Fact-checked by Grok 2 weeks ago

Hyperfunction

In , hyperfunctions are a class of generalized functions introduced by in 1959–1960, defined as boundary values of holomorphic functions taken across a real in complex Euclidean space, extending the scope beyond traditional distributions by embedding the real space into a complexified domain without relying on limits or dual spaces of smooth functions. This construction leverages sheaf cohomology to ensure local exactness, forming a flabby sheaf that captures analytic singularities and supports operations like , , and multiplication by holomorphic functions. Sato's innovation arose in the mid-1950s amid the dominance of Laurent Schwartz's distribution theory for solving partial differential equations (PDEs), but sought to address limitations in handling analytic phenomena, such as the unsolvability highlighted by Hans Lewy's example of a PDE with smooth coefficients lacking smooth solutions. Unlike distributions, which rely on Fourier transforms and test functions, hyperfunctions treat generalized functions as direct "jumps" between holomorphic extensions, enabling precise control over singularities via local cohomology groups—a concept Sato developed concurrently with Alexander Grothendieck's work. This approach proved particularly powerful for linear PDEs with analytic coefficients, allowing solutions to be represented as hyperfunctions and facilitating computations free from the topological constraints of classical . The theory's impact extended rapidly, laying the groundwork for in the late 1960s and early 1970s, where Sato, along with collaborators Masaki Kashiwara and Takahiro Kawai, introduced microfunctions to localize singularities on the , revolutionizing the study of PDE propagation and pseudo-differential operators. Hyperfunctions also underpin and D-module theory, providing tools for holonomic systems and , with applications in physics for modeling wave propagation and scattering. By the 1970s, the framework had influenced international mathematics, as evidenced by its presentation at the 1970 and subsequent monographs, establishing hyperfunctions as a cornerstone of modern analytic methods.

Introduction

Basic Concept

Hyperfunctions represent a class of generalized functions introduced to address singularities in solutions to partial differential equations that exceed the capabilities of classical distributions, such as those arising from boundary values of holomorphic functions across real hypersurfaces. They extend the framework of distributions by incorporating discontinuities or "jumps" along real submanifolds, enabling the rigorous treatment of objects like the delta function as prototypical examples within this broader structure. In one complex variable, a hyperfunction on an open interval \Omega \subset \mathbb{R} is defined as an equivalence class of holomorphic functions on the complex domain V \setminus \Omega, where V is an open neighborhood of \Omega in \mathbb{C}. Specifically, two such functions F, G \in \mathcal{O}(V \setminus \Omega) (the space of holomorphic functions) are equivalent if F - G extends holomorphically to the full V. This construction typically involves holomorphic functions f_+ on the upper half-plane and f_- on the lower half-plane, with the hyperfunction B_F given by the boundary value difference B_F = \lim_{\epsilon \to 0^+} \left( f_+(x + i\epsilon) - f_-(x - i\epsilon) \right), where the limit is taken in the sense of distributions. These functions must satisfy prerequisite conditions, including holomorphic extendibility across the real axis except on \Omega and suitable growth conditions at infinity ensuring the boundary values exist in the sense of distributions, such as controlled growth in the imaginary direction locally. A key result establishing their analytic character is a Paley-Wiener-type : hyperfunctions with compact in one variable have Fourier transforms that are entire functions of exponential type, precisely bounding the growth and linking their spectral properties to holomorphic extensions. Hyperfunctions thus generalize , as every distribution can be embedded into the space of hyperfunctions via with analytic functions, but the converse does not hold due to the enhanced ability to capture singular behaviors along the real line.

