Fact-checked by Grok 2 weeks ago

Mollifier

A mollifier is a , non-negative \eta: \mathbb{R}^n \to [0, \infty) with compact support contained in the unit ball B(0,1) and satisfying \int_{\mathbb{R}^n} \eta(x) \, dx = 1. These , also known as approximations to the identity, are used in and partial differential equations (PDEs) to regularize or distributions and via , producing approximations that converge to the original in appropriate norms as the scale approaches zero. The technique of mollification traces its origins to the work of Sergei L. Sobolev in 1938, where integral operators akin to mollifiers were employed to prove embedding theorems for function spaces, laying foundational groundwork for modern theory. A canonical example of a mollifier in one dimension is given by \eta(x) = C \exp\left(\frac{1}{|x|^2 - 1}\right) for |x| < 1 and \eta(x) = 0 otherwise, where the constant C > 0 is chosen to ensure the integral equals 1; this construction extends naturally to higher dimensions using radial symmetry. For a general function f \in L^p(\mathbb{R}^n) with $1 \leq p < \infty, the mollified approximation f_\epsilon = f * \eta_\epsilon, where \eta_\epsilon(x) = \epsilon^{-n} \eta(x/\epsilon), is infinitely differentiable and satisfies \|f_\epsilon - f\|_{L^p} \to 0 as \epsilon \to 0^+. Mollifiers play a crucial role in establishing the density of smooth compactly supported functions (C_c^\infty) in Sobolev spaces W^{k,p}(\Omega), enabling the approximation of weak solutions to PDEs by classical smooth solutions and facilitating proofs of regularity and existence results. Beyond PDEs, they appear in harmonic analysis for studying Fourier transforms of tempered distributions, in numerical methods for regularization of ill-posed problems, and in probability for kernel density estimation, where scaled mollifiers act as smoothing kernels.

Introduction and Definition

Basic Definition

A mollifier is a family of smooth functions \{\phi_\epsilon : \mathbb{R}^n \to \mathbb{R} \mid \epsilon > 0\} used to smooth other functions via convolution, where each \phi_\epsilon(x) = \epsilon^{-n} \phi(x/\epsilon) for a fixed smooth function \phi \in C^\infty(\mathbb{R}^n) that is compactly supported in the unit ball B(0,1), non-negative (\phi \geq 0), and normalized so that \int_{\mathbb{R}^n} \phi(x) \, dx = 1. The base function \phi is often chosen to be radially symmetric, expressed as \phi(x) = \mu(|x|) where \mu is a positive, decreasing function on [0,1] that ensures the required properties. This rescaling ensures that \phi_\epsilon is also non-negative, smooth, compactly supported in the ball of radius \epsilon, and integrates to 1 over \mathbb{R}^n. For a locally integrable f: \mathbb{R}^n \to \mathbb{R}, the mollification f * \phi_\epsilon is defined by the (f * \phi_\epsilon)(x) = \int_{\mathbb{R}^n} f(x - y) \phi_\epsilon(y) \, dy, which produces a to f supported in a slightly enlarged version of the original support. This operation "softens" discontinuities or irregularities in f while preserving its properties, as the of \phi_\epsilon ensures that if f is integrable, then \int (f * \phi_\epsilon) = \int f. In the modern perspective from distribution theory, mollifiers serve as approximations to the identity, converging weakly to the Dirac delta distribution \delta as \epsilon \to 0^+, meaning \langle \phi_\epsilon, \psi \rangle \to \psi(0) for any test function \psi \in C_c^\infty(\mathbb{R}^n). This weak convergence underpins their role in embedding rough functions into spaces of smooth functions and in proving density results in Sobolev spaces.

