Fact-checked by Grok 2 weeks ago

Singular function

In , a singular function is a real-valued function defined on an that is differentiable with respect to but has equal to zero . Such functions are continuous and typically of , yet they fail to satisfy the in its classical form because their cannot be recovered from integrating their (vanishing) derivatives. Singular functions arise prominently in the Lebesgue decomposition theorem for functions of , where any such function f on [a, b] can be uniquely expressed as the sum f = f_{ac} + f_s + f_j, with f_{ac} absolutely continuous, f_s singular, and f_j a function. The singular component f_s captures the part of the increase that is not due to an absolutely continuous measure or discrete jumps, instead corresponding to a singular continuous measure in the associated Stieltjes measure. This decomposition highlights the pathological behavior of singular functions, which increase despite having zero nearly everywhere, illustrating the limitations of differentiability in . A canonical example of a singular function is the Cantor–Lebesgue function, also known as the , defined on [0, 1]. This function is continuous and strictly increasing from 0 to 1, constant on the intervals removed in the construction of the , and differentiable with derivative zero, yet it maps the (of measure zero) onto a set of positive measure. Introduced by in 1883 and later analyzed by , it serves as a fundamental counterexample in measure theory and has applications in fractal geometry and probability. More generally, singular functions like Lebesgue's singular function with parameter a (for $0 < a < 1, a \neq 1/2) are strictly increasing with derivative zero , providing models for studying sets where differentiability fails in specific ways.

Definition and basic properties

Formal definition

In real analysis, a singular function is a real-valued function f: [a, b] \to \mathbb{R} that is continuous on the closed interval [a, b], non-constant, and differentiable with derivative f'(x) = 0 almost everywhere on [a, b] with respect to Lebesgue measure. The continuity of f ensures the absence of jumps or discontinuities, allowing the function to map the compact interval continuously while exhibiting pathological differentiability behavior. The phrase "almost everywhere" refers to the property holding except on a set of Lebesgue measure zero, meaning the exceptional set where f' fails to exist or equals zero has negligible "size" in the measure-theoretic sense. This condition implies that the total variation contributed by the derivative vanishes, so that \int_a^b |f'(x)| \, dx = 0, yet the net change f(b) - f(a) \neq 0 due to the function's non-constancy. The term "singular function" originated in the early 20th century, introduced by in his studies of the differentiability of monotone functions, where such functions arise as the "singular" components in the decomposition of functions of bounded variation.

Key properties

Singular functions possess several fundamental properties stemming directly from their definition within the framework of functions of bounded variation. They are non-constant functions that are differentiable almost everywhere with respect to , yet their derivative vanishes almost everywhere. This counterintuitive behavior allows them to accumulate variation without "moving" on sets of full measure. Singular functions are inherently of on a closed interval [a, b]. For the typical case of an increasing singular function f, the total variation coincides with the net increase, satisfying V(f; [a, b]) = f(b) - f(a) > 0, while the absolutely continuous component contributes nothing to this variation. They are generally monotonic, often taken to be non-decreasing by considering the positive variation part via the Jordan decomposition of bounded variation functions. A hallmark property is their role in the Lebesgue decomposition of functions of . Any such f admits a unique representation f(x) = f(a) + \int_a^x f'(t) \, dt + s(x), where the term is the absolutely continuous part (with respect to ) and s(x) is the singular satisfying s'(t) = 0 , though s is non-constant and thus the of |s'(t)| over [a, b] is zero. This decomposition isolates the singular as the "purely singular" component, which is unique. Regarding differentiability, singular functions are differentiable with zero , but they exhibit non-differentiability on dense sets, reflecting their pathological nature despite the almost everywhere regularity. This aligns with their placement in the broader Lebesgue decomposition theorem, where the singular part is orthogonal to the absolutely continuous measures.

