i}\int_{\partial\Omega} {f(z)\over z-w}\, dz} defines a holomorphic function ''F'' in Ω such that its restrictions to the level curves ∂Ω''s'' in a tubular neighbourhood of ∂Ω have uniformly bounded L2 norms. The classical definition of Hardy space is of holomorphic functions on Ω with this property. Identifying the level curves with ∂Ω, it follows that the restrictions of ''F'' tend to ''f'' in L2 norm. Writing H2(Ω) for the classical Hardy space, identified with H2(∂Ω) by taking L2 boundary values, it follows that Hardy space H2(Ω) is a dense subspace of Bergman space A2(Ω). Define the conjugate Cauchy transform of ''f'' by{{harvnb, Krzyż, Partyka, 1993 :\displaystyle{Cf(w)={1\over 2\pi i}\int_{\partial\Omega} {\overline{f(z)}\over z-w}\, d\overline{z}.} It lies in H2(Ω). Moreover for ''w'' in Ω :\displaystyle{Cf(w)=T_\Omega F(w),} since by Green's theorem :\displaystyle{{1\over \pi}\iint_{\Omega,\,\,\,, z-w, >\varepsilon} {\overline{F(z)}\over (z-w)^2} \,dx\,dy = Cf(w) -{1\over 2\pi i} \int_{, z-w, =\varepsilon} {\overline{F(z)}\over z-w}\, d\overline{z}=Cf(w).} For a smooth Jordan curve ∂Ω, the Fredholm eigenfunctions of ''T''Ω all lie in H2(Ω).


See also

*
Grunsky matrix 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 pai ...
* Planar Riemann surface


