Fact-checked by Grok 2 weeks ago

Rademacher's theorem

Rademacher's theorem states that if U is an open of \mathbb{R}^m and f: U \to \mathbb{R}^n is a , then f is differentiable with respect to . This means that at almost every point x \in U, the \lim_{h \to 0} \frac{f(x + h) - f(x) - Df_x(h)}{\|h\|} = 0 exists, where Df_x: \mathbb{R}^m \to \mathbb{R}^n is a bounded representing the . Named after the German mathematician Hans Rademacher, the theorem was originally proved in 1919 as part of a broader study on partial and total differentiability of functions of several variables. Rademacher's work built on earlier results in , such as the one-dimensional case for monotone functions being differentiable , and extended differentiability guarantees to higher dimensions under the milder condition rather than stricter smoothness assumptions. The theorem is a cornerstone of and , providing essential structure for functions that arise in optimization, partial differential equations, and variational problems. It implies that the Df_x, when it exists, satisfies \|Df_x\| \leq \Lip(f), the constant of f, ensuring controlled growth. In the context of Sobolev spaces, the theorem connects to the fact that functions are in W^{1,\infty}, with weak bounded by the constant . Beyond spaces, Rademacher's theorem has been generalized to measure spaces satisfying conditions like doubling measures and Poincaré inequalities, where functions are asymptotically linear . These extensions, developed in works like those of Cheeger in , apply to spaces such as Heisenberg groups and Carnot-Carathéodory spaces, facilitating analysis in non-smooth geometries. The theorem's proofs typically rely on tools like the Vitali covering lemma and , highlighting its deep ties to measure theory.

Background Concepts

Lipschitz Continuity

In , a f: U \to \mathbb{R}^m, where U \subset \mathbb{R}^n is an , is said to be K- continuous (or simply K-Lipschitz) if there exists a constant K \geq 0 such that \| f(x) - f(y) \| \leq K \| x - y \| for all x, y \in U, with norms denoting the Euclidean norms in the respective spaces. This condition, named after the German mathematician , imposes a uniform bound on the rate at which the can change, ensuring that the distance between function values is controlled by the distance between inputs scaled by K, the Lipschitz constant. Lipschitz continuity exhibits several key properties that distinguish it from weaker forms of continuity. It implies on U, as the inequality guarantees that for any \epsilon > 0, choosing \delta = \epsilon / K suffices to ensure \| f(x) - f(y) \| < \epsilon whenever \| x - y \| < \delta. In the one-dimensional case, where U \subset \mathbb{R} and f: U \to \mathbb{R}, Lipschitz continuity further implies absolute continuity, meaning f can be expressed as the integral of a locally integrable function. Moreover, if f is differentiable at every point in U, then f is K-Lipschitz if and only if the supremum of the norms of its derivatives is bounded by K, i.e., \sup_{x \in U} \| Df(x) \| \leq K, establishing an equivalence between the global Lipschitz condition and local boundedness of the derivative. Common examples illustrate the breadth of Lipschitz functions. The distance function d(x) = \| x - x_0 \| to a fixed point x_0 \in \mathbb{R}^n is 1-Lipschitz, as \| d(x) - d(y) \| \leq \| x - y \| by the triangle inequality. Orthogonal projections onto closed subspaces of \mathbb{R}^n are also 1-Lipschitz, preserving distances in a non-expansive manner. Simple linear maps f(x) = Ax + b, where A is an m \times n matrix with operator norm \| A \| \leq K, are K-Lipschitz, providing a foundational class of such functions. In real analysis, Lipschitz continuity plays a crucial role by controlling the growth rates of functions; for instance, on a bounded domain, it ensures linear boundedness, as \| f(x) \| \leq \| f(0) \| + K \| x \|. This property facilitates measure-theoretic arguments, such as the preservation of sets of measure zero under Lipschitz mappings, which is essential for studying function behavior in higher dimensions.