Examples

Cantor function

The , also known as the , serves as the prototypical example of a singular function. It was introduced by in a letter dated November 1883 and elaborated in his paper "Über unendliche, lineare Punktmannichfaltigkeiten, V," published in 1883. The function was later recognized as singular—continuous and non-decreasing but with derivative zero —by in 1904. The c: [0,1] \to [0,1] is constructed alongside the middle-thirds C, which is formed by iteratively removing the open middle third from each subinterval of [0,1], starting with C_0 = [0,1], C_1 = [0,1/3] \cup [2/3,1], and so on, yielding C = \bigcap_{n=0}^\infty C_n with zero. On each open interval removed at stage n, c is defined to be constant, taking the value it attains at the endpoints of that interval (which belong to C_n); the values are chosen so that c increases from 0 to 1 only across the , ensuring monotonicity. This iterative process defines c on the complement of C and extends continuously to all of [0,1]. An explicit formula arises from ternary expansions: for x \in [0,1], express x = \sum_{k=1}^\infty a_k 3^{-k} with a_k \in \{0,1,2\}, choosing representations without infinite tails of 1's when possible. For x \in C (where no a_k = 1), set b_k = a_k / 2 (so b_k \in \{0,1\}) and define c(x) = \sum_{k=1}^\infty b_k 2^{-k}. For x \notin C, c(x) equals c(y) where y is the left endpoint of the removed interval containing x. The Cantor function is continuous and non-decreasing, with c(0) = 0 and c(1) = 1, and it is surjective onto [0,1]. Its derivative exists almost everywhere and equals zero a.e., as c is constant on the complement of C (measure 1) and C has measure zero, verifying its singularity. It remains constant on each complementary interval to C. The restriction c|_C: C \to [0,1] is a homeomorphism, establishing a topological equivalence between the Cantor set and the unit interval.

Lebesgue's singular function

Another important example is Lebesgue's singular function L_a, parameterized by a \in (0,1) \setminus \{1/2\}. Defined on [0,1], it is constructed iteratively similar to the but with asymmetric removals or mass distributions, resulting in a strictly increasing with zero . Unlike the , L_a has no intervals of constancy and increases throughout, providing a model for singular functions without flat parts. It was introduced by to study points of differentiability failure.

Constructions via measures

Singular functions can be constructed generally as the cumulative distribution functions (CDFs) of singular continuous measures on a closed interval [a, b]. A singular continuous measure \mu is a probability measure that is continuous (assigning zero mass to singletons, i.e., no atoms) and mutually singular with respect to Lebesgue measure \lambda, meaning there exists a set E \subset [a, b] such that \mu(E^c) = 0 and \lambda(E) = 0. Such measures are supported on sets of Lebesgue measure zero but carry positive total mass, \mu([a, b]) > 0. The associated singular function f: [a, b] \to [0, 1] is then defined by normalizing the CDF, ensuring f(a) = 0 and f(b) = 1, and is continuous, non-decreasing, and differentiable almost everywhere with derivative zero, yet f(b) - f(a) = 1 > 0. To construct such a function, begin with a singular continuous measure \mu supported on a set of zero, such as a set like the or a more general self-similar structure. The function is obtained by integrating the measure: compute f(x) = \mu([a, x]) for x \in [a, b], which yields a non-decreasing function that increases only on the support of \mu. This integration leverages the Lebesgue-Stieltjes construction, where \mu defines a measure via the differences of f, and the singularity ensures the function has no intervals of positive length over which it varies in a way compatible with . Explicit constructions often involve iterative processes, such as removing intervals from [a, b] while assigning masses that concentrate on the remaining set of measure zero. The defining equation for the singular function is f(x) = \int_{[a, x]} d\mu(t), where \mu \perp \lambda and \mu([a, b]) > 0. This formulation directly ties the function to measure theory, as the total variation of f equals the total mass of \mu. All singular functions of on [a, b] arise precisely in this manner, as the CDFs of singular continuous measures, providing a complete characterization within the framework of Lebesgue . For instance, the serves as a concrete example of this construction, where \mu is the supported on the middle-thirds . Beyond the Cantor case, singular functions can be derived from invariant measures of certain dynamical systems on . For piecewise-smooth circle homeomorphisms with break points (points of derivative discontinuity) and irrational rotation number, the unique ergodic measure is singular with respect to , and its CDF yields a singular function. Similarly, devil's staircases appear on other fractal supports, such as self-similar sets generated by systems with unequal contraction ratios, where the associated induces a singular continuous distribution whose CDF is a generalized singular function.