Notes

{{reflist, 30em


References

*{{citation, last=Ahlfors, first= Lars V., authorlink=Lars Ahlfors, title=Remarks on the Neumann–Poincaré integral equation, journal=Pacific J. Math., volume= 2, issue= 3, year=1952, pages =271–280, doi=10.2140/pjm.1952.2.271, doi-access=free *{{citation, last1=Bergman, first1= S., authorlink=Stefan Bergman, last2=Schiffer, first2= M., title=Kernel functions and conformal mapping, journal=
Compositio Mathematica ''Compositio Mathematica'' is a monthly peer-reviewed mathematics journal established by L.E.J. Brouwer in 1935. It is owned by the Foundation Compositio Mathematica, and since 2004 it has been published on behalf of the Foundation by the London M ...
, volume= 8, year=1951, pages=205–249 *{{citation, first1=J. , last1=Blumenfeld, first2=W., last2= Mayer, title= Über Poincaresche fundamentalfunktionen, journal= Sitz. Wien. Akad. Wiss., Math.-Nat. Klasse , volume=122, year=1914, pages= 2011–2047 *{{citation, last=Burbea, first= Jacob, title=Fredholm spectrum and Grunsky inequalities in general domains, journal=Studia Math., volume= 83, issue= 2, year=1986, pages= 167–200, doi= 10.4064/sm-83-2-167-200, doi-access= free *{{citation, last=Folland, first= Gerald B., authorlink=Gerald Folland, title= Introduction to partial differential equations, edition=2nd, publisher=Princeton University Press, year=1995, isbn= 978-0-691-04361-6 *{{citation, title=Integral Equations: Theory and Numerical Treatment, volume= 120, series= International Series of Numerical Mathematics, first=Wolfgang, last= Hackbusch, publisher=Springer, year= 1995, isbn=978-3764328719 *{{citation, last1=Hsiao, first1=George C., last2= Wendland, first2= Wolfgang L., title=Boundary integral equations, series=Applied Mathematical Sciences, volume= 164, publisher= Springer-Verlag, year= 2008, isbn=978-3-540-15284-2 *{{citation, last=Kellogg, first= Oliver Dimon, authorlink=Oliver Dimon Kellogg, title=Foundations of potential theory, series= Die Grundlehren der Mathematischen Wissenschaften, volume=31, publisher= Springer-Verlag, year=1929 *{{citation, last1=Khavinson, first1= D., last2= Putinar, first2= M., last3= Shapiro, first3= H. S., authorlink3=Harold S. Shapiro, title=Poincaré's variational problem in potential theory, journal=Arch. Ration. Mech. Anal., volume= 185 , issue= 1, year=2007, pages=143–184, doi=10.1007/s00205-006-0045-1, bibcode= 2007ArRMA.185..143K, citeseerx= 10.1.1.569.7145, s2cid= 855706 *{{citation, title=Linear Integral Equations, volume= 82, series=Applied Mathematical Sciences, first=Rainer, last=Kress, edition=2nd, publisher=Springer, year=1999, isbn=978-0387987002 *{{citation, last1=Krzyż, first1= Jan G., last2= Partyka, first2= Dariusz, title= Generalized Neumann–Poincaré operator, chord-arc curves and Fredholm eigenvalues, journal= Complex Variables Theory Appl., volume= 21, issue= 3–4, year=1993, pages= 253–263, doi=10.1080/17476939308814634 *{{citation, last=Landkof, first=N. S., title= Foundations of modern potential theory, series=Die Grundlehren der mathematischen Wissenschaften, volume= 180, publisher= Springer-Verlag, year= 1972 *{{citation, title=Mathematical physics: an advanced course, first=S. G., last= Mikhlin, publisher=North Holland, year= 1971 *{{citation, first=Carl, last= Neumann, authorlink=Carl Neumann, title=Untersuchungen über das logarithmische und Newton'sche Potential, location= Leipzig, publisher=
Teubner The Bibliotheca Teubneriana, or ''Bibliotheca Scriptorum Graecorum et Romanorum Teubneriana'', also known as Teubner editions of Greek and Latin texts, comprise one of the most thorough modern collection published of ancient (and some medieval) ...
, year=1877, url= https://archive.org/details/untersuchungen00neumiala *{{citation, last=Partyka, first= Dariusz, title=The generalized Neumann–Poincaré operator and its spectrum, series= Dissertationes Math, volume= 366, year=1997, url=http://pldml.icm.edu.pl/pldml/element/bwmeta1.element.zamlynska-7660e1af-389e-4bf7-91a1-fb5f3b08165e *{{citation, last=Plemelj, first= J., title= Potentialtheoretische Untersuchungen, publisher=
Teubner The Bibliotheca Teubneriana, or ''Bibliotheca Scriptorum Graecorum et Romanorum Teubneriana'', also known as Teubner editions of Greek and Latin texts, comprise one of the most thorough modern collection published of ancient (and some medieval) ...
, year= 1911 *{{citation, last=Poincaré, first= H., authorlink=Henri Poincaré, title=La méthode de Neumann et le problème de Dirichlet, journal=Acta Math., volume= 20, pages=59–152, year=1897, doi=10.1007/bf02418028, doi-access=free *{{citation, title=Periodic Integral and Pseudodifferential Equations with Numerical Approximation, first1=Jukka, last1= Saranen, first2= Gennadi, last2= Vainikko, publisher=Springer, year= 2001, isbn=978-3540418788 *{{citation, last=Schiffer, first=M., title=The Fredholm eigenvalues of plane domains, journal=Pacific J. Math., volume= 7, issue=2, year= 1957, pages= 1187–1225, doi=10.2140/pjm.1957.7.1187, doi-access=free *{{citation, last=Schiffer, first= M., title=Fredholm eigenvalues of multiply connected domains, journal=Pacific J. Math., volume= 9, year= 1959, pages= 211–269, doi=10.2140/pjm.1959.9.211, doi-access=free *{{citation, last1=Schiffer, first1= M., last2= Hawley, first2= N. S., title=Connections and conformal mapping, journal=Acta Math., volume= 107 , issue= 3–4, year=1962, pages= 175–274, doi=10.1007/bf02545790, doi-access= free *{{citation, last=Schiffer, first= M., title=Fredholm eigenvalues and Grunsky matrices, journal=Ann. Polon. Math., volume= 39, year= 1981, pages= 149–164, doi= 10.4064/ap-39-1-149-164, doi-access= free *{{citation, series=Autovalori e autosoluzioni, C.I.M.E. Summer Schools, year=2011, volume= 27, pages=203–234, title= Fredholm Eigenvalues and Conformal Mapping, publisher=Springer, last=Schiffer, first=Menahem *{{citation, last=Shapiro, first=H. S., authorlink=Harold S. Shapiro, title=The Schwarz function and its generalization to higher dimensions, series= University of Arkansas Lecture Notes in the Mathematical Sciences, volume= 9, publisher= Wiley-Interscience, year= 1992, isbn=978-0-471-57127-8 *{{citation, last=Taylor, first= Michael E., authorlink=Michael E. Taylor, title= Partial differential equations II: Qualitative studies of linear equations, edition=2nd, series=Applied Mathematical Sciences, volume= 116, publisher=Springer, year=2011, isbn= 978-1-4419-7051-0 {{DEFAULTSORT:Neumann-Poincare operator Potential theory Partial differential equations Complex analysis Operator theory