Differential Forms On A Riemann Surface
   HOME

TheInfoList



OR:

In
mathematics Mathematics is a field of study that discovers and organizes methods, Mathematical theory, theories and theorems that are developed and Mathematical proof, proved for the needs of empirical sciences and mathematics itself. There are many ar ...
, differential forms on a Riemann surface are an important special case of the general theory of
differential forms In mathematics, differential forms provide a unified approach to define integrands over curves, surfaces, solids, and higher-dimensional manifolds. The modern notion of differential forms was pioneered by Élie Cartan. It has many applications, ...
on
smooth manifold In mathematics, a differentiable manifold (also differential manifold) is a type of manifold that is locally similar enough to a vector space to allow one to apply calculus. Any manifold can be described by a collection of charts (atlas). One may ...
s, distinguished by the fact that the
conformal structure In mathematics, conformal geometry is the study of the set of angle-preserving (conformal map, conformal) transformations on a space. In a real two dimensional space, conformal geometry is precisely the geometry of Riemann surfaces. In space highe ...
on the
Riemann surface In mathematics, particularly in complex analysis, a Riemann surface is a connected one-dimensional complex manifold. These surfaces were first studied by and are named after Bernhard Riemann. Riemann surfaces can be thought of as deformed vers ...
intrinsically defines a
Hodge star operator In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a Dimension (vector space), finite-dimensional orientation (vector space), oriented vector space endowed with a Degenerate bilinear form, nonde ...
on
1-forms In differential geometry, a one-form (or covector field) on a differentiable manifold is a differential form of degree one, that is, a smooth section of the cotangent bundle. Equivalently, a one-form on a manifold M is a smooth mapping of the ...
(or differentials) without specifying a
Riemannian metric In differential geometry, a Riemannian manifold is a geometric space on which many geometric notions such as distance, angles, length, volume, and curvature are defined. Euclidean space, the N-sphere, n-sphere, hyperbolic space, and smooth surf ...
. This allows the use of
Hilbert space In mathematics, a Hilbert space is a real number, real or complex number, complex inner product space that is also a complete metric space with respect to the metric induced by the inner product. It generalizes the notion of Euclidean space. The ...
techniques for studying function theory on the Riemann surface and in particular for the construction of harmonic and holomorphic differentials with prescribed singularities. These methods were first used by in his variational approach to the Dirichlet principle, making rigorous the arguments proposed by Riemann. Later found a direct approach using his method of orthogonal projection, a precursor of the modern theory of
elliptic differential operator In the theory of partial differential equations, elliptic operators are differential operators that generalize the Laplace operator. They are defined by the condition that the coefficients of the highest-order derivatives be positive, which im ...
s and
Sobolev space In mathematics, a Sobolev space is a vector space of functions equipped with a norm that is a combination of ''Lp''-norms of the function together with its derivatives up to a given order. The derivatives are understood in a suitable weak sense ...
s. These techniques were originally applied to prove the
uniformization theorem In mathematics, the uniformization theorem states that every simply connected Riemann surface is conformally equivalent to one of three Riemann surfaces: the open unit disk, the complex plane, or the Riemann sphere. The theorem is a generali ...
and its generalization to planar Riemann surfaces. Later they supplied the analytic foundations for the harmonic integrals of . This article covers general results on differential forms on a Riemann surface that do not rely on any choice of Riemannian structure.


Hodge star on 1-forms

On a Riemann surface the Hodge star is defined on 1-forms by the local formula :\displaystyle It is well-defined because it is invariant under holomorphic changes of coordinate. Indeed, if is holomorphic as a function of , then by the
Cauchy–Riemann equations In the field of complex analysis in mathematics, the Cauchy–Riemann equations, named after Augustin-Louis Cauchy, Augustin Cauchy and Bernhard Riemann, consist of a system of differential equations, system of two partial differential equatio ...
and . In the new coordinates : \displaystyle so that :-q_1 \,du + p_1 \,dv= -(px_v+qy_v)du + (px_u + q y_u)dv = -q(x_u du + x_v dv) + p(y_u du + y_vdv) = -q\,dx + p\, dy, proving the claimed invariance. Note that for 1-forms and :\displaystyle In particular if then :\displaystyle Note that in standard coordinates : \star dz = -i dz,\,\, \star d\overline = i d\overline. Recall also that : = \left( - i \right),\,\,\, = \left( +i \right), so that : df= f_z \, dz + f_\, d\overline = \partial f + \barf. The decomposition d = \partial + \bar is independent of the choice of local coordinate. The 1-forms with only a dz component are called (1,0) forms; those with only a d\overline component are called (0,1) forms. The operators \partial and \overline are called the Dolbeault operators. It follows that : \star df= -i\partial f + i\barf. The Dolbeault operators can similarly be defined on 1-forms and as zero on 2-forms. They have the properties :d=\partial+\bar :\partial^2=\bar^2=\partial\bar+\bar\partial=0.


Poincaré lemma

On a Riemann surface the
Poincaré lemma In mathematics, the Poincaré lemma gives a sufficient condition for a closed differential form to be exact (while an exact form is necessarily closed). Precisely, it states that every closed ''p''-form on an open ball in R''n'' is exact for ''p'' ...
states that every closed 1-form or 2-form is locally exact. Thus if ''ω'' is a smooth 1-form with then in some open neighbourhood of a given point there is a smooth function ''f'' such that in that neighbourhood; and for any smooth 2-form Ω there is a smooth 1-form ''ω'' defined in some open neighbourhood of a given point such that in that neighbourhood. If is a closed 1-form on , then . If then and . Set :\displaystyle so that . Then must satisfy and . The right hand side here is independent of ''x'' since its partial derivative with respect to ''x'' is 0. So :\displaystyle and hence :\displaystyle Similarly, if then with . Thus a solution is given by and :\displaystyle Comment on differential forms with compact support. Note that if ''ω'' has compact support, so vanishes outside some smaller rectangle with and , then the same is true for the solution ''f''(''x'',''y''). So the Poincaré lemma for 1-forms holds with this additional conditions of compact support. A similar statement is true for 2-forms; but, since there is some choices for the solution, a little more care has to be taken in making those choices. In fact if Ω has compact support on and if furthermore , then with ''ω'' a 1-form of compact support on . Indeed, Ω must have support in some smaller rectangle with and . So vanishes for or and for or . Let ''h''(''y'') be a smooth function supported in (''c''1,''d''1) with . Set : it is a smooth function supported in (''a''1,''b''1). Hence is smooth and supported in . It now satisfies . Finally set :P(x,y)=\int_c^y R(x,y) dy,\,\,\, Q(x,y)=h(y)\int_a^x k(s) \, ds. Both ''P'' and ''Q'' are smooth and supported in with and . Hence is a smooth 1-form supported in with :d\omega = (Q_x + P_y)\,dx \wedge dy = r \,dx \wedge dy =\Omega.


Integration of 2-forms

If Ω is a continuous 2-form of compact support on a Riemann surface ''X'', its support ''K'' can be covered by finitely many coordinate charts ''U''''i'' and there is a partition of unity χ''i'' of smooth non-negative functions with compact support such that Σ χ''i'' = 1 on a neighbourhood of ''K''. Then the integral of Ω is defined by :\displaystyle where the integral over ''U''''i'' has its usual definition in local coordinates. The integral is independent of the choices here. If Ω has the local representation ''f''(''x'',''y'') ''dx'' ∧ ''dy'', then , Ω, is the density , ''f''(''x'',''y''), ''dx'' ∧ ''dy'', which is well defined and satisfies , ∫''X'' Ω, ≤ ∫''X'' , Ω, . If Ω is a non-negative continuous density, not necessarily of compact support, its integral is defined by :\displaystyle If Ω is any continuous 2-form it is integrable if ∫''X'' , Ω, < ∞. In this case, if ∫''X'' , Ω, = lim ∫''X'' ψ''n'' , Ω, , then ∫''X'' Ω can be defined as lim ∫''X'' ψ''n'' Ω. The integrable continuous 2-forms form a complex normed space with norm , , Ω, , 1 = ∫''X'' , Ω, .