Theoretical context

Relation to bounded variation

A function f: [a, b] \to \mathbb{R} is said to be of if its V(f; [a, b]) = \sup_P \sum_{i=1}^n |f(x_{i+1}) - f(x_i)| is finite, where the supremum is taken over all partitions P = \{a = x_1 < x_2 < \cdots < x_{n+1} = b\} of [a, b]. This concept was introduced by Camille Jordan in his 1881 paper on Fourier series, where he studied the properties of such functions in the context of convergence. Jordan's key contribution was the decomposition theorem stating that every function of bounded variation can be expressed as the difference of two non-decreasing functions, f = g - h, where g and h are non-decreasing on [a, b]. This Jordan decomposition highlights the structure of bounded variation functions and implies that they are differentiable almost everywhere with respect to Lebesgue measure, as non-decreasing functions possess this property. Building on Jordan's work, Henri Lebesgue extended the decomposition in his 1902 doctoral thesis by showing that every function f of bounded variation admits a unique decomposition f = f_{ac} + f_j + f_s, where f_{ac} is absolutely continuous, f_j is a jump function, and f_s is singular. Here, f_{ac}(x) = \int_a^x f'(t) \, dt represents the absolutely continuous part, with f' denoting the almost everywhere derivative of f, f_j(x) is the sum of the jumps of f up to x, while the singular part f_s = f - f_{ac} - f_j is continuous, satisfies f_s' = 0 almost everywhere, and is not constant. Singular functions thus form a subclass of continuous functions of bounded variation, characterized by their continuity, bounded variation, and vanishing derivative almost everywhere. A brief sketch of the proof for the Lebesgue decomposition proceeds as follows: since f is of bounded variation, Jordan's theorem yields f = g - h with g, h non-decreasing, and each is differentiable almost everywhere by Lebesgue's differentiation theorem for monotone functions; define f'(t) = g'(t) - h'(t) almost everywhere, set f_{ac}(x) = f(a) + \int_a^x f'(t) \, dt; the remainder f - f_{ac} has derivative zero almost everywhere and is further decomposed into its jump part f_j (saltus function, capturing discontinuities) and continuous singular part f_s = (f - f_{ac}) - f_j; the total variation remains bounded, and f_s is continuous with derivative zero almost everywhere. Uniqueness follows from the measure-theoretic decomposition, where if f = f_{ac_1} + f_{j_1} + f_{s_1} = f_{ac_2} + f_{j_2} + f_{s_2}, the differences between corresponding components are constant almost everywhere.

Lebesgue decomposition

The Lebesgue decomposition theorem provides a canonical way to express any monotone function as the sum of three components that are mutually singular with respect to Lebesgue measure: an absolutely continuous part, a discrete (purely atomic or jump) part, and a singular continuous part. This decomposition arises from applying the Radon-Nikodym theorem to the Lebesgue-Stieltjes measure induced by the monotone function and is essential for analyzing the distributional properties of such functions. For a non-decreasing function f: [a, b] \to \mathbb{R}, the associated measure \mu_f (defined by \mu_f((c, d]) = f(d) - f(c) for a \leq c < d \leq b) decomposes uniquely as \mu_f = \mu_{ac} + \mu_d + \mu_s, where \mu_{ac} \ll \lambda (absolutely continuous with respect to Lebesgue measure \lambda), \mu_d is purely atomic (concentrated on a countable set of points corresponding to jumps), and \mu_s \perp \lambda with no atoms (singular continuous). The corresponding function decomposition is f(x) = f_{ac}(x) + f_d(x) + f_s(x), where f_{ac}(x) = \int_a^x g(t) \, dt for some g \in L^1[a, b] (with g = f' almost everywhere, by the differentiability theorem for monotone functions), f_d(x) = \sum_{\{t \leq x : f \text{ discontinuous at } t\}} (f(t+) - f(t-)) captures the countable sum of jump sizes up to x, and f_s is the singular continuous component, which is continuous but increases only on a set of Lebesgue measure zero. The Radon-Nikodym derivative provides the density g = \frac{d\mu_{ac}}{d\lambda}, ensuring the absolutely continuous part recovers the integral of the derivative, while the singular parts account for the remaining increase without density with respect to \lambda. This structure highlights how monotone functions of bounded variation extend beyond the broader class by incorporating measure-theoretic singularity. The decomposition is unique up to equivalence classes modulo sets of Lebesgue measure zero, meaning that if f = f_{ac}' + f_d' + f_s' is another such representation, then the differences between corresponding components are constant almost everywhere. A key property of the singular continuous part f_s is that f_s'(x) = 0 for almost every x \in [a, b], yet f_s is strictly increasing over intervals where it concentrates its variation on null sets, exemplifying pathological behavior absent in absolutely continuous functions. This theorem, originally developed for monotone functions, underpins the general of measures and was introduced by in his 1902 doctoral thesis Intégrale, longueur, aire.