Historical Development

The concept of mollifiers originated in the work of Sergei Sobolev, who introduced integral operators resembling mollifiers in his 1938 paper "Sur un théorème d'analyse fonctionnelle" proving the Sobolev embedding theorem, applying them to establish continuity properties in function spaces for applications in . These operators, though not explicitly named, served to smooth functions and approximate solutions in Sobolev spaces, laying foundational groundwork for later developments in . Independently, Kurt Otto Friedrichs developed mollifiers in 1944 as a tool for analyzing partial differential equations, particularly in demonstrating the equivalence of weak and strong solutions for elliptic systems. In his seminal paper "The identity of weak and strong extensions of differential operators," Friedrichs employed these smoothing operators with compact support to regularize distributions and prove differentiability properties, marking a pivotal advancement in the modern theory of . This work built on but did not initially reference Sobolev's earlier contributions. The term "mollifier" was coined as a playful pun by Donald Alexander Flanders, a colleague of Friedrichs at New York University, who suggested it to evoke the smoothing effect of the operator; Friedrichs approved the name, which quickly gained acceptance in the mathematical community. Over time, the terminology shifted from denoting the integral operator itself to primarily referring to its kernel function, reflecting evolving usage in approximation theory. Friedrichs later acknowledged Sobolev's prior introduction of similar mollifiers, stating that they had been developed independently by both researchers. Mollifiers found early applications in during the 1940s, where they facilitated the manipulation of generalized functions in solving boundary value problems. Their connections to Laurent Schwartz's emerging theory of distributions in the late 1940s and further propelled their adoption, as mollifiers provided a means to approximate and extend distributions while preserving key analytical properties.

Examples and Constructions

Standard Example

A standard example of a mollifier in one dimension is the given by \phi(x) = \begin{cases} C \exp\left( -\frac{1}{1 - x^2} \right) & \text{if } |x| < 1, \\ 0 & \text{otherwise}, \end{cases} where C is the normalization constant chosen such that \int_{-\infty}^{\infty} \phi(x) \, dx = 1. This function is nonnegative, smooth (C^\infty), has compact support on the interval [-1, 1], and integrates to unity over the real line. Qualitatively, \phi(x) peaks sharply at the origin and decays smoothly to zero as |x| approaches 1, providing a smooth approximation to the Dirac delta distribution when appropriately scaled. The construction extends naturally to n dimensions via the radially symmetric form \phi(x) = \begin{cases} C_n \exp\left( -\frac{1}{1 - |x|^2} \right) & \text{if } |x| < 1, \\ 0 & \text{otherwise}, \end{cases} where |x| denotes the Euclidean norm and C_n > 0 is the normalization constant ensuring \int_{\mathbb{R}^n} \phi(x) \, dx = 1. This multidimensional version retains the same properties: it is C^\infty-smooth, nonnegative, supported on the unit ball, and has unit integral. To create approximations to the at scale \varepsilon > 0, the scaled mollifier is defined as \phi_\varepsilon(x) = \varepsilon^{-n} \phi(x / \varepsilon) in \mathbb{R}^n, which preserves the unit \int_{\mathbb{R}^n} \phi_\varepsilon(x) \, dx = 1 and has compact support on the ball of radius \varepsilon. The ensures the function concentrates near the origin as \varepsilon \to 0, exemplifying the general properties of mollifiers.

General Constructions

Mollifiers can be constructed from any non-negative \psi \in C_c^\infty(\mathbb{R}^n) that integrates to a positive constant a \neq 1 over \mathbb{R}^n, by normalizing it to \phi = \frac{1}{a} \psi, ensuring the resulting \phi is non-negative, smooth, compactly supported, and integrates to 1. This normalization preserves the infinite differentiability and compact support of \psi while achieving the required unit property essential for mollifiers. Radial mollifiers, which depend only on the Euclidean norm r = |x|, are often constructed by defining a profile function \mu \in C^\infty([0,1]) with \mu(0) > 0 and \mu(1) = 0, extended to \phi(x) = \mu(|x|) for |x| \leq 1 and 0 otherwise. To normalize such radial functions, the constant C is chosen as C = \left( \int_0^1 \mu(r) \, \omega_{n-1} r^{n-1} \, dr \right)^{-1}, where \omega_{n-1} denotes the surface area of the unit in \mathbb{R}^n, ensuring \int_{\mathbb{R}^n} C \mu(|x|) \, dx = 1. This leverages spherical coordinates for the integration, guaranteeing radial and up to the boundary of the . Although non-standard due to lacking compact support, Gaussian-like mollifiers can be formed as \phi_\varepsilon(x) = (\varepsilon \sqrt{\pi})^{-n} \exp(-|x|^2 / \varepsilon^2) for \varepsilon > 0, which integrate to 1 over \mathbb{R}^n and provide analytic smoothing. These trade compact support for faster decay at infinity and stronger regularity properties, such as producing real-analytic convolutions, but may complicate applications requiring localized effects. Infinite differentiability in mollifier constructions is ensured by techniques like in the profile function, as in the Gaussian case, or by composing with smooth cutoffs that avoid derivative discontinuities at the boundary. Alternatively, partitions of can be employed to glue together smooth pieces, maintaining C^\infty regularity while controlling the .