Differentiability Almost Everywhere

A function f: \mathbb{R}^n \to \mathbb{R}^m is differentiable at a point x \in \mathbb{R}^n if there exists a linear transformation L: \mathbb{R}^n \to \mathbb{R}^m such that \lim_{h \to 0} \frac{\|f(x + h) - f(x) - L(h)\|}{\|h\|} = 0. This linear map L is called the derivative of f at x, often denoted Df(x) or \nabla f(x). In the context of Lebesgue measure theory, a property is said to hold almost everywhere (a.e.) if the set where it fails has Lebesgue measure zero. The Lebesgue measure on \mathbb{R}^n, denoted m_n or simply m, extends the notion of length, area, and volume to more general subsets, assigning to each open set the sum of volumes of countably many cubes covering it, and completing this to a measure on the sigma-algebra of Lebesgue measurable sets. Sets of measure zero, such as countable sets or the Cantor set, are "negligible" in this sense, allowing analysis to focus on the "typical" behavior of functions over \mathbb{R}^n. For example, continuous functions on \mathbb{R} are differentiable almost everywhere if they are of bounded variation, but counterexamples exist showing that continuity alone does not suffice; the Weierstrass function w(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x) with $0 < a < 1 and ab > 1 + \frac{3\pi}{2} is continuous everywhere yet differentiable nowhere. This pathological example, constructed by in 1872, highlights how oscillations can prevent differentiability at every point despite uniform continuity on compact intervals. Measure-theoretic tools like the Vitali covering lemma facilitate proofs of differentiability almost everywhere by selecting disjoint subcollections from coverings by balls. Specifically, if a set E \subset \mathbb{R}^n of finite is covered in the Vitali sense by a family of balls (meaning for every x \in E and \delta > 0, there is a ball centered at x with radius less than \delta in the family), then there exists a disjoint finite or countable subcollection whose union has measure at least a positive (such as $1/5^n) of m^*(E), the of E. This lemma, without its full proof here, underpins density arguments essential for theorems.

Statement of the Theorem

Formal Statement

Rademacher's theorem asserts that if U \subset \mathbb{R}^m is an open set and f: U \to \mathbb{R}^n is a Lipschitz continuous function, then f is differentiable in the Fréchet sense at Lebesgue almost every point x \in U. At each such point of differentiability, the Fréchet derivative Df(x): \mathbb{R}^m \to \mathbb{R}^n exists and is a bounded linear transformation. This result applies to vector-valued functions, where the theorem guarantees differentiability with respect to , meaning the set of points in U where f fails to be differentiable has Lebesgue measure zero. The Lipschitz condition, which requires that there exists a constant K > 0 such that \|f(y) - f(x)\| \leq K \|y - x\| for all x, y \in U, imposes a controlled growth on f that precludes pathological non-differentiability on sets of positive measure.

Historical Development

Rademacher's theorem is named after the German mathematician Hans Rademacher, who established the result in his 1919 paper published in Mathematische Annalen, where he demonstrated the almost everywhere differentiability of continuous functions of several variables. This work formed part of his thesis under at the and addressed key issues in , including partial and total differentiability as well as transformations of double integrals. Rademacher's proof extended earlier one-dimensional results to higher dimensions, marking a significant advancement in understanding the structure of mappings. The theorem built directly on foundational developments in measure theory from the early 1900s, particularly Henri Lebesgue's differentiation theorem, which showed that monotone functions on the real line are differentiable almost everywhere (originally in Lebesgue's 1902 thesis and elaborated in subsequent works around 1904–1910). Essential tools like Giuseppe Vitali's covering lemma, introduced in 1908, provided the covering arguments necessary for handling the almost everywhere properties in proofs of differentiability. These precursors were motivated by broader efforts to rigorize integration and differentiation in the context of variational problems, where controlling the regularity of functions was crucial for minimizing functionals in the calculus of variations. In the 1930s, Charles B. Morrey Jr. offered an influential alternative proof of the theorem, employing integral identities derived from Fubini's theorem to establish the existence of partial derivatives almost everywhere, followed by verification of the full differentiability condition. Morrey's approach, detailed in his early papers on multiple integrals and quasi-linear equations (e.g., 1938), emphasized applications to elliptic partial differential equations and variational methods, highlighting the theorem's utility in higher-dimensional settings. This extension solidified the result's role in analysis while addressing subtleties in the multidimensional case that Rademacher's original proof had outlined. Rademacher himself transitioned from real analysis to analytic number theory later in his career, contributing seminal results like the exact formula for the partition function in 1937.