Applications

In probability

In probability theory, a singular continuous distribution is a probability measure that is continuous, meaning it has no atoms or point masses, and is singular with respect to the Lebesgue measure, meaning it is concentrated on a set of Lebesgue measure zero. The cumulative distribution function (CDF) of such a distribution is a singular function: it is continuous and non-decreasing, but its derivative exists and equals zero almost everywhere with respect to Lebesgue measure. Formally, for a random variable X with a singular continuous distribution, the CDF satisfies P(X \leq x) = f(x), where f is singular, ensuring no probability density exists while the distribution remains continuous. Singular continuous distributions often have support on Cantor-like sets, which are compact, nowhere dense subsets of the real line with positive but zero Lebesgue measure. These distributions are useful in modeling probabilistic phenomena that exhibit fractal structure, lacking both a density with respect to Lebesgue measure and discrete point masses, such as certain self-similar processes or limit laws in dynamical systems. Examples include distributions arising from infinite convolutions, such as , which are the infinite convolution products of symmetric scaled by a parameter \lambda \in (1/2, 1); for most \lambda, these yield singular continuous measures supported on . Another class involves associated with defined with respect to singular continuous weights, as constructed in studies of where the orthogonality measure is singular. In quantum mechanics, the singular continuous spectrum of self-adjoint operators corresponds to singular continuous components in the associated spectral measures, linking these distributions to the analysis of wave propagation and scattering without bound states or absolutely continuous parts.

In physics

Singular functions arise in physical models of solids and dynamical systems, where they describe phenomena involving incommensurate structures, phase transitions, and invariant measures that exhibit zero derivative almost everywhere but continuous variation. These functions, often manifesting as devil's staircases, capture the locking of phases or displacements on dense sets of parameters while remaining constant elsewhere, reflecting the absence of smooth variation in certain physical responses. In the Frenkel-Kontorova model, which simulates atomic chains interacting via nearest-neighbor springs and a substrate potential, singular functions describe atomic displacements in incommensurate structures. The model exhibits a devil's staircase in the ground-state configuration as a function of the misfit parameter between the natural lattice spacing and the substrate periodicity, where the average atomic spacing locks into rational values over intervals, forming a singular continuous function akin to the . This staircase arises from the competition between harmonic interactions and the sinusoidal potential, leading to discommensurations—localized regions of mismatch—that mediate the phase locking, with the overall displacement field showing singular behavior in its derivative. The axial next-nearest-neighbor Ising (ANNNI) model, used to study competing ferromagnetic and antiferromagnetic interactions along different axes in magnetic systems, also features singular functions in its phase diagram. Phase transitions exhibit a devil's staircase structure in the modulation wave number versus the frustration parameter, where commensurate phases occupy dense intervals, resulting in a singular continuous variation of the order parameter. This reflects the separating ferromagnetic, paramagnetic, and antiphase regions, with the staircase capturing the multiplicity of modulated phases driven by frustration. In the fractional quantum Hall effect, wavefunctions and associated probability densities involve singular continuous measures, particularly at quantum Hall transitions. Multifractal analysis reveals that the local density of states or squared wavefunction amplitudes scale with singular continuous spectra, where the measure is supported on sets of Lebesgue measure zero but exhibits intricate scaling properties. This multifractality underscores the critical nature of the integer and fractional quantum Hall plateau transitions, linking to the topological robustness of edge states and delocalization. Dynamical systems, such as those governed by circle diffeomorphisms, yield singular functions through their invariant measures. For critical circle homeomorphisms with irrational rotation numbers, the unique ergodic invariant measure is absolutely singular with respect to Lebesgue measure, implying a cumulative distribution function that is continuous but has derivative zero almost everywhere. These measures describe the long-term distribution of orbits in quasiperiodic motions, relevant to physical systems like Josephson junctions or coupled oscillators exhibiting mode locking. Singular functions appear in models of dislocations and defects, where the variation in strain or displacement occurs on zero-measure sets, such as slip planes or core regions. In the Frenkel-Kontorova framework for dislocations, the plastic distortion is concentrated along lines of measure zero, leading to singular components in the total deformation field that capture the topological defects without smooth gradients elsewhere.