Mathematical Properties

Smoothing Property

A fundamental property of mollifiers is their ability to smooth functions through convolution. Specifically, if f is a locally integrable function on \mathbb{R}^n, then for any standard mollifier \phi_\varepsilon with \varepsilon > 0, the convolution u_\varepsilon = f * \phi_\varepsilon is a smooth function, belonging to C^\infty(\mathbb{R}^n). This result holds because the mollifier \phi_\varepsilon itself is smooth and compactly supported, ensuring that the convolution inherits infinite differentiability regardless of the initial regularity of f. The derivatives of u_\varepsilon admit explicit expressions via convolution. For any multi-index \alpha, the partial derivative satisfies \partial^\alpha u_\varepsilon(x) = \int_{\mathbb{R}^n} f(x - y) \partial^\alpha \phi_\varepsilon(y) \, dy = (f * \partial^\alpha \phi_\varepsilon)(x). If f possesses higher regularity, such as being k-times differentiable with \partial^\alpha f locally integrable for |\alpha| \leq k, integration by parts yields the alternative form \partial^\alpha u_\varepsilon(x) = (-1)^{|\alpha|} \int_{\mathbb{R}^n} (\partial^\alpha f)(x - y) \phi_\varepsilon(y) \, dy = (-1)^{|\alpha|} ((\partial^\alpha f) * \phi_\varepsilon)(x). These formulas demonstrate how the smoothing process transfers derivatives from the mollifier to the convolved function or vice versa. The proof of smoothness relies on the ability to differentiate under the integral sign, justified by the compact support of \phi_\varepsilon and the local integrability of f. To verify that \partial^\alpha u_\varepsilon exists and equals the convolution above, consider the difference quotient for the derivative; the dominated convergence theorem applies since |\partial^\alpha \phi_\varepsilon(y)| is bounded and supported in a fixed ball scaled by \varepsilon, allowing passage of the limit inside the integral. Iterating this process for all orders of differentiation establishes the C^\infty regularity. Moreover, u_\varepsilon is locally uniformly continuous and locally bounded. The boundedness follows from the estimate |u_\varepsilon(x)| \leq \int_{B(0,\varepsilon)} |f(x - y)| \cdot \varepsilon^{-n} |\phi(y/\varepsilon)| \, dy, which, upon substitution z = y/\varepsilon, becomes |u_\varepsilon(x)| \leq \int_{B(0,1)} |f(x - \varepsilon z)| |\phi(z)| \, dz. Local integrability of f ensures this integral is finite and bounded on compact sets, as the average of |f| over balls of radius \varepsilon remains controlled. Uniform continuity on compact domains arises similarly, by controlling the variation of the integral over small shifts in x. A concrete illustration of this smoothing occurs when convolving the H, which is discontinuous at the , with a standard mollifier such as \phi_\varepsilon(x) = \varepsilon^{-n} \phi(x/\varepsilon) where \phi is a nonnegative smooth with 1 and support in the unit ball. The result H * \phi_\varepsilon is a C^\infty function that transitions smoothly from 0 to 1 over an interval of width proportional to \varepsilon, effectively regularizing the jump discontinuity.