Proof Overview

One-Dimensional Case

In the one-dimensional case, the proof of Rademacher's theorem relies on classical real analysis tools, establishing that a Lipschitz continuous function f: [a, b] \to \mathbb{R} is differentiable almost everywhere with respect to Lebesgue measure. This serves as the foundational instance of the general theorem, where the domain is \mathbb{R} instead of \mathbb{R}^n. The proof begins with a reduction showing that Lipschitz continuity implies absolute continuity on the interval [a, b]. Specifically, if f satisfies |f(x) - f(y)| \leq K |x - y| for some constant K > 0 and all x, y \in [a, b], then for any \epsilon > 0, choose \delta = \epsilon / K. For any finite collection of disjoint subintervals [c_i, d_i] \subset [a, b] with \sum (d_i - c_i) < \delta, the sum \sum |f(d_i) - f(c_i)| \leq K \sum (d_i - c_i) < \epsilon, satisfying the definition of absolute continuity. Absolute continuity further implies that f has bounded variation, as the total variation over [a, b] is controlled by the Lipschitz constant. Functions of bounded variation can be decomposed as the difference of two non-decreasing functions, say f = g - h where g and h are non-decreasing. Each non-decreasing function is differentiable almost everywhere, a result proved using the (also known as ). This lemma states that for a non-decreasing function F: [a, b] \to \mathbb{R}, the set where the upper and lower derivatives differ has measure zero; it constructs a countable collection of disjoint intervals covering points of non-differentiability, whose total length is at most F(b) - F(a). Thus, f' exists almost everywhere as the difference of the derivatives of g and h. Moreover, the derivative satisfies the integral representation f(x) = f(a) + \int_a^x f'(t) \, dt for all x \in [a, b], where f' \in L^1([a, b]). This follows from the fundamental theorem of calculus for absolutely continuous functions, as the derivative f' is the Radon-Nikodym derivative of the associated measure. Specifically, the limit f'(x) = \lim_{h \to 0} \frac{f(x + h) - f(x)}{h} exists almost everywhere. A canonical example is the absolute value function f(x) = |x| on \mathbb{R}, which is Lipschitz continuous with constant 1. It is differentiable everywhere except at the origin, a set of Lebesgue measure zero, where f'(x) = \operatorname{sign}(x) for x \neq 0. The integral representation holds, as |x| = \int_0^x \operatorname{sign}(t) \, dt.