Other uses of the term

Functions with singularities

In complex analysis, a function is said to have a singularity at a point z_0 in the complex plane if it fails to be analytic there, meaning it cannot be represented by a power series in a neighborhood of z_0. Such points disrupt the function's differentiability and continuity in the complex sense, often leading to unbounded or oscillatory behavior nearby./05:_Chapter_5/5.04:_Classification_of_Singularities) Singularities are broadly classified by their isolation and nature. An isolated singularity exists at z_0 if there is a punctured disk around it where the function is analytic; these are subdivided into removable singularities (where the limit exists, allowing redefinition to make the function analytic), poles (where the function grows like (z - z_0)^{-n} for finite positive integer n), and essential singularities (where the function exhibits wild, non-polynomial-like growth, as per Picard's theorem)./05:_Chapter_5/5.04:_Classification_of_Singularities) Non-isolated singularities, in contrast, occur when singularities accumulate, forming a cluster or natural boundary, such as the points on the unit circle for the function \sum_{n=1}^\infty z^{2^n}. Branch points represent another category, typically non-isolated, where the function becomes multi-valued, as in the complex logarithm \log z at z=0, requiring a branch cut to define a single-valued version./05:_Chapter_5/5.04:_Classification_of_Singularities) Classic examples illustrate these types. The function f(z) = \frac{1}{z} has a simple pole (a pole of order 1) at z=0, where |f(z)| \to \infty as z \to 0, and its residue is 1./05:_Chapter_5/5.04:_Classification_of_Singularities) In contrast, f(z) = e^{1/z} possesses an essential singularity at z=0, as its Laurent series has infinitely many negative powers, and by the Casorati-Weierstrass theorem, the image of any neighborhood of 0 densely fills the complex plane. The structure of a singularity, particularly isolated ones, is analyzed via the Laurent series expansion around z_0: f(z) = \sum_{n=-\infty}^{\infty} a_n (z - z_0)^n, which converges in an annulus r < |z - z_0| < R. The principal part \sum_{n=-\infty}^{-1} a_n (z - z_0)^n classifies the singularity: it vanishes for removable cases, is finite for poles, and infinite for essential singularities./08:_Taylor_and_Laurent_Series/8.07:_Laurent_Series) This interpretation of singular functions in complex analysis differs from its occasional use in real analysis to describe distributions like the Dirac delta, which regularize improper integrals.

Singularity functions in engineering

In structural engineering, singularity functions, often referred to as , provide a mathematical framework for modeling discontinuous loading conditions in beams and frames. These functions are defined as \langle x - a \rangle^n, where x is the position along the beam, a is the location of the discontinuity, and n is a non-negative integer; the function equals zero for x < a and (x - a)^n for x \geq a. This notation allows engineers to express the effects of abrupt changes, such as point loads or moments, without piecewise definitions. The functions extend to negative exponents to represent distributions like the ; for instance, \langle x - a \rangle^{-1} corresponds to a unit impulse at x = a, and \langle x - a \rangle^{-2} to a unit doublet. Introduced by William H. Macaulay in 1919, these functions were developed to derive closed-form deflection equations for indeterminate beams under complex loading, overcoming the limitations of prior methods that required separate expressions for different beam segments. Prior to this, beam deflection analysis relied on fragmented moment equations, making it cumbersome for structures with multiple supports or loads. Macaulay's approach, published in "Note on the Deflection of Beams," enabled a unified elastic curve equation by incorporating the functions into the moment expression of the . In beam theory, singularity functions facilitate the analysis of shear force V(x), bending moment M(x), slope \theta(x), and deflection v(x) using a single governing equation, even for beams with discontinuous loads. The bending moment is expressed as M(x) = EI \frac{d^2 v}{dx^2}, where E is the modulus of elasticity and I is the moment of inertia; substituting singularity terms for loads allows double integration to yield deflection directly. This method is particularly advantageous for statically determinate and indeterminate beams, reducing the need for multiple integrations and boundary condition applications across segments. Boundary conditions are applied only at the ends, with intermediate discontinuities handled inherently by the functions. Representative examples illustrate their application. A unit step function \langle x - a \rangle^0 models a point load's onset, producing a constant shear change beyond a. A linear ramp \langle x - a \rangle^1 represents a uniformly distributed load starting at a, leading to a parabolic moment increase. For concentrated forces, the Dirac delta \langle x - a \rangle^{-1} captures an impulsive shear jump, essential for point loads in shear diagrams. These functions ensure continuity in higher-order responses like deflection while accommodating load discontinuities. The integration of singularity functions follows a consistent rule: \int \langle x - a \rangle^n \, dx = \frac{1}{n+1} \langle x - a \rangle^{n+1} + C for n \geq -1, with the constant C determined from boundary conditions; this holds because the functions are zero before a, preserving the integration limits. For negative exponents, differentiation reverses the process: \frac{d}{dx} \langle x - a \rangle^n = n \langle x - a \rangle^{n-1} for n > 0, and \frac{d}{dx} \langle x - a \rangle^0 = \langle x - a \rangle^{-1}. This property streamlines the derivation of from or load from in analysis.