Approximation to the Identity

A fundamental role of mollifiers lies in their capacity to approximate functions via convolution in the limit as the scaling parameter \varepsilon approaches zero. Specifically, consider a continuous function f: \mathbb{R}^n \to \mathbb{R} with compact support. Let \phi_\varepsilon(y) = \varepsilon^{-n} \phi(y/\varepsilon), where \phi is a standard mollifier (nonnegative, smooth, compactly supported, and integrating to 1). Then, the convolution f * \phi_\varepsilon converges uniformly to f, satisfying \|f * \phi_\varepsilon - f\|_\infty \to 0 as \varepsilon \to 0. This uniform convergence extends more broadly to L^p spaces. For f \in L^p(\mathbb{R}^n) with $1 \leq p < \infty, the approximation holds in the L^p norm: \|f * \phi_\varepsilon - f\|_p \to 0 as \varepsilon \to 0. The case p = \infty requires additional assumptions like uniform continuity for global convergence, but local uniform convergence occurs on compact sets for bounded continuous functions. The proof for continuous functions with compact support leverages uniform continuity. Fix x \in \mathbb{R}^n; then, |(f * \phi_\varepsilon)(x) - f(x)| = \left| \int_{\mathbb{R}^n} (f(x - y) - f(x)) \phi_\varepsilon(y) \, dy \right| \leq \sup_{|y| < \varepsilon} |f(x - y) - f(x)| \cdot \int_{\mathbb{R}^n} \phi_\varepsilon(y) \, dy = \sup_{|y| < \varepsilon} |f(x - y) - f(x)|, since the integral of \phi_\varepsilon is 1 and \operatorname{supp} \phi_\varepsilon \subset B(0, \varepsilon) for small \varepsilon. Uniform continuity of f on its compact support implies the supremum tends to 0 as \varepsilon \to 0, uniformly in x. For L^p convergence, one applies density arguments: approximate f by continuous compactly supported functions, use Young's inequality to bound the operator norm of convolution with \phi_\varepsilon, and pass to the limit via dominated convergence. In the distributional sense, mollifiers provide a sequence approximating the Dirac . The family \{\phi_\varepsilon\}_{\varepsilon > 0} converges weakly to \delta in the space of distributions: for any test \psi \in C_c^\infty(\mathbb{R}^n), \langle \phi_\varepsilon, \psi \rangle = \int_{\mathbb{R}^n} \phi_\varepsilon(y) \psi(y) \, dy \to \psi(0) = \langle \delta, \psi \rangle as \varepsilon \to 0, by a and dominated convergence, since \phi integrates to 1 and is nonnegative. For a T, this yields T * \phi_\varepsilon \to T weakly. In particular, for a locally integrable f inducing the regular T_f(\psi) = \int f \psi, one has \langle f * \phi_\varepsilon, \psi \rangle \to \langle f, \psi \rangle for all test functions \psi. The refines these limits under higher regularity. If f \in C^k(\mathbb{R}^n) for k \geq 1, then \|f * \phi_\varepsilon - f\|_\infty = O(\varepsilon^k) as \varepsilon \to 0, with explicit constants depending on the C^k norm of f and the moments of \phi up to order k. This estimate arises from : expand f(x - y) to order k-1 around x, integrate the polynomial terms (which vanish if \phi has vanishing moments up to k-1, or otherwise contribute lower-order terms), and bound the remainder by the k-th times \varepsilon^k integrated against \phi_\varepsilon.

Support of the Convolution

When a f is convolved with a mollifier \phi_\epsilon, the of the resulting f * \phi_\epsilon is contained within the Minkowski sum of the of f and the of \phi_\epsilon. Specifically, if \phi is a mollifier with \operatorname{supp}(\phi) = B(0,1), then \phi_\epsilon(x) = \epsilon^{-n} \phi(x/\epsilon) has in the B(0,\epsilon), yielding the : \operatorname{supp}(f * \phi_\epsilon) \subseteq \operatorname{supp}(f) + B(0, \epsilon). Equality holds in this inclusion when \operatorname{supp}(\phi) = B(0,1), as the precisely fills the \epsilon-neighborhood of \operatorname{supp}(f). To prove this, suppose x \notin \operatorname{supp}(f) + B(0, \epsilon). Then, for all y \in B(0, \epsilon), x - y \notin \operatorname{supp}(f), so f(x - y) = 0. It follows that (f * \phi_\epsilon)(x) = \int_{\mathbb{R}^n} f(x - y) \phi_\epsilon(y) \, dy = 0, since the integrand vanishes everywhere, confirming x \notin \operatorname{supp}(f * \phi_\epsilon). This geometric containment implies that mollification enlarges the support by at most the radius \epsilon, preserving the essential location of non-zero values while smoothing them. For distributions u \in \mathcal{D}'(\mathbb{R}^n), the convolution u * \phi_\epsilon (well-defined since \phi_\epsilon \in \mathcal{D}) satisfies \operatorname{supp}(u * \phi_\epsilon) \subseteq \operatorname{supp}(u) + B(0, \epsilon), mirroring the function case. Similarly, the singular support obeys \operatorname{sing\, supp}(u * \phi_\epsilon) \subseteq \operatorname{sing\, supp}(u) + B(0, \epsilon), ensuring mollification does not introduce singularities outside this \epsilon-neighborhood. If u has compact support, then u * \phi_\epsilon also has compact support, as the Minkowski sum of two compact sets is compact.