Integration of 1-forms along paths

If ''ω'' is a 1-form on a Riemann surface ''X'' and ''γ''(''t'') for is a smooth path in ''X'', then the mapping ''γ'' induces a 1-form ''γ''∗''ω'' on 'a'',''b'' The integral of ''ω'' along ''γ'' is defined by :\displaystyle This definition extends to piecewise smooth paths ''γ'' by dividing the path up into the finitely many segments on which it is smooth. In local coordinates if and then :\displaystyle so that :\displaystyle Note that if the 1-form ''ω'' is exact on some connected open set ''U'', so that for some smooth function ''f'' on ''U'' (unique up to a constant), and ''γ''(''t''), , is a smooth path in ''U'', then :\displaystyle This depends only on the difference of the values of ''f'' at the endpoints of the curve, so is independent of the choice of ''f''. By the Poincaré lemma, every closed 1-form is locally exact, so this allows ∫γ ω to be computed as a sum of differences of this kind and for the integral of closed 1-forms to be extended to continuous paths: Monodromy theorem. If ''ω'' is a closed 1-form, the integral can be extended to any continuous path ''γ''(''t''), so that it is invariant under any
homotopy In topology, two continuous functions from one topological space to another are called homotopic (from and ) if one can be "continuously deformed" into the other, such a deformation being called a homotopy ( ; ) between the two functions. ...
of paths keeping the end points fixed. ::In fact, the image of ''γ'' is compact, so can be covered by finitely many connected open sets ''U''''i'' on each of which ω can be written ''df''''i'' for some smooth function ''f''''i'' on ''U''''i'', unique up to a constant. It may be assumed that 'a'',''b''is broken up into finitely many closed intervals with and so that . From the above if ''γ'' is piecewise smooth, :::\displaystyle ::Now ''γ''(''t''''i'') lies in the open set , hence in a connected open component ''V''''i''. The difference satisfies , so is a constant ''c''''i'' independent of ''γ''. Hence :::\displaystyle ::The formula on the right hand side also makes sense if ''γ'' is just continuous on 'a'',''b''and can be used to define . The definition is independent of choices: for the curve ''γ'' can be uniformly approximated by piecewise smooth curves ''δ'' so close that for all ''i''; the formula above then equals and shows the integral is independent of the choice of ''δ''. The same argument shows that the definition is also invariant under small homotopies fixing endpoints; by compactness, it is therefore invariant under any homotopy fixing endpoints. The same argument shows that a homotopy between closed continuous loops does not change their integrals over closed 1-forms. Since , the integral of an exact form over a closed loop vanishes. Conversely if the integral of a closed 1-form ''ω'' over any closed loop vanishes, then the 1-form must be exact. ::Indeed a function ''f''(''z'') can be defined on ''X'' by fixing a point ''w'', taking any path ''δ'' from ''w'' to ''z'' and setting . The assumption implies that ''f'' is independent of the path. To check that , it suffices to check this locally. Fix ''z''0 and take a path ''δ''1 from ''w'' to ''z''0. Near ''z''0 the Poincaré lemma implies that for some smooth function ''g'' defined in a neighbourhood of ''z''0. If ''δ''2 is a path from ''z''0 to ''z'', then , so ''f'' differs from ''g'' by a constant near ''z''0. Hence near ''z''0. A closed 1-form is exact if and only if its integral around any piecewise smooth or continuous Jordan curve vanishes. ::In fact the integral is already known to vanish for an exact form, so it suffices to show that if for all piecewise smooth closed Jordan curves ''γ'' then for all closed continuous curves ''γ''. Let ''γ'' be a closed continuous curve. The image of ''γ'' can be covered by finitely many opens on which ''ω'' is exact and this data can be used to define the integral on ''γ''. Now recursively replace ''γ'' by smooth segments between successive division points on the curve so that the resulting curve ''δ'' has only finitely many intersection points and passes through each of these only twice. This curve can be broken up as a superposition of finitely many piecewise smooth Jordan curves. The integral over each of these is zero, so their sum, the integral over ''δ'', is also zero. By construction the integral over ''δ'' equals the integral over ''γ'', which therefore vanishes. The above argument also shows that given a continuous Jordan curve ''γ''(''t''), there is a finite set of simple smooth Jordan curves ''γ''''i''(''t'') with nowhere zero derivatives such that :\int_\gamma \omega = \sum_i \int_ \omega for any closed 1-form ''ω''. Thus to check exactness of a closed form it suffices to show that the vanishing of the integral around any regular closed curve, i.e. a simple smooth Jordan curve with nowhere vanishing derivative. The same methods show that any continuous loop on a Riemann surface is homotopic to a smooth loop with nowhere zero derivative.


Green–Stokes formula

If ''U'' is a bounded region in the complex plane with boundary consisting of piecewise smooth curves and ''ω'' is a 1-form defined on a neighbourhood of the closure of ''U'', then the Green–Stokes formula states that :\int_ \omega = \int_U d\omega . In particular if ''ω'' is a 1-form of compact support on C then :\int_ d\omega = 0 , since the formula may be applied to a large disk containing the support of ω. Similar formulas hold on a Riemann surface ''X'' and can be deduced from the classical formulas using
partitions of unity In mathematics, a partition of unity on a topological space is a Set (mathematics), set of continuous function (topology), continuous functions from to the unit interval ,1such that for every point x\in X: * there is a neighbourhood (mathem ...
. Thus if is a connected region with compact closure and piecewise smooth boundary ∂''U'' and ''ω'' is a 1-form defined on a neighbourhood of the closure of ''U'', then the Green–Stokes formula states that :\displaystyle Moreover, if ''ω'' is a 1-form of compact support on ''X'' then :\int_ d\omega = 0 . To prove the second formula take a partition of unity ''ψ''''i'' supported in coordinate charts covering the support of ''ω''. Then , by the planar result. Similarly to prove the first formula it suffices to show that :\displaystyle when ''ψ'' is a smooth function compactly supported in some coordinate patch. If the coordinate patch avoids the boundary curves, both sides vanish by the second formula above. Otherwise it can be assumed that the coordinate patch is a disk, the boundary of which cuts the curve transversely at two points. The same will be true for a slightly smaller disk containing the support of ''ψ''. Completing the curve to a Jordan curve by adding part of the boundary of the smaller disk, the formula reduces to the planar Green-Stokes formula. The Green–Stokes formula implies an adjoint relation for the Laplacian on functions defined as Δ''f'' = −''d''∗''df''. This gives a 2-form, given in local coordinates by the formula :\Delta f =\left( - - \right) \, dx \wedge dy . Then if ''f'' and ''g'' are smooth and the closure of ''U'' is compact :\int_U f\Delta g - g \Delta f = \int_ g - f . Moreover, if ''f'' or ''g'' has compact support then :\int_X f \Delta g = \int_X g \Delta f .