Historical Development

The theory of hyperfunctions originated in the late 1950s through the work of Japanese mathematician , who sought to extend the framework of generalized functions beyond the limitations of . Building on earlier concepts of analytic functionals developed by Luigi Fantappiè in and , Sato first announced hyperfunctions in a 1958 Japanese publication. This foundational idea was influenced by Laurent Schwartz's theory from the , particularly his 1945-1951 "Théorie des distributions," which provided tools for handling singularities in real variables but struggled with non-tempered growth. Additionally, Jean Leray's 1950s contributions to and complex singularities inspired Sato's shift toward complex structures to address these shortcomings. In the early 1960s, Sato formalized the theory in English through his seminal papers "Theory of Hyperfunctions, I" (1959) and "Theory of Hyperfunctions, II" (1960), published in the Journal of the Faculty of Science, University of Tokyo, where he demonstrated that hyperfunctions encompass Schwartz's distributions while extending to non-tempered singularities via local cohomology in complex space. These works marked a pivotal milestone, establishing hyperfunctions as a module over analytic functions with operations like differentiation. By 1962, Sato had further refined the framework, integrating it with sheaf theory—a tool popularized by Alexander Grothendieck in the 1950s—to handle local properties and enable generalization to several complex variables. Collaborators such as Akira Kaneko contributed to early expositions and applications in the 1960s, while Takahiro Kawai assisted in advancing the multivariable aspects through sheaf cohomology. The 1970s saw hyperfunction theory evolve into a cornerstone of , with Sato's 1970 presentation "Regularity of Hyperfunction Solutions of Partial Differential Equations" at the in highlighting its role in describing singularity structures. This integration culminated in the 1973 collaborative work "Microfunctions and Pseudo-Differential Equations" by Sato, Kawai, and Masaki Kashiwara, which embedded hyperfunctions into the framework, influencing subsequent developments in . Through these milestones, hyperfunctions bridged distribution theory's real-analytic focus with , providing a robust tool for singularities beyond tempered growth.

Mathematical Formulation

In One Complex Variable

In the case of one complex variable, hyperfunctions on an open subset \Omega \subset \mathbb{R} are defined as equivalence classes of pairs (F^+, F^-), where F^+ is a on an in the upper half-plane containing \Omega, and F^- is holomorphic on the corresponding in the lower half-plane. Two such pairs (F^+, F^-) and (G^+, G^-) are equivalent, representing the same hyperfunction, if there exists a holomorphic function H on a complex neighborhood of \Omega such that F^+ - G^+ = H on the upper part and F^- - G^- = H on the lower part; in other words, the differences extend analytically across the real line. This construction generalizes distributions by allowing the "jump" across \Omega to capture more singular behaviors. The hyperfunction is concretely realized through its boundary value representation along the real line. For a representative pair (F^+, F^-), the associated hyperfunction BF is given by the distributional limit BF(x) = \lim_{\epsilon \to 0^+} \left[ F^+(x + i\epsilon) - F^-(x - i\epsilon) \right], where the limit is understood in the sense of distributions on \Omega. For this limit to exist, the functions F^\pm must satisfy suitable growth conditions near the real axis, such as |F^\pm(z)| \leq C (1 + |z|)^N \exp\left( \frac{\pi |\Im z|}{\epsilon} \right) for constants C, N > 0 and every \epsilon > 0, ensuring the convergence despite potentially rapid growth as \Im z \to 0. This boundary value formulation embeds the hyperfunction as the "difference" of the analytic continuations from each side. Hyperfunctions contain the space of distributions as a , with the embedding preserving s. Locally on \Omega, every hyperfunction can be decomposed as the sum of a smooth (which is a ) and a hyperfunction with compact . For hyperfunctions with compact support, the structure theorem provides a explicit finite-dimensional description: such a hyperfunction is a finite of distributional derivatives of the \delta(x) and of the principal value distribution \mathrm{Pv}(1/x). This theorem highlights the finite nature of the singular components at each point, distinguishing hyperfunctions from more general ultradistributions.