References

  1. [1]
    [PDF] 3.5 absolutely continuous and singular functions - Christopher Heil
    Definition 3 (Singular Function). A function f : [a, b] → C or f : R → C is singular if f is differentiable at almost every point in its domain and f′ = 0 a.e.
  2. [2]
    [PDF] Real Analysis lecture notes for MA 645/646
    where is called a singular function. Continuous functions of bounded ... Real analysis. Pure and Applied Mathematics (New York). John Wiley & Sons,. Inc ...
  3. [3]
    On the set of points where Lebesgue's singular function has the ...
    As is well known, La(x) L a ( x ) is strictly increasing and has a derivative equal to zero almost everywhere.
  4. [4]
    [PDF] Proving the characterization of Archimedean copulas via Dini ...
    We understand a strictly increasing (decreasing) singular function as a continuous and strictly increasing (decreasing) function with derivative zero almost ...
  5. [5]
    [PDF] arXiv:2008.13217v2 [math.DS] 3 Oct 2020
    Oct 3, 2020 · F is uniformly continuous,. 3. F is differentiable with derivative zero almost everywhere, and. 4. F is a singular function. Proof of Theorem 4 ...
  6. [6]
    [PDF] Companion to Real Analysis - Portland State University
    Mar 8, 2010 · ) The Lebesgue singular function is increasing and continuous. It is differentiable, with derivative zero, almost everywhere so that ψ(1) ...<|control11|><|separator|>
  7. [7]
    [PDF] Sur les intégrales singulières - Numdam
    lequel une intégrale indéfinie a presque partout pour dérivée la fonction intégrée il résulte que la fonction c~ = F (x) a une dérivée nulle presque partout.
  8. [8]
    [PDF] Real Analysis MAA 6616 Lecture 22 Absolutely Continuous Functions
    gives a decomposition of a function with bounded variation as the sum of an absolutely continuous function and a singular function. Theorem (7). Let f ∈ BV ...
  9. [9]
    [PDF] Appendix: Functions of one variable - UC Davis Mathematics
    has bounded variation on [0, 1]. The total variation of both functions is ... Sometimes a singular function is required to be continuous, but our definition ...
  10. [10]
    [PDF] A Note on the History of the Cantor Set and Cantor Function
    In this note, we would like to give some of that history, a sketch of the ideas under consideration at the time of their discovery, and a hypothesis regarding ...Missing: Scheeffer 1905
  11. [11]
    [PDF] The Devil's Staircase - UCSD Math
    It is called the Cantor Function or the Devil's staircase. Its graph is shown in Figure 2. It can actually be described in simple terms.Missing: original sources
  12. [12]
    [PDF] The Cantor function
    In fact, Proposition 4.10 remains valid for an arbitrary singular function of bounded variation. Theorem 4.13 Let F : R → R be a function of bounded variation ...<|control11|><|separator|>
  13. [13]
    ON THE DECOMPOSITION THEOREMS OF LEBESGUE AND ...
    If F is an additive function of bounded variation of an interval, the derivative. F. 0 is summable, and the function F is the sum of a singular additive ...
  14. [14]
    References - Fourier Analysis
    Jordan, C. 1881. Sur la série de Fourier. C. R. Acad. Sci. Paris, 92, 228–230.Google Scholar. Kahane, J.-P. 2000. Baire's category theorem and trigonometric ...<|control11|><|separator|>
  15. [15]
    [PDF] The reverse mathematics of theorems of Jordan and Lebesgue - arXiv
    Jan 8, 2021 · The Jordan decomposition theorem states that every function f : [0, 1] → R of bounded variation can be written as the difference of two non- ...
  16. [16]
    [PDF] Intégrale, Longueur, aire - Internet Archive
    Intégrale, Longueur, Aire. 2e ...
  17. [17]