Duality between 1-forms and closed curves

Theorem. If ''γ'' is a continuous Jordan curve on a Riemann surface ''X'', there is a smooth closed 1-form ''α'' of compact support such that for any closed smooth 1-form ''ω'' on ''X''. ::It suffices to prove this when ''γ'' is a regular closed curve. By the
inverse function theorem In mathematics, the inverse function theorem is a theorem that asserts that, if a real function ''f'' has a continuous derivative near a point where its derivative is nonzero, then, near this point, ''f'' has an inverse function. The inverse fu ...
, there is a tubular neighbourhood of the image of ''γ'', i.e. a smooth diffeomorphism of the annulus into ''X'' such that . Using a bump function on the second factor, a non-negative function ''g'' with compact support can be constructed such that ''g'' is smooth off ''γ'', has support in a small neighbourhood of ''γ'', and in a sufficiently small neighbourhood of ''γ'' is equal to 0 for and 1 for . Thus ''g'' has a jump discontinuity across ''γ'', although its differential ''dg'' is smooth with compact support. But then, setting , it follows from Green's formula applied to the annulus that :::\int_X \omega \wedge \alpha = \int_X dg \wedge \omega = \int_ dg \wedge \omega = \int_d(g\omega) = \int_\gamma \omega. Corollary 1. A closed smooth 1-form ''ω'' is exact if and only if for all smooth 1-forms ''α'' of compact support. ::In fact if ''ω'' is exact, it has the form ''df'' for ''f'' smooth, so that by Green's theorem. Conversely, if for all smooth 1-forms ''α'' of compact support, the duality between Jordan curves and 1-forms implies that the integral of ''ω'' around any closed Jordan curve is zero and hence that ''ω'' is exact. Corollary 2. If ''γ'' is a continuous closed curve on a Riemann surface ''X'', there is a smooth closed 1-form ''α'' of compact support such that for any closed smooth 1-form ''ω'' on ''X''. The form ''α'' is unique up to adding an exact form and can be taken to have support in any open neighbourhood of the image of ''γ''. ::In fact ''γ'' is homotopic to a piecewise smooth closed curve ''δ'', so that . On the other hand there are finitely many piecewise smooth Jordan curves ''δ''''i'' such that . The result for ''δ''''i'' thus implies the result for ''γ''. If ''β'' is another form with the same property, the difference satisfies for all closed smooth 1-forms ''ω''. So the difference is exact by Corollary 1. Finally, if ''U'' is any neighbourhood of the image of ''γ'', then the last result follows by applying first assertion to ''γ'' and ''U'' in place of ''γ'' and ''X''.


Intersection number of closed curves

The
intersection number In mathematics, and especially in algebraic geometry, the intersection number generalizes the intuitive notion of counting the number of times two curves intersect to higher dimensions, multiple (more than 2) curves, and accounting properly for ta ...
of two closed curves ''γ''1, ''γ''2 in a Riemann surface ''X'' can be defined analytically by the formula : I(\gamma_1 , \gamma_2) = \int_X \alpha_1 \wedge \alpha_2, where ''α''1 and ''α''2 are smooth 1-forms of compact support corresponding to ''γ''1 and ''γ''2. From the definition it follows that . Since ''αi'' can be taken to have its support in a neighbourhood of the image of γ''i'', it follows that if ''γ''1 and ''γ''2 are disjoint. By definition it depends only on the homotopy classes of ''γ''1 and ''γ''2. More generally the intersection number is always an integer and counts the number of times ''with signs'' that the two curves intersect. A crossing at a point is a positive or negative crossing according to whether ''d''γ1 ∧ ''d''γ2 has the same or opposite sign to , for a local holomorphic parameter ''z'' = ''x'' + ''iy''. ::Indeed, by homotopy invariance, it suffices to check this for smooth Jordan curves with nowhere vanishing derivatives. The ''α''1 can be defined by taking α1''df'' with ''f'' of compact support in a neighbourhood of the image of γ1 equal to 0 near the left hand side of ''γ''1, 1 near the right hand side of ''γ''1 and smooth off the image of ''γ''1. Then if the points of intersection of ''γ''2(''t'') with γ1 occur at ''t'' = ''t''1, ..., ''t''''m'', then :::I(\gamma_1,\gamma_2) = \int_ \alpha_1 = \int_ df = \sum f\circ \gamma_2(t_i+) - f\circ\gamma_2(t_i-). ::This gives the required result since the jump is + 1 for a positive crossing and −1 for a negative crossing.


Holomorphic and harmonic 1-forms

A holomorphic 1-form ω is one that in local coordinates is given by an expression ''f''(''z'') ''dz'' with ''f'' holomorphic. Since dg=\partial_z g \, dz + \partial_ g\, d\overline, it follows that ''dω'' = 0, so any holomorphic 1-form is closed. Moreover, since ∗''dz'' = −''i'' ''dz'', ω must satisfy ∗''ω'' = −''iω''. These two conditions characterize holomorphic 1-forms. For if ω is closed, locally it can be written as ''dg'' for some ''g'', The condition ∗''dg'' = ''i'' ''dg'' forces \partial_ g = 0, so that ''g'' is holomorphic and ''dg'' = ''g'' '(''z'') ''dz'', so that ω is holomorphic. Let ''ω'' = ''f'' ''dz'' be a holomorphic 1-form. Write ''ω'' = ''ω''1 + ''iω''2 with ''ω''1 and ''ω''2 real. Then ''dω''1 = 0 and ''dω''2 = 0; and since ∗''ω'' = −''iω'', ∗''ω''1 = ''ω''2. Hence ''d''∗''ω''1 = 0. This process can clearly be reversed, so that there is a one-one correspondence between holomorphic 1-forms and real 1-forms ω1 satisfying ''d''ω1 = 0 and ''d''∗ω1 = 0. Under this correspondence, ''ω''1 is the real part of ''ω'' while ''ω'' is given by ''ω'' = ''ω''1 + ''i''∗''ω''1. Such forms ω1 are called harmonic 1-forms. By definition ω1 is harmonic if and only if ∗''ω''1 is harmonic. Since holomorphic 1-forms locally have the form ''df'' with ''f'' a holomorphic function and since the real part of a holomorphic function is harmonic, harmonic 1-forms locally have the form ''dh'' with ''h'' a
harmonic function In mathematics, mathematical physics and the theory of stochastic processes, a harmonic function is a twice continuously differentiable function f\colon U \to \mathbb R, where is an open subset of that satisfies Laplace's equation, that i ...
. Conversely if ''ω''1 can be written in this way locally, ''d''∗ω1 = ''d''∗''dh'' = (''h''''xx'' + ''h''''yy'') ''dx''∧''dy'' so that ''h'' is harmonic. Remark. The definition of harmonic functions and 1-forms is intrinsic and only relies on the underlying Riemann surface structure. If, however, a conformal metric is chosen on the Riemann surface, the adjoint ''d''* of ''d'' can be defined and the Hodge star operation extended to functions and 2-forms. The Hodge Laplacian can be defined on ''k''-forms as ∆''k'' = ''dd''* +''d''*''d'' and then a function ''f'' or a 1-form ''ω'' is harmonic if and only if it is annihilated by the Hodge Laplacian, i.e. ∆0''f'' = 0 or ∆1''ω'' = 0. The metric structure, however, is not required for the application to the uniformization of simply connected or planar Riemann surfaces.