In Several Complex Variables

In several complex variables, the concept of hyperfunctions generalizes the one-variable theory to real analytic manifolds of arbitrary dimension n, relying on sheaf cohomology rather than explicit boundary representations. Introduced by Mikio Sato, the sheaf of hyperfunctions \mathcal{B}_M on a real analytic manifold M embedded as a totally real submanifold in its complexification X = M \times \mathbb{C} (or more generally a complex neighborhood) is defined as the direct image under the inclusion i: M \hookrightarrow X of the local cohomology sheaf \underline{H}^n_M(\mathcal{O}_X) \otimes_{\mathcal{Z}_M} \mathrm{or}_{M/X}, where \mathcal{O}_X is the sheaf of holomorphic functions on X, \mathcal{Z}_M is the constant sheaf \mathbb{Z} on M, and \mathrm{or}_{M/X} is the relative orientation sheaf accounting for the codimension n embedding. Sections of \mathcal{B}_M over an open set U \subset M are thus cohomology classes in H^n(U, U \setminus V; \mathcal{O}_X) (or relative versions thereof), where V is a suitable complex neighborhood of U, capturing "jumps" across the real submanifold in a cohomological sense. This construction ensures that \mathcal{B}_M is a flabby sheaf, allowing global sections to be computed locally, and it forms a module over the sheaf of real analytic functions on M. Unlike the one-variable case, where hyperfunctions admit direct representations as differences of holomorphic functions from upper and lower half-planes, the several-variables extension loses such explicit boundary formulas due to the higher codimension of the real locus (codimension n in \mathbb{C}^n) and instead depends on derived functors of cohomology. A representational approach in \mathbb{C}^n views hyperfunctions on open sets \Omega \subset \mathbb{R}^n as boundary values of holomorphic functions on tube domains T^C = \Omega + iC, where C \subset \mathbb{R}^n is a convex cone ensuring the domain is a tube over \Omega; these boundary values are taken in the sense of relative cohomology to handle the non-simply connected structure. This leads to an exact sequence $0 \to \mathcal{O} \to \mathcal{B}^+ \oplus \mathcal{B}^- \to \mathcal{B} \to 0, where \mathcal{O} is the sheaf of holomorphic functions on the full complex domain, \mathcal{B}^+ (resp. \mathcal{B}^-) consists of hyperfunctions extending holomorphically to tube domains over the "positive" (resp. "negative") cones, and the map to \mathcal{B} quotients by the holomorphic functions common to both sides. A fundamental result in this theory is that every hyperfunction on M has its support contained in a real analytic subvariety of M, with the singular support (or microsupport) localized along real analytic hypersurfaces defining the boundaries of the tube domains; specifically, for a hyperfunction represented via a defining pair of holomorphic functions on adjacent tube domains separated by a real analytic hypersurface S, the support is the zero set of the real analytic function describing S, intersected with the singularities of the extension. This localization arises from the coherence of \mathcal{B}_M over the structure sheaf of real analytic functions and the fact that M is purely n-codimensional with respect to \mathcal{O}_X. The reliance on derived functors introduces the role of microsupport in the cotangent bundle T^*M, which refines singularity localization beyond global supports and highlights the abstract nature of the theory compared to the one-variable setting, where singularities are simply points on the real line.

Examples

Elementary Examples

One of the simplest hyperfunctions is the , which extends the distributional delta beyond standard test function duality by representing it as a boundary value in the . Specifically, it is given by \delta(x) = \frac{1}{2\pi i} \lim_{\epsilon \to 0^+} \left( \frac{1}{x + i \epsilon} - \frac{1}{x - i \epsilon} \right). This expression derives from the jump discontinuity across the real axis of the function $1/z, illustrating how hyperfunctions formalize such singularities as differences of holomorphic boundary values. The H(x), a classic , acquires a hyperfunction through its boundary value representation, emphasizing its jump at the despite being locally integrable elsewhere. It is expressed as H(x) = \lim_{\epsilon \to 0^+} \frac{1}{2\pi i} \log\left( \frac{z - i \epsilon}{z + i \epsilon} \right), where the logarithm captures the shift across the real axis, confirming its embedding within the hyperfunction space while highlighting the boundary value mechanism. Another elementary case is the principal value \mathrm{Pv}(1/x), whose hyperfunction form distinguishes it from the purely distributional symmetric limit by incorporating logarithmic boundary values. It is defined as \mathrm{Pv}\left( \frac{1}{x} \right) = \lim_{\epsilon \to 0^+} \frac{1}{2\pi i} \left( \log(x + i \epsilon) - \log(x - i \epsilon) \right), where the difference in branches of the logarithm yields the singular behavior at zero, separate from the average of one-sided limits used in distribution theory. Simple linear combinations of these basic hyperfunctions further exemplify their structure, such as sums or differences that preserve the boundary value representation. For instance, near , the hyperfunction associated with e^{1/x} (extended smoothly to the left) is constructed via the boundary value of e^{1/z} in the upper half-plane paired with a suitable lower half-plane extension, underscoring non-analytic smooth behavior and the capacity of hyperfunctions to handle essential singularities without relying on distributional growth restrictions.