Multidimensional Extension

The multidimensional extension of Rademacher's theorem addresses Lipschitz continuous functions f: \mathbb{R}^n \to \mathbb{R}^m for n > 1, establishing that such functions are differentiable almost everywhere with respect to Lebesgue measure. The proof strategy relies on measure-theoretic tools to extend the one-dimensional case, where differentiability follows from bounded variation, to higher dimensions by reducing the problem to slices and ensuring the derivative forms a linear map. Specifically, the approach begins by applying Fubini's theorem to show the existence of partial derivatives almost everywhere, then verifies that these partials assemble into a bounded linear operator approximating the function. To establish partial derivatives, consider the function restricted to lines parallel to the coordinate . By Fubini's theorem, for almost every fixed point in the orthogonal , the one-dimensional restriction of f along the is and thus differentiable on that line, yielding the \partial f / \partial x_j in \mathbb{R}^n. This holds for each coordinate direction j = 1, \dots, n, with the partials satisfying the condition inherited from f. The set where any partial fails to exist has zero, as the exceptional sets for each direction form a countable of measure-zero sets. Next, directional derivatives are introduced to connect the partials into a full . For a v \in \mathbb{R}^n, the is defined as D_v f(x) = \lim_{t \to 0} \frac{f(x + t v) - f(x)}{t}, which exists by a similar slicing argument via Fubini, treating the restriction along the direction v as a one-dimensional function. The partial derivatives then relate to these via the chain rule in , with D_v f(x) = \sum_{j=1}^n v_j \partial f / \partial x_j (x) holding . To prove of the candidate Df(x), is applied: for vectors h, k \in \mathbb{R}^n, \langle Df(x) h, Df(x) k \rangle = \frac{1}{4} \left( \|D_{h+k} f(x)\|^2 - \|D_{h-k} f(x)\|^2 \right), ensuring Df(x) acts as a from the partials , with boundedness following from the constant of f. The core of the proof uses Vitali covering lemma to establish approximate differentiability, selecting a dense countable set of directions on the unit sphere to cover potential non-differentiability points with disjoint balls where the difference quotient converges uniformly. A key integral identity confirms the linear approximation: for fixed h \in \mathbb{R}^n, \int_{\mathbb{R}^n} \left\| Df(x) h - \big( f(x + h) - f(x) \big) \right\| \, dx \to 0 \quad \text{as} \quad |h| \to 0, proved by splitting the integral over regions where approximate differentiability holds and applying the dominated convergence theorem, leveraging the Lipschitz bound to dominate the integrand by an integrable function. This identity extends to all directions by density, implying full differentiability almost everywhere. Challenges arise in handling non-differentiability on measure-zero sets, such as Cantor-like sets of positive but zero , where slicing may encounter pathological behaviors. The Vitali covering and Fubini arguments mitigate this by ensuring the exceptional set—where partials exist but fail to linearize or approximate—is contained in a G_\delta set of measure zero, often refined using the of rational directions to avoid such pathologies.

Applications

In Sobolev Spaces

functions defined on an open subset \Omega \subset \mathbb{R}^n belong to the W^{1,p}(\Omega) for every $1 \leq p \leq \infty, as their weak partial derivatives exist and satisfy \|\nabla u\|_{L^p(\Omega)} \leq K |\Omega|^{1/p} , where K is the constant. Rademacher's theorem ensures that such functions are classically differentiable in \Omega, and the classical derivative coincides with the . A fundamental consequence is the preservation of Sobolev regularity under bi-Lipschitz transformations. If \psi: V \to U is a bi-Lipschitz between open sets in \mathbb{R}^n, then for any u \in W^{1,p}(U) with $1 \leq p < \infty, the composition u \circ \psi lies in W^{1,p}(V), and its weak gradient is given by the chain rule formula \nabla (u \circ \psi)(y) = D u(\psi(y)) \cdot D \psi(y) \quad \text{a.e. in } V, where the derivatives exist almost everywhere by Rademacher's theorem applied to \psi and u. This property is essential for change-of-variables formulas in multiple integrals, enabling the transport of Sobolev norms and facilitating proofs of invariance under domain transformations. In Sobolev embedding theorems, Rademacher's theorem plays a key role when p > n. Here, W^{1,p}(\Omega) embeds continuously into the Hölder space C^{0,1 - n/p}(\overline{\Omega}), and in the boundary case approaching (as p \to \infty), the almost everywhere differentiability follows directly. Lipschitz functions in these spaces can also be approximated in the W^{1,p}-norm by smooth C^\infty functions through mollification, with Rademacher's theorem verifying that the classical derivatives of the limit recover the weak ones . These tools underpin regularity results in Calderón-Zygmund theory for elliptic partial differential equations, where weak solutions in W^{1,p} with p > n exhibit classical differentiability properties .