Sobolev spaces on T2

The theory of Sobolev spaces on can be found in , an account which is followed in several later textbooks such as and . It provides an analytic framework for studying function theory on the torus C/Z+''i'' Z = R2 / Z2 using
Fourier series A Fourier series () is an Series expansion, expansion of a periodic function into a sum of trigonometric functions. The Fourier series is an example of a trigonometric series. By expressing a function as a sum of sines and cosines, many problems ...
, which are just eigenfunction expansions for the Laplacian . The theory developed here essentially covers tori C / Λ where Λ is a lattice in C. Although there is a corresponding theory of Sobolev spaces on any compact Riemann surface, it is elementary in this case, because it reduces to
harmonic analysis Harmonic analysis is a branch of mathematics concerned with investigating the connections between a function and its representation in frequency. The frequency representation is found by using the Fourier transform for functions on unbounded do ...
on the compact Abelian group . Classical approaches to Weyl's lemma use harmonic analysis on the non-compact Abelian group C = R2, i.e. the methods of
Fourier analysis In mathematics, Fourier analysis () is the study of the way general functions may be represented or approximated by sums of simpler trigonometric functions. Fourier analysis grew from the study of Fourier series, and is named after Joseph Fo ...
, in particular convolution operators and 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 ...
of the Laplacian. Let T2 = = R2/Z2 = C/Λ where Λ = Z + ''i'' Z. For λ = ''m'' + ''i'' ''n'' ≅ (''m'',''n'') in Λ, set . Furthermore, set ''D''''x'' = −''i''∂/∂''x'' and ''D''''y'' = −''i''∂/∂''y''. For α = (''p'',''q'') set ''D''α =(''D''''x'')''p'' (''D''''y'')''q'', a differential operator of total degree , ''α'', = ''p'' + ''q''. Thus , where . The (''e''λ) form an
orthonormal basis In mathematics, particularly linear algebra, an orthonormal basis for an inner product space V with finite Dimension (linear algebra), dimension is a Basis (linear algebra), basis for V whose vectors are orthonormal, that is, they are all unit vec ...
in C(T2) for the inner product , so that . For ''f'' in C(''T2) and ''k'' an integer, define the ''k''th Sobolev norm by :\, f\, _ = \left(\sum , \widehat(\lambda), ^2 (1 + , \lambda, ^2)^k\right)^. The associated inner product :\displaystyle makes C(T2) into an inner product space. Let ''H''''k''(T2) be its Hilbert space completion. It can be described equivalently as the Hilbert space completion of the space of trigonometric polynomials—that is finite sums —with respect to the ''k''th Sobolev norm, so that ''H''''k''(T2) = with inner product :(Σ ''a''λ ''e''λ, Σ ''b''μ ''e''μ)(''k'') = Σ ''a''λ (1 + , λ, 2)''k''. As explained below, the elements in the intersection ''H''(T2) = \cap ''H''''k''(T2) are exactly the smooth functions on T2; elements in the union ''H''−∞(T2) = \cup ''H''''k''(T2) are just distributions on T2 (sometimes referred to as "periodic distributions" on R2). The following is a (non-exhaustive) list of properties of the Sobolev spaces. *Differentiability and Sobolev spaces. for ''k'' ≥ 0 since, using the
binomial theorem In elementary algebra, the binomial theorem (or binomial expansion) describes the algebraic expansion of powers of a binomial. According to the theorem, the power expands into a polynomial with terms of the form , where the exponents and a ...
to expand (1 + , λ, 2)''k'', ::\displaystyle *Differential operators. ''D''α ''H''''k''(T2) ⊂ ''H''''k''−, α, (T2) and ''D''α defines a bounded linear map from ''H''''k''(T2) to ''H''''k''−, α, (T2). The operator ''I'' + Δ defines a unitary map of ''H''''k''+2(T2) onto ''H''''k''(T2); in particular (''I'' + Δ)''k'' defines a unitary map of ''H''''k''(T2) onto ''H''−''k''(T2) for ''k'' ≥ 0. ::The first assertions follow because ''D''α ''e''λ = λα ''e''λ and , λα, ≤ , λ, , α, ≤ (1 + , λ, 2), α, /2. The second assertions follow because ''I'' + Δ acts as multiplication by 1 + , λ, 2 on ''e''λ. *Duality. For ''k'' ≥ 0, the pairing sending ''f'', ''g'' to (''f'',''g'') establishes a duality between ''H''''k''(T2) and ''H''−''k''(T2). ::This is a restatement of the fact that (''I'' + Δ)''k'' establishes a unitary map between these two spaces, because . *Multiplication operators. If ''h'' is a smooth function then multiplication by ''h'' defines a continuous operator on ''H''''k''(T2). ::For ''k'' ≥ 0, this follows from the formula for , , ''f'', , above and the Leibniz rule. Continuity for ''H''−''k''(T2) follows by duality, since . *Sobolev spaces and differentiability (Sobolev's embedding theorem). For ''k'' ≥ 0, and sup, α, ≤''k'' , ''D''α''f'', ≤ ''C''''k'' ⋅ , , ''f'', , (''k''+2). ::The inequalities for trigonometric polynomials imply the containments. The inequality for ''k'' = 0 follows from :::\sup \left, \sum a_\lambda e_\lambda\ \le \left(\sum (1+, \lambda, ^2)^\right)^ \cdot \sum , a_\lambda, \le \left(\sum , a_\lambda, ^2 (1+, \lambda, ^2)^2\right)^ = \left(\sum (1+, \lambda, ^2)^\right)^ \cdot \left\, \sum a_\lambda e_\lambda \right\, _, ::by the Cauchy-Schwarz inequality. The first term is finite by the
integral test In mathematics, an integral is the continuous analog of a Summation, sum, which is used to calculate area, areas, volume, volumes, and their generalizations. Integration, the process of computing an integral, is one of the two fundamental oper ...