Advanced Examples

In several complex variables, a key example is the hyperfunction associated with the real line in \mathbb{C}^2, defined as the sheaf of sections of the relative cohomology sheaf \mathcal{B}/\mathcal{O} over the real \mathbb{R}^2 \subset \mathbb{C}^2. This hyperfunction captures the boundary values of holomorphic functions on the upper and lower half-spaces separated by the real line, with its singularity concentrated along the diagonal \{(z,w) \mid z = \bar{w}\}, reflecting the intrinsic of the real in higher dimensions. A classic non-distributional hyperfunction is defined by e^{1/x^2} for x < 0 and 0 for x > 0. The boundary value computation from the upper half-plane involves the holomorphic extension, revealing an at x = 0 due to the explosive as x \to 0^-, which exceeds any bound and thus cannot arise from a but fits within the hyperfunction class via difference of boundary values.

Properties

Analytic Properties

Hyperfunctions possess a close relationship to holomorphic functions through their representation as values. In one complex variable, a hyperfunction on an open interval \Omega \subset \mathbb{R} can be expressed as the difference BF = h^+ - h^-, where h^+ and h^- are the values of holomorphic functions defined on the upper half-plane tuboid \Omega + i(0,\infty) and the lower half-plane tuboid \Omega + i(-\infty,0), respectively. This representation extends to several variables, where hyperfunctions on \Omega \subset \mathbb{R}^n are values of holomorphic functions on suitable tuboids over \Omega. A fundamental analytic is the analytic theorem, which asserts that every hyperfunction defined on an \Omega admits a meromorphic to a neighborhood of \Omega in the complex domain, with possible poles confined to the real . This is achieved via the boundary value representation, ensuring that the hyperfunction corresponds to a holomorphic off the real axis. The ness follows from the quotient structure in the definition of hyperfunctions as sheaves of holomorphic functions on the complement of \Omega. Uniqueness of hyperfunctions is further guaranteed by an adaptation of the edge-of-the-wedge theorem. Specifically, if two hyperfunctions agree on an open subset of their and their supports coincide, then they agree globally on the entire . This result relies on the local determination property of hyperfunctions as a sheaf and the compatibility of their singularity spectra, ensuring no extraneous extensions across wedges. In terms of growth estimates, the holomorphic functions representing hyperfunctions exhibit sub-exponential growth in the complex directions away from the real axis. For a representing holomorphic function F on a tuboid V \setminus \Omega, the growth is bounded by \sup_{x \in K} |F(x + i t)| \leq C \exp G(|t|), where G is a growth function of class *, satisfying G(t)/t^\epsilon \to 0 as t \to \infty for every \epsilon > 0. This contrasts with the polynomial growth bounds for distributions, allowing hyperfunctions to capture more singular behaviors while maintaining analytic control.