Applications

In the Theory of Distributions

Mollifiers play a central role in the regularization of distributions, allowing any distribution T \in \mathcal{D}'(\mathbb{R}^n) to be approximated by smooth functions. Specifically, the convolution T * \rho_\epsilon, where \rho_\epsilon is a standard mollifier with support shrinking to the origin as \epsilon \to 0^+, yields a C^\infty function that converges to T in the distributional sense: \langle T * \rho_\epsilon - T, \phi \rangle \to 0 for every test function \phi \in \mathcal{D}(\mathbb{R}^n). This process, known as regularization, transforms singular objects like the Dirac delta into sequences of smooth approximations, preserving the underlying distributional structure while enabling pointwise evaluation and differentiation. In the context of defining products involving distributions, mollifiers facilitate the multiplication of a general distribution u by a smooth function v \in C^\infty(\mathbb{R}^n). The product u v is defined distributionally by \langle u v, \phi \rangle = \langle u, v \phi \rangle for \phi \in \mathcal{D}(\mathbb{R}^n), leveraging the smoothness of v to ensure v \phi remains a valid test function. When v has compact support, an alternative construction uses mollifiers: approximate u by u_\epsilon = u * \rho_\epsilon, which is smooth, form the pointwise product u_\epsilon v (also smooth and compactly supported), and take the limit \lim_{\epsilon \to 0} u_\epsilon v = u v in the distributional topology, justified by the continuity of multiplication by fixed smooth functions. This approach aligns with the identification u v = u * (v \delta), where \delta is the Dirac distribution, since convolution with the compactly supported distribution v \delta is well-defined. Mollifiers extend the notion of multiplication to pairs of distributions u, v \in \mathcal{D}'(\mathbb{R}^n) under suitable conditions, such as when the singular support of one avoids the support of the other. If \operatorname{supp}(u) \cap \operatorname{sing}\operatorname{supp}(v) = \emptyset, local smoothing via mollifiers allows defining u v by regularizing u near the singularities of v (or vice versa), ensuring the product converges distributionally without ambiguity. This local regularization exploits the approximation to the identity property of mollifiers to handle interactions away from singular sets, enabling tensor product constructions and other algebraic operations in distribution theory. The integration of mollifiers into distribution theory was pivotal in Laurent Schwartz's foundational work during the late 1940s, where they provided tools for rigorous approximation and operational extensions in his development of the framework, culminating in his seminal treatise that formalized these concepts.

In Partial Differential Equations

Mollifiers play a crucial role in establishing the equivalence between weak and strong solutions for partial differential equations (PDEs), particularly through "weak = strong" theorems. In these theorems, a weak solution u to a PDE is mollified to produce a smooth approximation u_\epsilon = u * \phi_\epsilon, where \phi_\epsilon is a standard mollifier with parameter \epsilon > 0. This mollified function u_\epsilon satisfies the PDE in the classical (strong) sense due to the smoothing properties of convolution, and as \epsilon \to 0, u_\epsilon converges to u in appropriate norms, leveraging the approximation-to-the-identity property of mollifiers to infer that the original weak solution possesses the necessary regularity to satisfy the strong formulation. In elliptic PDEs, mollifiers were originally employed by Friedrichs to prove interior regularity for solutions in Sobolev spaces. For a weak solution u \in W^{k,p}(\Omega) to an elliptic equation like -\Delta u = f in an open set \Omega, mollification yields u_\epsilon \in C^\infty(\Omega) that approximately satisfies the equation, and estimates on the difference u - u_\epsilon allow passage to the limit to show higher-order regularity, such as u \in W^{k+2,p}_{\text{loc}}(\Omega) under suitable assumptions on the coefficients and right-hand side. This technique extends to more general linear and quasilinear elliptic systems, providing local C^\infty-regularity for solutions in appropriate spaces. Bootstrap arguments further exploit mollification to iteratively improve regularity. Starting from a weak solution in, say, L^2(\Omega), one mollifies to obtain a smooth approximant satisfying the PDE, applies elliptic estimates to gain one or two derivatives (e.g., into H^2), and repeats the process; each iteration uses the previous regularity to control error terms in the mollified equation, eventually bootstrapping to local C^\infty-smoothness. This method is standard in proving higher regularity for elliptic problems, where the gain in derivatives per step depends on the order of the operator and embedding theorems. A specific example arises in the Laplace equation \Delta u = 0, where weak (distributional) solutions are harmonic functions. Mollifying a weak solution u produces u_\epsilon = u * \phi_\epsilon, which approximately preserves harmonicity since \Delta u_\epsilon = (\Delta u) * \phi_\epsilon = 0, and the mean-value property holds exactly for such convolutions over balls; passing to the limit \epsilon \to 0 shows that u is smooth and satisfies the classical Laplace equation locally. This illustrates how mollification bridges weak and strong notions, leading to the conclusion that all weak solutions to Laplace's equation are C^\infty.