, since ∬C (1 + , ''z'', 2)−2 ''dx'' ''dy'' = < ∞ using
polar coordinates In mathematics, the polar coordinate system specifies a given point (mathematics), point in a plane (mathematics), plane by using a distance and an angle as its two coordinate system, coordinates. These are *the point's distance from a reference ...
. In general if , α, ≤ k, then , sup ''D''α''f'', ≤ ''C''0 , , ''D''α''f'', , 2 ≤ ''C''0 ⋅ ''C''α ⋅ , , ''f'', , ''k''+2 by the continuity properties of ''D''α. *Smooth functions. ''C''(T2) = \cap ''H''''k''(T2) consists of Fourier series Σ ''a''λ ''e''λ such that for all ''k'' > 0, (1 + , λ, 2)''k'' , ''a''λ, tends to 0 as , λ, tends to ∞, i.e. the Fourier coefficients ''a''λ are of "rapid decay". ::This is an immediate consequence of the Sobolev embedding theorem. *Inclusion maps (Rellich's compactness theorem). If ''k'' > ''j'', the space ''H''''k''(T2) is a subspace of ''H''''j''(T2) and the inclusion ''H''''k''(T2) \rightarrow ''H''''j''(T2) is
compact Compact as used in politics may refer broadly to a pact or treaty; in more specific cases it may refer to: * Interstate compact, a type of agreement used by U.S. states * Blood compact, an ancient ritual of the Philippines * Compact government, a t ...
. ::With respect to the natural orthonormal bases, the inclusion map becomes multiplication by (1 + , λ, 2)−(''k''−''j'')/2. It is therefore compact because it is given by a diagonal matrix with diagonal entries tending to zero. *Elliptic regularity (Weyl's lemma). Suppose that ''f'' and ''u'' in ''H''−∞(T2) = \cup ''H''''k''(T2) satisfy ∆''u'' = ''f''. Suppose also that ψ ''f'' is a smooth function for every smooth function ψ vanishing off a fixed open set ''U'' in T2; then the same is true for ''u''. (Thus if ''f'' is smooth off ''U'', so is ''u''.) ::By the Leibniz rule , so . If it is known that ''φu'' lies in ''H''''k''(T2) for some ''k'' and all ''φ'' vanishing off ''U'', then differentiating shows that φ''u''''x'' and φ''u''''y'' lie in ''H''''k''−1(T2). The square-bracketed expression therefore also lies in ''H''''k''−1(T2). The operator (''I'' + Δ)−1 carries this space onto ''H''''k''+1(T2), so that ψ''u'' must lie in ''H''''k''+1(T2). Continuing in this way, it follows that ψ''u'' lies in \cap ''H''''k''(T2) = ''C''(T2). *Hodge decomposition on functions. ''H''0(T2) = ∆ ''H''2(T2) \oplus ker ∆ and ''C''(T2) = ∆ ''C''(T2) \oplus ker ∆. ::Identifying ''H''2(T2) with ''L''2(T2) = ''H''0(T2) using the unitary operator ''I'' + Δ, the first statement reduces to proving that the operator ''T'' = ∆(''I'' + Δ)−1 satisfies ''L''2(T2) = im ''T'' \oplus ker ''T''. This operator is bounded, self-adjoint and diagonalized by the orthonormal basis ''e''λ with eigenvalue , λ, 2(1 + , λ, 2)−1. The operator ''T'' has kernel C ''e''0 (the constant functions) and on (ker ''T'') = it has a bounded inverse given by ''S'' ''e''λ = , λ, −2(1 + , λ, 2) ''e''λ for λ ≠ 0. So im ''T'' must be closed and hence ''L''2(T2) = (ker ''T'') \oplus ker ''T'' = im ''T'' \oplus ker ''T''. Finally if ''f'' = ∆''g'' + ''h'' with ''f'' in ''C''(T2), ''g'' in ''H''2(T2) and ''h'' constant, ''g'' must be smooth by Weyl's lemma. *Hodge theory on T2. Let Ω''k''(T2) be the space of smooth ''k''-forms for 0 ≤ ''k'' ≤ 2. Thus Ω0(T2) = ''C''(T2), Ω1(T2) = ''C''(T2) ''dx'' \oplus ''C''(T2) ''dy'' and Ω2(T2) = ''C''(T2) ''dx'' ∧ ''dy''. The Hodge star operation is defined on 1-forms by ∗(''p'' dx + ''q'' ''dy'') = −''q'' ''dx'' + ''p'' ''dy''. This definition is extended to 0-forms and 2-forms by *''f'' = ''f'' ''dx'' ∧ ''dy'' and *(''g'' ''dx'' ∧ ''dy'') = ''g''. Thus ** = (−1)''k'' on ''k''-forms. There is a natural complex inner product on Ω''k''(T2) defined by ::\displaystyle :Define . Thus δ takes Ω''k''(T2) to Ω''k''−1(T2), annihilating functions; it is the adjoint of ''d'' for the above inner products, so that . Indeed by the Green-Stokes formula ::\displaystyle :The operators ''d'' and δ = ''d''* satisfy ''d''2 = 0 and ''δ''2 = 0. The Hodge Laplacian on ''k''-forms is defined by . From the definition . Moreover and . This allows the Hodge decomposition to be generalised to include 1-forms and 2-forms: *Hodge theorem. Ω''k''(T2) = ker ''d'' \cap ker ''d''∗ \oplus im ''d'' \oplus im ∗''d'' = ker ''d'' \cap ker ''d''* \oplus im ''d'' \oplus im ''d''*. In the Hilbert space completion of Ω''k''(T2) the orthogonal complement of is , the finite-dimensional space of harmonic ''k''-forms, i.e. the constant ''k''-forms. In particular in , , the space of harmonic ''k''-forms. Thus the
de Rham cohomology In mathematics, de Rham cohomology (named after Georges de Rham) is a tool belonging both to algebraic topology and to differential topology, capable of expressing basic topological information about smooth manifolds in a form particularly adapte ...
of T2 is given by harmonic (i.e. constant) ''k''-forms. ::From the Hodge decomposition on functions, Ω''k''(T2) = ker ∆''k'' \oplus im ∆''k''. Since ∆''k'' = ''dd''* + ''d''*''d'', ker ∆''k'' = ker ''d'' \cap ker ''d''*. Moreover im (''dd''* + ''d''*''d'') ⊊ im ''d'' \oplus im ''d''*. Since ker ''d'' \cap ker ''d''* is orthogonal to this direct sum, it follows that Ω''k''(T2) = ker ''d'' \cap ker ''d''* \oplus im ''d'' \oplus im ''d''*. The last assertion follows because ker ''d'' contains and is orthogonal to im ''d''* = im ∗''d''.