Support and Singularities

The of a hyperfunction BF on an \Omega \subset \mathbb{R}^n is defined as the complement of the largest open subset of \Omega on which BF vanishes identically. This makes it the smallest K \subset \Omega such that BF extends by zero to \Omega \setminus K. Unlike distributions, whose supports can be arbitrary closed sets, the of a hyperfunction is always contained in some real analytic subvariety V of \Omega. This containment reflects the real analytic nature of the test functions to hyperfunctions and ensures that hyperfunctions are real analytic outside their supports. Singularities of hyperfunctions arise solely along real hypersurfaces within their supports, distinguishing them from more general generalized functions. Specifically, if BF is a hyperfunction on \Omega, its singularities are confined to points where the defining holomorphic functions fail to extend analytically across the real boundary, and the theorem asserts that \operatorname{supp} BF \subset V for some real analytic set V. This structure implies that hyperfunctions exhibit real analytic behavior away from these hypersurfaces, with jumps or discontinuities localized precisely on them. For instance, the delta function, as a hyperfunction, has its singularity exactly on the real hyperplane \{x=0\}, contained in the trivial real analytic variety \{0\}. To capture the directional nature of these singularities, the microlocal support is described by the wavefront set \operatorname{WF}(BF) \subset T^* \mathbb{R}^n \setminus 0, which consists of points (x, \xi) where BF fails to be microlocally real analytic in the direction \xi. This set projects onto the singular support \operatorname{SS}(BF), the locus of spatial singularities, and is defined via the absence of suitable boundary value representations with conic supports avoiding certain half-spaces in the cotangent space. For hyperbolic partial differential operators, singularities in hyperfunction solutions propagate along bicharacteristic curves in the wavefront set. A representative example occurs in solutions to the wave equation \partial_t^2 u - \Delta u = 0, where initial data with a at a point (t,x) = (0,0) lead to hyperfunction solutions whose set propagates along the \{(t,x) : |x| = |t|\}, spreading singularities at the while remaining real analytic outside this cone. This propagation aligns with the characteristic variety of the operator, ensuring that the microlocal structure respects the geometry of the bicharacteristics.

Operations

Linear Operations

Hyperfunctions on an open interval I \subset \mathbb{R} form a complex vector space B(I), where elements are equivalence classes of pairs of holomorphic functions (F^+, F^-) with F^+ holomorphic in the upper half-neighborhood and F^- in the lower, representing the hyperfunction as the difference of their boundary values. Addition is defined componentwise: for hyperfunctions BF = F^+ - F^- and BG = G^+ - G^-, (BF + BG)^+ = F^+ + G^+ and (BF + BG)^- = F^- + G^-. Scalar multiplication by c \in \mathbb{C} is similarly (c \cdot BF)^+ = c F^+ and (c \cdot BF)^- = c F^-. These operations are well-defined on equivalence classes and endow B(I) with the structure of a complex vector space. The space of hyperfunctions also forms a over the ring of smooth functions C^\infty(I). For \phi \in C^\infty(I) and hyperfunction BF = F^+ - F^-, the product is defined by (\phi \cdot BF)^+ = \phi F^+ and (\phi \cdot BF)^- = \phi F^-, where multiplication by the smooth \phi is performed after if necessary. This action is compatible with the structure and extends the module structure of distributions. Convolution of two hyperfunctions F and G on \mathbb{R} is defined when their supports satisfy suitable conditions, such as one having compact support, ensuring the converges in the sense of values. Specifically, (F * G)(x) = \int_{-\infty}^\infty F(y) G(x - y) \, dy, where the is interpreted as the value of the corresponding holomorphic extension in the complex domain. A fundamental theorem states that the convolution of two hyperfunctions is again a hyperfunction, preserving the under this operation. Differentiation preserves the class of hyperfunctions. For a hyperfunction BF, where F(z) is holomorphic in the defining domain above and below the real axis, the derivative is given by \frac{d}{dx} (BF) = B \left( \frac{d F}{dz} \right), the boundary value of the complex derivative \frac{d F}{dz}. This extends the distributional derivative and applies to higher-order derivatives as well, maintaining the hyperfunction structure. The of a hyperfunction BF extends the transform for tempered distributions to non-tempered cases. It is defined as the boundary value \hat{BF}(\xi) = \widehat{F}(\xi + i\eta), where \widehat{F} is the holomorphic extension of the of F, taken as the limit from the appropriate half-plane. Equivalently, \hat{BF}(\xi) = F_+(\xi + i0) - F_-(\xi - i0), with F_+ and F_- the analytic continuations from the upper and lower half-planes, respectively. This operation maps hyperfunctions to .