In Geometric Measure Theory

In , Rademacher's theorem plays a foundational role by guaranteeing the existence of approximate tangent planes to the images of maps almost everywhere, which allows for the precise of such images as generalized manifolds. Specifically, for a f: \mathbb{R}^n \to \mathbb{R}^m, the theorem ensures that the approximate T_x f exists at \mathcal{L}^n-almost every point x, providing a that aligns with the \nabla f(x). This structure is essential for defining orientations and vectors almost everywhere on the image, where the space is the of the , facilitating the study of multiplicity and density in measure-theoretic contexts. The underpins key integral formulas in , notably the area and coarea formulas, which relate the of images to integrals of the determinant. For instance, the area formula states that for a map f, the measure \mathcal{H}^m(f(E)) can be expressed as \int_E J_f(x) \, d\mathcal{H}^n(x) for measurable E \subset \mathbb{R}^n, relying directly on the differentiability to define the J_f(x) = \sqrt{\det((\nabla f(x))^T \nabla f(x))}. Similarly, the coarea formula decomposes integrals over preimages using level sets, again hinging on the tangent structure provided by Rademacher's . These formulas are crucial for computing masses of currents and varifolds. Additionally, the supports proofs of differentiability for the characteristic functions of certain sets, such as those of finite perimeter, by approximating them with functions whose derivatives yield reduced boundary measures. Representative examples illustrate these applications vividly. Lipschitz graphs, defined as sets of the form \{(x, f(x)) : x \in \mathbb{R}^n\} for f: \mathbb{R}^n \to \mathbb{R}, serve as prototypical rectifiable currents, where Rademacher's theorem ensures they admit an approximate plane and normal vector , allowing representation as oriented currents with multiplicity. In the study of varifolds, blow-up limits—rescalings of varifolds at singular points—often converge to varifolds that are flat or conical, with the theorem providing the differentiable structure needed to analyze stability and regularity; for example, varifolds with positive inherit cones that are graphs or planes . These concepts tie into Federer's structure theorem, which characterizes rectifiable sets via countable unions of images and their approximations, resolving the of measures into rectifiable and unrectifiable parts.

Generalizations

To Sobolev Functions

extended Rademacher's theorem to functions of in the Sobolev sense, establishing that every function u \in W^{1,p}(\Omega), where \Omega \subset \mathbb{R}^n is open and p > n, is classically differentiable in \Omega. This result replaces the global Lipschitz condition with the weaker requirement that the weak \nabla u belongs to L^p(\Omega)^n, leveraging the higher integrability exponent p > n to control local oscillations via singular integral estimates and maximal function techniques. Furthermore, under this embedding, the classical derivative at points of differentiability coincides with the weak derivative almost everywhere, ensuring that the pointwise gradient \nabla u(x) equals the precise representative of the distributional gradient. Specifically, for almost every x \in \Omega, \lim_{h \to 0} \frac{u(x + h) - u(x) - \nabla u(x) \cdot h}{|h|} = 0, where \nabla u(x) is the weak gradient, and this limit exists in the classical sense. The conclusion fails for p \leq n, as there exist functions in W^{1,p}(\Omega) that are nowhere differentiable or fail to be differentiable on sets of positive measure; such counterexamples can be constructed using modifications of Weierstrass-type nowhere differentiable functions adapted to have weak gradients in L^p. For instance, in dimensions n \geq 2, explicit examples in W^{1,n}(\mathbb{R}^n) exhibit non-differentiability on dense open sets, highlighting the sharpness of the exponent p > n. Post-1960s advances refined these results by incorporating and other Orlicz-type conditions to recover differentiability for lower exponents. In particular, if the weak gradient lies in the Lorentz space L^{n,1}(\Omega), then differentiability holds even for the borderline case p=1, extending the theorem beyond the critical integrability threshold. These developments, building on Calderón-Zygmund theory, have informed applications in nonlinear elliptic PDEs and compensated compactness.