As Smooth Cutoff Functions

Mollifiers provide a means to construct smooth approximations to the of a bounded , yielding cutoff functions that transition gradually between 1 inside the and 0 outside. Consider an \Omega \subset \mathbb{R}^n and a compact \overline{\Omega'} \subset K \subset \Omega for some compact K. Let \eta be a standard mollifier—a nonnegative smooth function with compact support in the unit ball and integral 1—and define the scaled version \eta_\epsilon(x) = \epsilon^{-n} \eta(x/\epsilon) for \epsilon > 0. The convolution \psi_\epsilon = \eta_\epsilon * \chi_K, where \chi_K is the of K, is then a smooth function satisfying $0 \leq \psi_\epsilon \leq 1, \psi_\epsilon \equiv 1 on \Omega', and \psi_\epsilon \equiv 0 outside the \epsilon-neighborhood of K. As \epsilon \to 0, \psi_\epsilon converges to \chi_K pointwise almost everywhere and in L^p norms for $1 \leq p < \infty. This construction allows mollifiers to smoothly cut off functions supported in \Omega while preserving smoothness. For a function f \in C^\infty(\Omega) supported in \Omega, the product f \psi_\epsilon extends f to a smooth function on \mathbb{R}^n with compact support in the \epsilon-enlargement of \Omega, without introducing discontinuities at the boundary. Such cutoffs maintain the original smoothness class inside \Omega' and enable controlled decay of derivatives near the transition region, with estimates like |\nabla \psi_\epsilon| \leq C/\epsilon for some constant C independent of the domain geometry. This is particularly useful for functions in Sobolev spaces, where the cutoff ensures the product remains in the same space locally. In proofs of variational problems and integration by parts formulas, these smooth cutoffs approximate rough domains by replacing characteristic functions with \psi_\epsilon, facilitating passage to smooth subdomains \Omega_\epsilon = \{x \in \Omega : \dist(x, \partial \Omega) > \epsilon\} where direct mollification applies without boundary issues. For instance, in elliptic PDE theory, multiplying test functions by \psi_\epsilon localizes integrals over irregular boundaries, allowing arguments to extend results from smooth to general domains. Fixed-\epsilon versions of these convolutions, such as \eta_\epsilon * \chi_{B_r(0)} for a B_r(0), serve as prototypes for bump functions—smooth, compactly supported functions that form the test C_c^\infty(\mathbb{R}^n) in distribution . These bump functions are dense in various spaces and for defining weak derivatives and generalized solutions.