Nonlinear Operations

Multiplication of a hyperfunction by a is always defined, as the of C^\infty(\Omega) embeds into the of hyperfunctions B(\Omega) on an \Omega \subset \mathbb{R}^n, since admit holomorphic extensions to neighborhoods. The product \phi \cdot F, for \phi \in C^\infty(\Omega) and F \in B(\Omega), is given pointwise by (\phi \cdot F)(x) = \phi(x) F(x), and it satisfies the over C^\infty(\Omega), meaning \langle \phi \cdot F, \psi \rangle = \langle F, \phi \psi \rangle for test \psi. The support of the product is contained in the union of the supports of \phi and F, ensuring the operation preserves the sheaf of hyperfunctions. In contrast to linear operations like , the of two general hyperfunctions F, G \in B(\Omega) is more restricted and does not always exist, reflecting the higher singularity of hyperfunctions compared to distributions. The product F \cdot G is defined as a hyperfunction only under specific microlocal conditions on their analytic wave front sets (AWFS), denoted WF_a(F) and WF_a(G), which capture the directions of singularities in the cotangent bundle T^* \Omega. The key result is the product theorem: F \cdot G \in B(\Omega) if WF_a(F) \cap WF_a(G) = \emptyset outside the conormal directions to the of \Omega in \mathbb{R}^n, ensuring the singularities do not interact destructively. For instance, if the singular supports are disjoint, the product vanishes and is well-defined as zero, but overlapping supports generally prevent the product from being a hyperfunction unless one factor is smooth or analytic. This contrasts with distributions, where products exist under weaker conditions on the classical wave front sets, but for hyperfunctions, the analytic nature imposes stricter requirements. A classic limitation is that the product of the Dirac delta with itself, \delta \cdot \delta, is undefined, as WF_a(\delta) = T_x^* \Omega \setminus \{0\} at the support point x, leading to non-empty intersection in all directions. Composition of a hyperfunction F \in B(\Omega) with a smooth function g: \Omega' \to \Omega is defined microlocally on open sets where the differential dg is invertible, i.e., g is a , allowing the pullback to respect the boundary value structure. Specifically, F \circ g is a hyperfunction on \Omega' when g extends holomorphically to a neighborhood and dg \neq 0, with the operation given by substituting the holomorphic representatives: if F = [F_+, F_-] in the upper and lower half-spaces, then F \circ g = [F_+ \circ g_+, F_- \circ g_-], modulo the of holomorphic functions on the extended domain. The formula provides explicit computation, such as for the Dirac : \delta(g(x)) = \sum_{g(a_i)=0} \frac{\delta(x - a_i)}{|g'(a_i)|}, where the sum is over simple zeros a_i of g, generalizing the coarea formula microlocally. This ensures associativity with multiplication by functions but fails globally if g has critical points, where singularities may propagate along characteristics.

Applications

In Partial Differential Equations

In the theory of hyperfunctions, fundamental solutions to linear partial differential equations with constant coefficients are constructed within the framework of boundary values of holomorphic functions. For a linear differential operator P(\partial) with constant coefficients, the fundamental solution E is a hyperfunction satisfying the equation P(\partial) E = \delta, where \delta denotes the Dirac delta distribution. This E arises as the boundary value (in the hyperfunction sense) of holomorphic solutions to the associated complex equation P(z) u = 0, defined in appropriate complex tubular neighborhoods of the real space. Such constructions ensure the existence of global fundamental solutions for hyperbolic operators on real analytic manifolds, leveraging the micro-support conditions and propagators in the D-module category. Hyperfunctions also facilitate the resolution of the for partial differential equations. When initial are prescribed as hyperfunctions on a non- , existence and uniqueness of solutions hold in the of hyperfunctions, provided the is . This well-posedness is established through isomorphisms between derived categories of D-modules restricted to the and the ambient , using to propagate solutions along bicharacteristics. For systems represented as D-modules with regular singularities, the admits a natural solution in hyperfunctions, extending classical results to more general settings. A representative example is the one-dimensional \partial_t^2 u - \partial_x^2 u = 0, a PDE with constant coefficients. Solutions with hyperfunction initial data u(0, x) = f(x) and \partial_t u(0, x) = g(x) propagate singularities along the characteristics x \pm t = constant, yielding a solution that generalizes the classical d'Alembert formula u(t, x) = \frac{1}{2} [f(x+t) + f(x-t)] + \frac{1}{2} \int_{x-t}^{x+t} g(y) \, dy to data with essential singularities or non-analytic behavior. This approach captures the sharp propagation of singularities inherent to systems. Compared to distributions, hyperfunctions offer the advantage of accommodating initial data or coefficients with essential singularities off the real axis, which distributions cannot represent without additional structure, thus enabling solutions to PDEs where classical distributional methods fail for non-tempered or highly singular data.