Hilbert space of 1-forms

In the case of the compact Riemann surface C / Λ, the theory of Sobolev spaces shows that the Hilbert space completion of smooth 1-forms can be decomposed as the sum of three pairwise orthogonal spaces, the closure of exact 1-forms ''df'', the closure of coexact 1-forms ∗''df'' and the harmonic 1-forms (the 2-dimensional space of constant 1-forms). The method of orthogonal projection of puts Riemann's approach to the Dirichlet principle on sound footing by generalizing this decomposition to arbitrary Riemann surfaces. If ''X'' is a Riemann surface Ω(''X'') denote the space of continuous 1-forms with compact support. It admits the complex inner product :\displaystyle for ''α'' and ''β'' in Ω(''X''). Let ''H'' denote the Hilbert space completion of Ω(''X''). Although ''H'' can be interpreted in terms of measurable functions, like Sobolev spaces on tori it can be studied directly using only elementary functional analytic techniques involving Hilbert spaces and bounded linear operators. Let ''H''1 denote the closure of ''d'' ''C''(''X'') and ''H''2 denote the closure of ∗''d'' ''C''(''X''). Since , these are orthogonal subspaces. Let ''H''0 denote the orthogonal complement (''H''1 \oplus ''H''2) = ''H'' \cap ''H''. Theorem (Hodge−Weyl decomposition). ''H'' = ''H''0 \oplus ''H''1 \oplus ''H''2. The subspace ''H''0 consists of square integrable harmonic 1-forms on ''X'', i.e. 1-forms ''ω'' such that ''dω'' = 0, ''d''∗''ω'' = 0 and , , ''ω'', , 2 = ∫''X'' ''ω'' ∧ ∗ < ∞. *Every square integrable continuous 1-form lies in ''H''. ::The space of continuous 1-forms of compact support is contained in the space of square integrable continuous 1-forms. They are both inner product spaces for the above inner product. So it suffices to show that any square integrable continuous 1-form can be approximated by continuous 1-forms of compact support. Let ''ω'' be a continuous square integrable 1-form, Thus the positive density Ω = ''ω'' ∧ ∗ is integrable and there are continuous functions of compact support ''ψn'' with 0 ≤ ''ψn'' ≤ 1 such that ∫''X'' ''ψn'' Ω tends to ∫''X'' Ω = , , ''ω'', , 2. Let , a continuous function of compact support with . Then ''ωn'' = ''φn'' ⋅ ω tends to ω in ''H'', since , , ''ω'' − ''ωn'', , 2 = ∫''X'' (1 − ''ψn'') Ω tends to 0. *If ω in ''H'' is such that ''ψ'' ⋅ ''ω'' is continuous for every ''ψ'' in ''C''''c''(''X''), then ''ω'' is a square integrable continuous 1-form. ::Note that the multiplication operator ''m''(''φ'') given by ''m''(''φ'')''α'' = ''φ'' ⋅ ''α'' for φ in ''C''''c''(''X'') and α in Ω(''X'') satisfies , , ''m''(φ)α, , ≤ , , φ, , , , α, , , where , , φ, , = sup , φ, . Thus ''m''(φ) defines a bounded linear operator with operator norm , , ''m''(φ), , ≤ , , φ, , . It extends continuously to a bounded linear operator on ''H'' with the same operator norm. For every open set ''U'' with compact closure, there is a continuous function φ of compact support with 0 ≤ φ ≤ 1 with φ ≅ 1 on ''U''. Then φ ⋅ ω is continuous on ''U'' so defines a unique continuous form ω''U'' on ''U''. If ''V'' is another open set intersecting ''U'', then ω''U'' = ω''V'' on ''U'' \cap ''V'': in fact if ''z'' lies in ''U'' \cap ''V'' and ψ in ''C''''c''(''U'' \cap ''V'') ⊂ ''C''''c''(''X'') with ''ψ'' = 1 near ''z'', then ''ψ'' ⋅ ''ωU'' = ''ψ'' ⋅ ''ω'' = ''ψ'' ⋅ ''ωV'', so that ''ωU'' = ''ωV'' near ''z''. Thus the ω''U'''s patch together to give a continuous 1-form ω0 on ''X''. By construction, ''ψ'' ⋅ ''ω'' = ''ψ'' ⋅ ''ω''0 for every ''ψ'' in ''C''''c''(''X''). In particular for ''φ'' in ''C''''c''(''X'') with , ∫ ''φ'' ⋅ ''ω''0 ∧ ∗ = , , ''φ''1/2 ⋅ ''ω''0, , 2 = , , ''φ''1/2 ⋅ ''ω'', , 2 ≤ , , ω, , 2. So ω0 ∧ ∗ is integrable and hence ω0 is square integrable, so an element of ''H''. On the other hand ω can be approximated by ''ωn'' in Ω(''X''). Take ψ''n'' in ''C''''c''(''X'') with 0 ≤ ψ''n'' ≤ 1 with . Since real-valued continuous functions are closed under lattice operations. it can further be assumed that ∫ ''ψ'' ''ω''0 ∧ ∗, and hence ∫ ''ψn'' ''ω''0 ∧ ∗, increase to , , ''ω''0, , 2. But then , , ψ''n'' ⋅ ω − ω, , and , , ψ''n'' ⋅ ω0 − ω0, , tend to 0. Since , this shows that . *Every square integrable harmonic 1-form ω lies in ''H''0. ::This is immediate because ω lies in ''H'' and, for ''f'' a smooth function of compact support, and . *Every element of ''H''0 is given by a square integrable harmonic 1-form. ::Let ω be an element of ''H''0 and for fixed ''p'' in ''X'' fix a chart ''U'' in ''X'' containing ''p'' which is conformally equivalent by a map ''f'' to a disc ''D'' ⊂ T2 with ''f''(0) = ''p''. The identification map from Ω(''U'') onto Ω(''D'') and hence into Ω1(T2) preserves norms (up to a constant factor). Let ''K'' be the closure of Ω(''U'') in ''H''. Then the above map extends uniquely to an isometry ''T'' of ''K'' into ''H''0(T2)''dx'' \oplus ''H''0(T2)''dy''. Moreover if ''ψ'' is in ''C''(''U'') then . The identification map ''T'' is also compatible with ''d'' and the Hodge star operator. Let ''D''1 be a smaller concentric disk in T2 and set ''V'' = ''f''(''V''). Take ''φ'' in ''C''(''U'') with ''φ'' ≡ 1 on ''V''. Then (''m''(''φ'') ''ω'',''dh'') = 0 = (''m''(''φ'') ''ω'',∗''dh'') for ''h'' in ''C''(''V''). Hence, if ''ω''1 = ''m''(''φ'')''ω'' and ''ω''2 = ''T''(''ω''1), then (''ω''2, ''dg'') = 0 = (''ω''2, ∗''dg'') for ''g'' in . ::Write ''ω''2 = ''a'' ''dx'' + ''b'' ''dy'' with ''a'' and ''b'' in ''H''0(T2). The conditions above imply (''dω''1, ∗''g'') = 0 = (''d''∗ ''ω''1, ∗''g''). Replacing ∗''g'' by ''dω''3 with ''ω''3 a smooth 1-form supported in ''D''1, it follows that ∆1 ''ω''2 = 0 on ''D''1. Thus ∆''a'' = 0 = ∆''b'' on ''D''1. Hence by Weyl's lemma, ''a'' and ''b'' are harmonic on ''D''1. In particular both of them, and hence ω2, are smooth on ''D''1; and ''dω''2 = 0 = ''d''∗''ω''2 on ''D''1. Transporting these equations back to ''X'', it follows that ''ω''1 is smooth on ''V'' and ''dω''1 = 0 = ''d''∗''ω''1 on ''V''. Since ''ω''1 = ''m''(''φ'')''ω'' and ''p'' was an arbitrary point, this implies in particular that ''m''(''ψ'')''ω'' is continuous for every ''ψ'' in ''C''''c''(''X''). So ''ω'' is continuous and square integrable. ::But then ω is smooth on ''V'' and ''d''ω = 0 = ''d''∗ω on ''V''. Again since ''p'' was arbitrary, this implies ω is smooth on ''X'' and ''d''ω = 0 = ''d''∗ω on ''X'', so that ω is a harmonic 1-form on ''X''. From the formulas for the Dolbeault operators \partial and \bar, it follows that :\displaystyle where both sums are orthogonal. The two subspaces in the second sum correspond to the ±''i'' eigenspaces of the Hodge ∗ operator. Denoting their closures by ''H''3 and ''H''4, it follows that ''H'' = ''H''3 ⊕ ''H''4 and that these subspaces are interchanged by complex conjugation. The smooth 1-forms in ''H''1, ''H''2, ''H''3 or ''H''4 have a simple description. *A smooth 1-form in ''H''1 has the form ''df'' for ''f'' smooth. *A smooth 1-form in ''H''2 has the form ∗''df'' for ''f'' smooth. *A smooth 1-form in ''H''3 has the form \partial''f'' for ''f'' smooth. *A smooth 1-form in ''H''3 has the form \bar''f'' for ''f'' smooth. ::In fact, in view of the decompositions of ''H'' and its invariance under the Hodge star operation, it suffices to prove the first of these assertions. Since ''H''1 is invariant under complex conjugation, it may be assumed that α is a smooth real 1-form in ''H''1. It is therefore a limit in ''H''1 of forms ''df''''n'' with ''f''''n'' smooth of compact support. The 1-form α must be closed since, for any real-valued ''f'' in ''C''(''X''), ::: \int_X f \, d\alpha = -\int_X df \wedge \alpha= (\alpha,*df)=0, ::so that ''d''α = 0. To prove that α is exact it suffices to prove that ∫''X'' α ∧ ∗β = 0 for any smooth closed real 1-form β of compact support. But by Green's formula :::\int_X \alpha \wedge \star \beta = (\alpha,\beta) = \lim (df_n,\beta)= \lim \int_X df_n \wedge \beta=\lim \int_X d(f_n\beta)=0. The above characterisations have an immediate corollary: *A smooth 1-form α in ''H'' can be decomposed uniquely as ''α'' = ''da'' + ∗''db'' = ''∂f'' + ''∂g'', with ''a'', ''b'', ''f'' and ''g'' smooth and all the summands square integrable. Combined with the previous Hodge–Weyl decomposition and the fact that an element of ''H''0 is automatically smooth, this immediately implies: Theorem (smooth Hodge–Weyl decomposition). If α is a smooth square integrable 1-form then α can be written uniquely as with ''ω'' harmonic, square integrable and smooth with square integrable differentials.