References

  1. [1]
    [PDF] Introduction to Sobolev Spaces
    Dec 23, 2018 · Definition 3.2.2 (Mollifier). A mollifier on Rn is a smooth ... Mathematics. European Mathematical Society (EMS), Zürich, 2016. [Ste70] ...
  2. [2]
    [PDF] NOTES ON Lp AND SOBOLEV SPACES - UC Davis Math
    where η is the standard mollifier defined in Definition 1.25. Then. (A) u ... Department of Mathematics, University of California, Davis, CA 95616, USA.
  3. [3]
    [PDF] Sobolev spaces and embedding theorems - Icmc-Usp
    The theory of Sobolev spaces has been originated by Russian mathematician S.L. Sobolev around 1938. [SO]. These spaces were not introduced for some theoretical ...
  4. [4]
    [PDF] MA612L-Partial Differential Equations - IIT Tirupati
    Oct 1, 2025 · Department of Mathematics and Statistics. IIT Tirupati, Tirupati ... Definition 2 (Mollifier). Define η ∈ C∞(Rn) by η(x) := (. Cexp. 1. |x ...
  5. [5]
    [PDF] A mollifier approach to the deconvolution of probability densities
    Dec 4, 2018 · Its definition relies not only on the regularization parameter a, but also on the shape ... SIAM Journal on Applied Mathematics, 56(5):1424–1444, ...
  6. [6]
    [PDF] Mollifiction (rough draft) - John McCuan
    Apr 14, 2020 · Mδ[φ] = ∫ µδφ. As distributions lim δց0. Mδ = δ0 where δ0 is the Dirac delta distribution (or evaluation functional) given by δ0[φ] = φ(0) ...
  7. [7]
    [PDF] arXiv:hep-lat/0602013v2 22 Sep 2008
    Sep 22, 2008 · (Standard mollifier.) A mollifier, η, also called an approximate identity, is a positive C∞. (Rn) function. The standard mollifier is defined in ...
  8. [8]
    [PDF] arXiv:1703.06299v1 [math.FA] 18 Mar 2017
    Mar 18, 2017 · His colleague, Donald Alexander. Flanders, suggested the name mollifiers. Friedrichs himself acknowledged Sobolev's work on mollifiers stating ...
  9. [9]
    [PDF] Test functions, mollifiers and convolution - Timo Rohner
    Definition 4. A mollifier is a smooth function ϕ : Rn → R, i.e. ϕ ∈ C∞(Rn), if the following conditions hold.
  10. [10]
    [PDF] Compact Polynomial Mollifiers For Poisson's Equation
    Abstract. In this paper we describe a family of polynomial mollifiers of compact support with a parameterized degree of differentiability.<|control11|><|separator|>
  11. [11]
    [PDF] A Ordinary Differential Equations - SLMath
    the Gaussian: then b is also a Gaussian, and moreover the mollifier Jε yields real-analytic functions (which is not the case if had compact support). Remark ...
  12. [12]
    [PDF] Real Analysis
    3.5 Smoothing in the entire space . ... harmonic analysis, functional analysis and partial differential equations.
  13. [13]
    [PDF] Partial Differential Equations (based on L.C. Evans's textbook) by ...
    One-dimensional wave equations and d'Alembert's formula. 81. §4.3. Higher ... standard radial mollifier. For each sufficiently small > 0, let. Ω = {x ∈ Ω ...
  14. [14]
    [PDF] expository notes on distribution theory - UMD MATH
    It is a finite linear combination of the delta function and its derivatives. An additional homogeneity argument shows ∆E = cδ.
  15. [15]
    None
    ### Summary of Approximation by Convolution with Mollifiers
  16. [16]
    Rate of convergence of mollifiers // Sobolev norms - MathOverflow
    Jul 9, 2019 · Rates of convergence of mollifiers with Sobolev norms on manifold · 3 · Friedrichs mollifiers and Sobolev spaces · 2 · Norms in Sobolev space W1 ...Missing: normalization constant 1.136
  17. [17]
    [PDF] Theory of function and distribution spaces
    Apr 27, 2022 · ... Theorem 7.10, which states that the support of the convolution of two functions is included in the closure of the sum of the supports ...
  18. [18]
    [PDF] Lecture Notes Advanced Analysis - Poisson
    May 11, 2018 · DISTRIBUTION THEORY. 62. 4.6 Regularization by Mollification. The goal ... Rudin, Functional Analysis, International Series in Pure and Applied ...
  19. [19]
    None
    ### Summary of Products of Distributions with Smooth Functions, Regularization Using Mollifiers, and Approximation in Distribution Theory
  20. [20]
    245C, Notes 3: Distributions
    ### Summary of Mollifiers, Regularization, and Products with Smooth Functions from https://terrytao.wordpress.com/2009/04/19/245c-notes-3-distributions/
  21. [21]
    The Identity of Weak and Strong Extensions of Differential Operators
    operator must be extended. Two such extensions offer themselves, a "weak" and a "strong" one. Existence theorems, when derived by variational meth-.
  22. [22]
    [PDF] Chapter 2: Laplace's equation - UC Davis Math
    For example, harmonic functions are smooth because local averages over a ball vary smoothly as the ball moves. We will prove this result by mollification, which ...