    [PDF] Differentiation Lecture 7, Following Folland, ch 3.1, 3.2
    Functions of Bounded variation and Lebesgue-Stieltjes measures. We know that each increasing right continuous function F corresponds to a Radon mea- sure on R.
  18. [18]
    [PDF] DIFFERENTIATION OF MONOTONE FUNCTIONS 1. Dini derivates ...
    |F(xi) − F(xi−1)| : a = x0 < x1 < ··· < xn = b} < ∞. Note that if F is a monotone function ... (Lebesgue Decomposition Theorem) Let F : [a, b] → R be an increas-.
  19. [19]
    Henri Lebesgue - Biography
    ### Summary of Lebesgue's 1902 Thesis
  20. [20]
    Singular distribution - Encyclopedia of Mathematics
    Aug 3, 2014 · A distribution is singular if the corresponding distribution function is continuous and its set of growth points has Lebesgue measure zero.
  21. [21]
    On Continuous Singular Infinitely Divisible Distribution Function
    A probability distribution function F F is said to be infinitely divisible if, for every integer n n , there is a distribution function Fn F n such that F F ...<|control11|><|separator|>
  22. [22]
  23. [23]
    [PDF] Sixty years of Bernoulli convolutions - Yale Math
    Bernoulli convolutions are the infinite convolution product of measures, studied since the 1930s, with connections to harmonic analysis and dynamical systems.
  24. [24]
    Polynomials orthogonal with respect to singular continuous measures
    Sep 28, 2006 · A class of invariant measures and the orthogonal polynomials associated with these measures are constructed and discussed.
  25. [25]
    Physics - Project Euclid
    It is already known from the inverse method of scattering theory [9] that singular continuous spectrum can occur. Aronszajn [10] has constructed an example ...
  26. [26]
    [PDF] Devil's staircase and order without periodicity in classical ... - HAL
    Feb 4, 2008 · Frenkel-Kontorova model (2). We describe our early results in terms of defect properties. (Here the defects are discommensurations and can ...
  27. [27]
    None
    Nothing is retrieved...<|separator|>
  28. [28]
    None
    Nothing is retrieved...<|separator|>
  29. [29]
    Singular measures in circle dynamics
    Critical circle homeomorphisms have an invariant measure totally singular with respect to the Lebesgue measure. We prove that singularities of the invarian.
  30. [30]
    Singularity -- from Wolfram MathWorld
    In general, a singularity is a point at which an equation, surface, etc., blows up or becomes degenerate. Singularities are often also called singular points.
  31. [31]
    [PDF] All about the Dirac Delta Function(?) --------~--------
    where Xo is any real number. The use of the delta function became more and more common after the appearance of Dirac's work. Other singular functions also made ...
  32. [32]
  33. [33]
    [PDF] beams: deformation by singularity functions
    A singularity function is an expression for x written as. , where n is any integer. (positive or negative) including zero, and x0 is a constant equal to the ...
  34. [34]
    (PDF) Deflection of Flexural Members - Macaulay's Method 3rd Year ...
    Before Macaulay's paper of 1919, the equation for the deflection of beams could not be found in closed form. Different equations for bending moment were ...
  35. [35]
    [PDF] Chapter 8 DEFLECTION OF BEAMS BY INTEGRATION
    8.3). Using singularity functions, express the bending moment corresponding to the beam and loading shown (Fig. 8.29a).
  36. [36]
    Singularity Functions - RoyMech
    A singularity function is expressed as where n = any integer (positive or negative) including zero a = distance on x axis along the beam,from the selected ...<|control11|><|separator|>