Holomorphic 1-forms with a double pole

The following result—reinterpreted in the next section in terms of harmonic functions and the Dirichlet principle—is the key tool for proving the
uniformization theorem In mathematics, the uniformization theorem states that every simply connected Riemann surface is conformally equivalent to one of three Riemann surfaces: the open unit disk, the complex plane, or the Riemann sphere. The theorem is a generali ...
for simply connected, or more generally planar, Riemann surfaces. Theorem. If ''X'' is a Riemann surface and ''P'' is a point on ''X'' with local coordinate ''z'', there is a unique holomorphic differential 1-form ''ω'' with a double pole at ''P'', so that the singular part of ''ω'' is ''z''−2''dz'' near ''P'', and regular everywhere else, such that ''ω'' is square integrable on the complement of a neighbourhood of ''P'' and the real part of ''ω'' is exact on ''X'' \ . The double pole condition is invariant under holomorphic coordinate change ''z'' \mapsto ''z'' + ''az''2 + ⋯. There is an analogous result for poles of order greater than 2 where the singular part of ''ω'' has the form ''z''–''k''''dz'' with ''k'' > 2, although this condition is not invariant under holomorphic coordinate change. ::To prove uniqueness, note that if ''ω''1 and ''ω''2 are two solutions then their difference ''ω'' = ''ω''1 − ''ω''2 is a square integrable holomorphic 1-form which is exact on ''X'' \ . Thus near ''P'', with ''f'' holomorphic near ''z'' = 0. There is a holomorphic function ''g'' on ''X'' \ such that ''ω'' = ''dg'' there. But then ''g'' must coincide with a primitive of ''f'' near ''z'' = 0, so that ''ω'' = ''dg'' everywhere. But then ''ω'' lies in ''H''0 ∩ ''H''1 = (0), i.e. ''ω'' = 0. ::To prove existence, take a bump function 0 ≤ ''ψ'' ≤ 1 in ''C''(''X'') with support in a neighbourhood of ''P'' of the form , ''z'', < ''ε'' and such that ''ψ'' ≡ 1 near ''P''. Set :::\alpha =-d \left(\right), ::so that ''α'' equals ''z''–2''dz'' near ''P'', vanishes off a neighbourhood of ''P'' and is exact on ''X'' \ . Let ''β'' = ''α'' − ''i''∗''α'', a smooth (0,1) form on ''X'', vanishing near ''z'' = 0, since it is a (1,0) form there, and vanishing off a larger neighbourhood of ''P''. By the smooth Hodge−Weyl decomposition, ''β'' can be decomposed as ''β'' = ''ω''0 + ''da'' – ''i''∗''da'' with ''ω''0 a harmonic and square integrable (0,1) form and ''a'' smooth with square integrable differential. Now set ''γ'' = ''α'' – ''da'' = ''ω''0 + ''i''∗''α'' − ''i''∗''da'' and ''ω'' = Re ''γ'' + ''i''∗ Re ''γ''. Then α is exact on ''X'' \ ; hence so is γ, as well as its real part, which is also the real part of ω. Near ''P'', the 1-form ω differs from ''z''–2''dz'' by a smooth (1,0) form. It remains to prove that \barω = 0 on ''X'' \ ; or equivalently that Re γ is harmonic on ''X'' \ . In fact γ is harmonic on ''X'' \ ; for ''d''γ = ''d''α − ''d''(''da'') = 0 on ''X'' \ because α is exact there; and similarly ''d''∗''γ'' = 0 using the formula ''γ'' = ''ω''0 + ''i''∗''α'' − ''i''∗''da'' and the fact that ''ω''0 is harmonic. Corollary of proof. If ''X'' is a Riemann surface and ''P'' is a point on ''X'' with local coordinate ''z'', there is a unique real-valued 1-form ''δ'' which is harmonic on ''X'' \ such that ''δ'' – Re ''z''−2''dz'' is harmonic near ''z'' = 0 (the point ''P'') such that δ is square integrable on the complement of a neighbourhood of ''P''. Moreover, if ''h'' is any real-valued smooth function on ''X'' with ''dh'' square integrable and ''h'' vanishing near ''P'', then (''δ'',''dh'') = 0. ::Existence follows by taking ''δ'' = Re ''γ'' = Re ''ω'' above. Since ''ω'' = ''δ'' + ''i''∗δ, the uniqueness of ω implies the uniqueness of δ. Alternatively if δ1 and δ2 are two solutions, their difference η = δ1 – δ2 has no singularity at ''P'' and is harmonic on ''X'' \ . It is therefore harmonic in a neighbourhood of ''P'' and therefore everywhere. So η lies in ''H''0. But also η is exact on ''X'' \ ''P'' and hence on the whole of ''X'', so it also lies in ''H''1. But then it must lie in ''H''0 ∩ ''H''1 = (0), so that η = 0. Finally, if ''N'' is the closure of a neighbourhood of ''P'' disjoint from the support of ''h'' and ''Y'' = ''X'' \ ''N'', then δ, ''Y'' lies in ''H''0(''Y'') and ''dh'' lies in the space ''H''1(''Y'') so that :::(\delta,dh) =\int_X \delta \wedge \star dh = \int_Y \delta \wedge \star dh = (\delta, _Y,dh)_Y = 0.