In Microlocal Analysis

In , the wavefront set of a hyperfunction provides a precise description of its analytic singularities in . For a hyperfunction u on a real-analytic manifold X, the analytic wavefront set \mathrm{WF}_a(u) is defined as the set of points (x, \xi) \in T^*X \setminus 0 where there exists no holomorphic function, defined in a suitable complex conic neighborhood of (x, \xi), that approximates u microlocally near (x, \xi). This non-extendability condition characterizes the directions in which the hyperfunction fails to admit local holomorphic extensions, making \mathrm{WF}_a(u) a closed conic subset of the conormal bundle to X. Pseudodifferential operators play a central role in the microlocal study of hyperfunctions, where they can be constructed using hyperfunctions as symbols or kernels. A key result establishes that properly supported pseudodifferential operators map the sheaf of hyperfunctions to itself, preserving the analytic structure and allowing for microlocal elliptic regularity. This continuity ensures that singularities of hyperfunctions under such operators are controlled by the principal symbol's characteristics. For operators, the propagation of singularities in hyperfunctions aligns with bicharacteristic flows, extending the classical Melrose-Sjöstrand to the analytic category. Specifically, if P is a pseudodifferential operator, then for a hyperfunction u to Pu = f, the analytic set satisfies \mathrm{WF}_a(u) \setminus \mathrm{WF}_a(f) is contained in the union of bicharacteristic strips issuing from \mathrm{WF}_a(f), with singularities propagating along these null bicharacteristics in the . This adaptation leverages the microlocal regularity properties of hyperfunctions via generalized FBI transforms. Hyperfunctions find significant applications in constructing parametrices for on manifolds with , where they resolve singularities arising from boundary interactions. For an on a manifold with , the parametrix can be built microlocally using hyperfunction kernels that capture the across the boundary, ensuring invertibility up to smoothing operators while accounting for conormal singularities. This approach facilitates the analysis of boundary value problems by localizing resolutions to the wavefront sets near the .