To Metric Spaces

The generalization of Rademacher's theorem to spaces replaces the classical linear derivative with a notion of differentiability, where the at a point x is an approximate (or ) between the tangent cones at x and f(x). Specifically, for a map f: (X, d_X) \to (Y, d_Y) between spaces, f is metrically differentiable at x if there exists a map \delta f(x): (X, x) \to (Y, f(x)) such that d_Y(f(x), f(z)) = d_Y(\delta f(x)(x), \delta f(x)(z)) + o(d_X(x, z)) as z \to x. This structure captures the local linear approximation in non-Euclidean settings without relying on operations. A foundational extension, due to Kirchheim, establishes that Lipschitz functions from open subsets of to arbitrary spaces are metrically differentiable with respect to . This theorem provides a prototype for broader settings by showing that the target structure suffices for differentiability when the domain is . For general metric measure spaces, Cheeger's theorem asserts that in doubling metric measure spaces supporting a (PI spaces), every function to \mathbb{R} is differentiable almost everywhere with respect to the measure. Here, differentiability is defined via a countable collection of coordinate charts forming a differentiable structure, with the derivative given by a linear functional on the approximate tangent space, which is a Euclidean space of controlled dimension. PI spaces are those where the measure is doubling—satisfying \mu(B(x, 2r)) \leq C \mu(B(x, r)) for some C > 0—and admit a : for f and balls B(y, r), \int_{B(y,r)} |f - f_{B(y,r)}| \, d\mu \lesssim r \left( \int_{\lambda B(y,r)} \mathrm{Lip} f^p \, d\mu \right)^{1/p} for some \lambda, p \geq 1. This result, proved in 1999, fills a significant gap by enabling a first-order calculus on non-smooth spaces like fractals. Independent proofs appeared later, confirming the theorem's robustness. Prominent examples include Carnot groups equipped with the Carnot-Carathéodory , such as the , which are PI spaces despite their sub-Riemannian geometry. In these spaces, the horizontal tangent structure allows functions to admit differentials almost , mirroring the case but with stratified approximations. An alternative generalization involves Alberti representations for currents, decomposing them into families of curves with approximate tangents, providing a rank-one-like structure without full differentiability. This approach, developed for currents in spaces, equates to Cheeger's in PI spaces and applies to representing measures via currents with bounded tangents. For instance, in the , such representations capture the non-trivial topology of 1-currents.