Dirichlet's principle on a Riemann surface

Theorem. If ''X'' is a Riemann surface and ''P'' is a point on ''X'' with local coordinate ''z'', there is a unique real-valued harmonic function ''u'' on ''X'' \ such that ''u''(''z'') – Re ''z''−1 is harmonic near ''z'' = 0 (the point ''P'') such that ''du'' is square integrable on the complement of a neighbourhood of ''P''. Moreover, if ''h'' is any real-valued smooth function on ''X'' with ''dh'' square integrable and ''h'' vanishing near ''P'', then (''du'',''dh'')=0. ::In fact this result is immediate from the theorem and corollary in the previous section. The harmonic form δ constructed there is the real part of a holomorphic form ω = ''dg'' where ''g'' is holomorphic function on ''X'' with a simple pole at ''P'' with residue -1, i.e. ''g''(''z'') = –''z''−1 + ''a''0 + ''a''1''z'' + ''a''2 ''z''2 + ⋯ near ''z'' = 0. So ''u'' = - Re ''g'' gives a solution with the claimed properties since δ = −''du'' and hence (''du'',''dh'') = −(δ,''dh'') = 0. This result can be interpreted in terms of Dirichlet's principle. Let ''D''''R'' be a parametric disk , ''z'', < ''R'' about ''P'' (the point ''z'' = 0) with ''R'' > 1. Let ''α'' = −''d''(''ψz''−1), where 0 ≤ ψ ≤ 1 is a bump function supported in ''D'' = ''D''1, identically 1 near ''z'' = 0. Let α1 = −χ''D''(''z'') Re ''d''(''z''−1) where χ''D'' is the
characteristic function In mathematics, the term "characteristic function" can refer to any of several distinct concepts: * The indicator function of a subset, that is the function \mathbf_A\colon X \to \, which for a given subset ''A'' of ''X'', has value 1 at points ...
of ''D''. Let γ= Re α and ''γ''1 = Re ''α''1. Since χ''D'' can be approximated by bump functions in L2, ''γ''1 − ''γ'' lies in the real Hilbert space of 1-forms Re ''H''; similarly α1 − α lies in ''H''. Dirichlet's principle states that the distance function :''F''(''ξ'') = , , ''γ''1 − ''γ'' – ''ξ'', , on Re ''H''1 is minimised by a smooth 1-form ξ0 in Re ''H''1. In fact −''du'' coincides with the minimising 1-form: ''γ'' + ''ξ''0 = −''du''. This version of Dirichlet's principle is easy to deduce from the previous construction of ''du''. By definition ''ξ''0 is the orthogonal projection of ''γ''1 – ''γ'' onto Re ''H''1 for the real inner product Re (''η''1,''η''2) on ''H'', regarded as a real inner product space. It coincides with the real part of the orthogonal projection ω1 of α1 – α onto ''H''1 for the complex inner product on ''H''. Since the Hodge star operator is a unitary map on ''H'' swapping ''H''1 and ''H''2, ''ω''2 = ∗''ω''1 is the orthogonal projection of ∗(''α''1 – ''α'') onto ''H''2. On the other hand, ∗''α''1 = −''i'' ''α''1, since ''α'' is a (1,0) form. Hence :(''α''1 – ''α'') − ''i''∗(''α''1 – ''α'') = ''ω''0 + ''ω''1 + ''ω''2, with ''ωk'' in ''H''''k''. But the left hand side equals –''α'' + ''i''∗''α'' = −''β'', with ''β'' defined exactly as in the preceding section, so this coincides with the previous construction. Further discussion of Dirichlet's principle on a Riemann surface can be found in , , , , and . Historical note. proved the existence of the harmonic function ''u'' by giving a direct proof of Dirichlet's principle. In , he presented his method of orthogonal projection which has been adopted in the presentation above, following , but with the theory of Sobolev spaces on T2 used to prove elliptic regularity without using measure theory. In the expository texts and , both authors avoid invoking results on measure theory: they follow Weyl's original approach for constructing harmonic functions with singularities via Dirichlet's principle. In Weyl's method of orthogonal projection, Lebesgue's theory of integration had been used to realise Hilbert spaces of 1-forms in terms of measurable 1-forms, although the 1-forms to be constructed were smooth or even analytic away from their singularity. In the preface to , referring to the extension of his method of orthogonal projection to higher dimensions by , Weyl writes: ::"Influenced by Kodaira's work, I have hesitated a moment as to whether I should not replace the Dirichlet principle by the essentially equivalent "method of orthogonal projection" which is treated in a paper of mine. But for reasons the explication of which would lead too far afield here, I have stuck to the old approach." In , after giving a brief exposition of the method of orthogonal projection and making reference to Weyl's writings, Kodaira explains: ::"I first planned to prove Dirichlet's Principle using the method of orthogonal projection in this book. However, I did not like to have to use the concept of Lebesgue measurability only for the proof of Dirichlet's Principle and therefore I rewrote it in such a way that I did not have to." The methods of Hilbert spaces, ''L''''p'' spaces and measure theory appear in the non-classical theory of Riemann surfaces (the study of
moduli space In mathematics, in particular algebraic geometry, a moduli space is a geometric space (usually a scheme (mathematics), scheme or an algebraic stack) whose points represent algebro-geometric objects of some fixed kind, or isomorphism classes of suc ...
s of Riemann surfaces) through the
Beltrami equation In mathematics, the Beltrami equation, named after Eugenio Beltrami, is the partial differential equation : = \mu . for ''w'' a complex distribution of the complex variable ''z'' in some open set ''U'', with derivatives that are locally L2 func ...
and Teichmüller theory.


Holomorphic 1-forms with two single poles

Theorem. Given a Riemann surface ''X'' and two distinct points ''A'' and ''B'' on ''X'', there is a holomorphic 1-form on ''X'' with simple poles at the two points with non-zero residues having sum zero such that the 1-form is square integrable on the complement of any open neighbourhoods of the two points. The proof is similar to the proof of the result on holomorphic 1-forms with a single double pole. The result is first proved when ''A'' and ''B'' are close and lie in a parametric disk. Indeed, once this is proved, a sum of 1-forms for a chain of sufficiently close points between ''A'' and ''B'' will provide the required 1-form, since the intermediate singular terms will cancel. To construct the 1-form for points corresponding to ''a'' and ''b'' in a parametric disk, the previous construction can be used starting with the 1-form : \alpha = d\left (\psi(z) \log\right), which locally has the form : - .


Poisson equation

Theorem (Poisson equation). If Ω is a smooth 2-form of compact support on a Riemann surface ''X'', then Ω can be written as Ω = ∆''f'' where ''f'' is a smooth function with ''df'' square integrable if and only if ∫''X'' Ω = 0. ::In fact, Ω can be written as Ω = ''d''α with α a smooth 1-form of compact support: indeed, using partitions of unity, this reduces to the case of a smooth 2-form of compact support on a rectangle. Indeed Ω can be written as a finite sum of 2-forms each supported in a parametric rectangle and having integral zero. For each of these 2-forms the result follows from Poincaré's lemma with compact support. Writing ''α'' = ''ω'' + ''da'' + *''db'', it follows that Ω = ''d''*''db'' = ∆''b''. In the case of the simply connected Riemann surfaces C, ''D'' and ''S''= C ∪ ∞, the Riemann surfaces are
symmetric space In mathematics, a symmetric space is a Riemannian manifold (or more generally, a pseudo-Riemannian manifold) whose group of isometries contains an inversion symmetry about every point. This can be studied with the tools of Riemannian geome ...
s ''G'' / ''K'' for the groups ''G'' = R2, SL(2,R) and SU(2). The methods of group representation theory imply the operator ∆ is ''G''-invariant, so that its fundamental solution is given by right convolution by a function on ''K'' \ ''G'' / ''K''. Thus in these cases Poisson's equation can be solved by an explicit integral formula. It is easy to verify that this explicit solution tends to 0 at ∞, so that in the case of these surfaces there is a solution ''f'' tending to 0 at ∞. proves this directly for simply connected surfaces and uses it to deduce the
uniformization theorem In mathematics, the uniformization theorem states that every simply connected Riemann surface is conformally equivalent to one of three Riemann surfaces: the open unit disk, the complex plane, or the Riemann sphere. The theorem is a generali ...
.


See also

* Differentials of the first kind * Abelian differential * Dolbeault complex


Notes


References

* * * *
2016 reprint
* * * * * * * * * *, 1989 reprint of 1941 edition with foreword by
Michael Atiyah Sir Michael Francis Atiyah (; 22 April 1929 – 11 January 2019) was a British-Lebanese mathematician specialising in geometry. His contributions include the Atiyah–Singer index theorem and co-founding topological K-theory. He was awarded the ...
*, reprint of 1941 edition incorporating corrections supplied by
Hermann Weyl Hermann Klaus Hugo Weyl (; ; 9 November 1885 – 8 December 1955) was a German mathematician, theoretical physicist, logician and philosopher. Although much of his working life was spent in Zürich, Switzerland, and then Princeton, New Jersey, ...
* *, Part III, Chapter 8: "Die Verallgemeinerung des Riemannschen Abbildungssatzes. Das Dirichletsche Prinzlp," by
Richard Courant Richard Courant (January 8, 1888 – January 27, 1972) was a German-American mathematician. He is best known by the general public for the book '' What is Mathematics?'', co-written with Herbert Robbins. His research focused on the areas of real ...
* * * * * * * * * * * * * * * * * *{{citation, last=Weyl, first= Hermann, title= The concept of a Riemann surface, translator= Gerald R. MacLane, publisher=Addison-Wesley, year= 1955, mr=0069903 Harmonic functions Riemann surfaces Differential forms