References

  1. [1]
    Mikio Sato, a Visionary of Mathematics
    Feb 1, 2007 · To understand the originality of Sato's theory of hyperfunctions, one has to place it in the mathematical landscape of the time. Mathematical.
  2. [2]
    Professor Mikio Sato and Microlocal Analysis - EMS Press
    In the summer of 1957, Sato created the theory of hyperfunctions. In formulating the idea of regarding boundary values of holomorphic functions as ...
  3. [3]
    Mikio Sato - Biography - MacTutor - University of St Andrews
    Hyperfunctions, together with integral Fourier operators, have become a major tool in linear partial differential equations. Along with his students, Sato ...<|control11|><|separator|>
  4. [4]
    [PDF] Introduction to Hyperfunctions and Their Integral Transforms
    Fortunately, some years ago, I found Isac Imai's Book Applied Hyperfunction Theory which explains and applies Sato's hyperfunctions in a concrete, but ...
  5. [5]
    [PDF] AN ELEMENTARY THEORY OF HYPERFUNCTIONS AND ...
    Sato's definition of hyperfunctions. The hyperfunctions are a class of generalized functions introduced by M. Sato [34], [35], [36] in 1958–60, only.
  6. [6]
    [PDF] Mikio Sato, a visionary of mathematics - IMJ-PRG
    It was in this environment that Sato defined hyperfunctions [Sat59] in 1959–1960 as boundary values of holomorphic functions, a discovery which allowed him to ...
  7. [7]
    [PDF] Theory of Hyperfunctions, I. - By Mikio SATO.
    § 4. Representation of a hyperfunction as the boundary value' of the defining function. § 5. Holomorphic functions as hyperfunctions.
  8. [8]
    [PDF] Theory of Hyperfunctions, II. - By Mikio SATO.
    A Summary of the Case of One Variable. § 1. Definitions. 11. Definitions. 1.2. An example. § 2. Localizability ...
  9. [9]
    Introduction to the Theory of Hyperfunctions (Mathematics and its ...
    Kaneko. Introduction to the Theory of Hyperfunctions (Mathematics and its Applications, 3). 1988th Edition. ISBN-13: 978-9027728371, ISBN-10: 9789027728371.Missing: collaborator Sato
  10. [10]
  11. [11]
    [PDF] Sato hyperfunctions via relative Dolbeault cohomology - arXiv
    Apr 5, 2022 · Sato, Theory of hyperfunctions I, II, J. Fac. Sci. Univ. Tokyo, 8 ... Faculty of Science. Hokkaido University. Sapporo 060-0810, Japan.
  12. [12]
    Introduction to Hyperfunctions
    The hyperfunction f(x) = [sin z, 0] = [sin(z) 1+(z)] has for every x the value sin x and thus interprets the ordinary function f(x) = sin x. Example 1.4.
  13. [13]
    Edge of the wedge theorem and hyperfunction - SpringerLink
    Aug 27, 2006 · 'Edge of the wedge theorem and hyperfunction' published in 'Hyperfunctions and Pseudo-Differential Equations'
  14. [14]
    Edge of the wedge theorem for tempered ultra-hyperfunctions
    The local form of the edge of the wedge theorem for hyperfunction has the following formulation ( and are related as in the above results): If in an open set , ...
  15. [15]
  16. [16]
  17. [17]
  18. [18]
    [PDF] Ordered linear space of hyperfunctions
    For every real number x 6= 0 the limit limε→0+ {F+(x+iε)−F−(x−iε)} exists and equal to 0. ... Introduce the subclass O(R+) of hyperfunctions f(x)=[F(z)] on ...Missing: boundary | Show results with:boundary
  19. [19]
    Convolution of Hyperfunctions | SpringerLink
    Convolution of Hyperfunctions. Chapter. pp 249–282; Cite this chapter. Download book PDF · Applied Hyperfunction Theory. Convolution of Hyperfunctions. Download ...
  20. [20]
    [PDF] Hyperfunctions
    Nov 3, 2023 · For simplicity we work in d = 1 from now on! Let u be a holomorphic function on the upper half plane. C. + := {z = x + iy ∈ C|y > 0} and let ...
  21. [21]
  22. [22]
    The Analysis of Linear Partial Differential Operators I - SpringerLink
    Distributions in Product Spaces. Lars Hörmander. Pages 126-132. Composition with ... Hyperfunctions. Lars Hörmander. Pages 325-370. Back Matter. Pages 371-440.
  23. [23]
    [PDF] arXiv:math/9906211v1 [math.AP] 30 Jun 1999
    Let us show how our results imply the existence of global fundamental solutions in the framework of hyperfunctions. ... [9] Mikio Sato, Theory of hyperfunctions.
  24. [24]
    Cauchy Problem for Hyperbolic D-modules with Regular Singularities
    The aim of this paper is to extend this last result to general systems of linear partial differential equations, that is, to D-modules. The assumption of having ...
  25. [25]
    [PDF] arXiv:1706.07388v1 [math.SG] 22 Jun 2017
    Jun 22, 2017 · Abstract. In this article we investigate the Duistermaat-Heckman theorem using the theory of hy- perfunctions.
  26. [26]
    Pseudo.differential Operators in the Theory of Hyperfunctions
    In this note we first give the definition of the sheaf of pseudo- differential operators, the notion of which is originally due to Sato.
  27. [27]
    [2104.11047] A new microlocal analysis of hyperfunctions - arXiv
    Apr 22, 2021 · Using a microlocal decomposition of a hyperfunction and the generalized FBI transforms we were able to characterize the wave-front set of ...