References

  1. [1]
    [PDF] Rademacher's theorem
    This theorem says that a Lipschitz function on Rn is differentiable almost everywhere. f(x + h) − f(x) = dfx(h) + 0(h). This means that for any > 0 there is a ...
  2. [2]
    [PDF] DIFFERENTIABILITY OF LIPSCHITZ FUNCTIONS ON METRIC ...
    In particular, we give a generalization of the theorem of Rademacher which asserts that a real valued Lipschitz function on Rn is differentiable almost ...
  3. [3]
    [PDF] GEOMETRIC ANALYSIS 1. The Rademacher theorem 1.1. The ...
    The Rademacher theorem. Lipschitz functions defined on one dimensional inter- vals are differentiable a.e. This is a classical result that is covered in ...
  4. [4]
    Lipschitz Function -- from Wolfram MathWorld
    A function f such that |f(x)-f(y)|<=C|x-y| for all x and y, where C is a constant independent of x and y, is called a Lipschitz function.Missing: continuity "real
  5. [5]
    Lipschitz Functions - Department of Mathematics at UTSA
    Nov 6, 2021 · In mathematical analysis, Lipschitz continuity, named after German mathematician Rudolf Lipschitz, is a strong form of uniform continuity for functions.
  6. [6]
    [PDF] 3.2 Function Space Preliminaries
    The Lipschitz property implies more than continuity, but less than differentiability. ... Continuous. Uniformly. Continuous. Locally Lipschitz. Lipschitz. C1.
  7. [7]
    [PDF] Notes for Optimization Algorithms Spring 2023 - Purdue Math
    Assume |f′(x)| is bounded by L for any x, we obtain Lipschitz continuity. Assume Lipschitz continuity, and take the limit y → x, we get |f′(x)| ≤. L. Thus ...
  8. [8]
    [PDF] LECTURES ON LIPSCHITZ ANALYSIS 1. Introduction A function f
    Distance functions are simple but important examples of Lipschitz functions. The distance can be taken either to a point x0 ∈ Rn,. (2.1) x 7→ dist(x, x0) ...
  9. [9]
    [PDF] On Lipschitz continuity of projections - arXiv
    Oct 6, 2019 · Our aim is to prove local lipschitzness of the projection mapping P : H×D→H, given by (1.1) at an arbitrary fixed (¯v, ¯p). The following ...
  10. [10]
    [PDF] Lipschitz Continuity - NET
    Example 12.3. A linear function f(x) = mx + b is Lipschitz continuous with Lipschitz constant Lf = |m| on the entire set of rational numbers Q. 2 − x2 1|. for ...
  11. [11]
    [PDF] Lebesgue Measure on Rn - UC Davis Mathematics
    Our goal is to construct a notion of the volume, or Lebesgue measure, of rather general subsets of Rn that reduces to the usual volume of elementary ...
  12. [12]
    [PDF] THE WEIERSTRASS PATHOLOGICAL FUNCTION - UCSD Math
    The example we give here is a faithful reproduction of Weierstrass's original 1872 proof. It is somewhat more complicated than the example given as Theorem 7.18 ...
  13. [13]
    Über partielle und totale differenzierbarkeit von Funktionen ...
    Über partielle und totale differenzierbarkeit von Funktionen mehrerer Variabeln und über die Transformation der Doppelintegrale. Download PDF. Hans Rademacher.
  14. [14]
  15. [15]
    [PDF] 11.2. Fundamental theorem of calculus. Question 11.12 ... - CMU Math
    Theorem 11.16. Let f : [a, b] → R be measurable. Then f is absolutely continuous if and only if f is differentiable almost everywhere, f ∈ L.
  16. [16]
    [PDF] FUNCTIONS OF BOUNDED VARIATION 1. Introduction In this paper ...
    We begin by defining uniform continuity and absolute continuity, and show that absolute continuity implies uniform continuity. ... properties of absolute.
  17. [17]
    245A, Notes 5: Differentiation theorems | What's new - Terry Tao
    Oct 16, 2010 · {f'} is equal almost everywhere to a monotone non-decreasing function, and so is itself almost everywhere differentiable. (Hint: Drawing the ...
  18. [18]
    THE LEBESGUE DIFFERENTIATION THEOREM VIA THE RISING ...
    Lebesgue's differentiation theorem [5] says that every monotone function is differentiable almost everywhere. Most proofs use Vitali's covering theorem.
  19. [19]
    [PDF] proof of rademacher's theorem - Arizona Math
    PROOF OF RADEMACHER'S THEOREM. Theorem. Let f(x) be a Lipschitz function in Rn. Then f(x) is differentiable almost everywhere. Proof. I will break the proof ...
  20. [20]
    [PDF] Introduction to Geometric Measure Theory - Stanford University
    The present text is a revision and updating of the author's 1983 “Lectures on Geomet- ric Measure Theory,” and is meant to provide an introduction to the ...
  21. [21]
    [PDF] Geometric Measure Theory
    May 8, 2021 · Theorem 2.7 (Rademacher's theorem). ... Next we prove a rectifiability result whose proof uses the Besicovitch-Federer structure theorem.
  22. [22]
    Local properties of solutions of elliptic partial differential equations
    Calderón, A., and Zygmund, A.. "Local properties of solutions of elliptic partial differential equations." Studia Mathematica 20.2 (1961): 181-225. <http ...Missing: non negative
  23. [23]
    [PDF] arXiv:1605.00233v1 [math.FA] 1 May 2016
    May 1, 2016 · In this paper, we prove that every Sobolev space Wl,p(Ω) with pl ≤ N if p 6= 1, or pl < N if p = 1, contains a closed infinite linear ...
  24. [24]
    [PDF] NOWHERE DIFFERENTIABLE SOBOLEV FUNCTIONS ... - metaphor
    ... p(B), for every p ∈ [1,n]. Thus, our counterexample shows that any Sobolev class W1,p(B) with p ∈ [1,n] contains a nowhere differentiable function.Missing: almost everywhere
  25. [25]
    DIFFERENTIABILITY OF MONOTONE SOBOLEV FUNCTIONS
    He proved that if u ∈ W1,1(Ω) is a function whose weak partial derivatives belong to the Lorentz space Ln,1(Ω), then u is differentiable almost everywhere. The.
  26. [26]
    Lp-Taylor approximations characterize the Sobolev space W1,p
    In this note, we introduce a variant of Calderón and Zygmund's notion of L p -differentiability – an L p -Taylor approximation.
  27. [27]
  28. [28]
    rectifiable metric spaces - American Mathematical Society
    We obtain the following substitution rule for integration with respect to Haus- dorff measure on rectifiable subsets of Banach spaces (see [4, Theorem 3.2.5].
  29. [29]