Fact-checked by Grok 2 weeks ago

Weierstrass function

In , the Weierstrass function is a real-valued defined on the real numbers that exemplifies a pathological case in : it is everywhere but differentiable nowhere. Constructed by the German mathematician , the was first presented in a to the on July 18, 1872, providing the inaugural explicit to the then-widespread belief among mathematicians that every is differentiable except possibly at a set of isolated points. This discovery upended foundational assumptions in and , demonstrating that alone does not guarantee even almost-everywhere differentiability, and it spurred deeper investigations into the nature of in functions. The standard form of the Weierstrass function is given by the infinite series
f(x) = \sum_{n=0}^{\infty} a^n \cos(b^n \pi x),
where 0 < a < 1 is the amplitude scaling factor and b > 1 is typically chosen as an odd integer to ensure the function is nowhere differentiable, with the key condition ab > 1 + \frac{3\pi}{2} guaranteeing the pathological behavior. The series converges uniformly on the real line due to the decay of a^n, establishing at every point, but the increasingly rapid oscillations induced by the b^n frequencies prevent the from having a anywhere, thus ensuring non-differentiability. Variations of this construction, such as using sine terms or random phases, preserve these core properties while allowing for generalizations in fractal geometry and stochastic processes.
Beyond its theoretical significance, the Weierstrass function has influenced diverse fields, including the study of —its graph is self-similar and possesses a strictly between 1 and 2 for appropriate parameters, reflecting its "jagged" structure at all scales. It also finds applications in modeling irregular phenomena, such as fractal media in physics and , where its nowhere-differentiable nature captures non-smooth, scale-invariant behaviors. Early extensions, like those explored in the early , further quantified its irregularity, confirming properties such as only on sets of measure zero.

Introduction

Definition and Overview

The Weierstrass function is a classic example in , defined as the infinite series w(x) = \sum_{n=0}^{\infty} a^n \cos(b^n \pi x), where $0 < a < 1, b is an odd integer greater than 1, and the parameters satisfy ab > 1 + \frac{3\pi}{2}. This construction ensures on the real line, yielding a that is continuous everywhere. The function's significance lies in its pathological behavior: it is continuous at every point but differentiable , directly countering the early 19th-century mathematical that continuous functions are differentiable except possibly at isolated points. This property challenged foundational assumptions in and , demonstrating that does not imply differentiability even on dense sets. Karl Weierstrass introduced this example in a 1872 lecture to exhibit a function with these counterintuitive traits, marking a pivotal moment in the study of function regularity.

Historical Background

In the early 19th century, mathematicians began questioning the intuitive assumptions about the regularity of continuous functions, particularly regarding differentiability. Augustin-Louis Cauchy, in his 1821 Cours d'analyse, introduced rigorous definitions of continuity and limits but operated under the prevailing belief that continuous functions were differentiable almost everywhere, though he acknowledged cases where derivatives might fail to exist at isolated points or exhibit discontinuities. Similarly, Bernard Bolzano, in an unpublished 1834 manuscript, constructed an example of a continuous nowhere differentiable function, highlighting early awareness of pathological behaviors without fully resolving the issue of nowhere differentiability. Karl Weierstrass advanced this inquiry decisively during a lecture to the on July 18, 1872, where he presented the first rigorous construction of a that is everywhere but differentiable . This example challenged the foundations of by demonstrating that does not imply differentiability at any point. Although Weierstrass's proof circulated privately, it was first published in summary form by his student Paul du Bois-Reymond in 1875, who also proposed related constructions with the weaker condition ab > 1 for non-differentiability, while the original condition ab > 1 + \frac{3\pi}{2} ensured the pathological behavior. The full details of Weierstrass's 1872 lecture appeared posthumously in 1895, solidifying its place as a landmark in . The introduction of Weierstrass's function provoked strong reactions among contemporaries, reflecting broader debates on the nature of mathematical objects. , in his 1886 work Les Méthodes nouvelles de la mécanique céleste, decried such pathological examples as an "outrage to ," while Hermite labeled continuous nowhere-differentiable functions a "deplorable evil." Du Bois-Reymond, despite his role in disseminating the result, contributed to the discourse by exploring similar irregular functions in his 1875 publication. These responses underscored the function's disruptive impact on classical intuitions about smoothness. Recent scholarship continues to reflect on the Weierstrass function's enduring influence, as seen in a 2025 Quanta Magazine article that revisits its role in upending 19th-century views of calculus and inspiring modern fractal geometry.

Construction

Infinite Series Formula

The Weierstrass function is constructed as the infinite series w(x) = \sum_{n=0}^{\infty} a^n \cos(b^n \pi x), where $0 < a < 1 and b > 1 is an odd integer satisfying certain conditions to ensure pathological behavior. This series arises as the pointwise limit of the partial sums s_N(x) = \sum_{n=0}^{N} a^n \cos(b^n \pi x), each of which is an infinitely differentiable function because it is a finite linear combination of the smooth cosine functions. As N \to \infty, the sequence \{s_N\} converges to w(x), yielding a function that inherits continuity from the uniform convergence of the series but exhibits more complex properties due to the accumulating oscillations. The of the series on \mathbb{R} follows from the . For each term, |\cos(b^n \pi x)| \leq 1 for all real x and nonnegative integers n, so |a^n \cos(b^n \pi x)| \leq a^n = M_n. The series \sum_{n=0}^{\infty} M_n = \sum_{n=0}^{\infty} a^n = \frac{1}{1-a} < \infty since a < 1. Thus, by the M-test, the original series converges uniformly and absolutely on \mathbb{R}, implying that w(x) is continuous everywhere as the uniform limit of continuous partial s. The cosine function is selected for its inherent periodicity, which facilitates self-similarity across scales: each successive term a^n \cos(b^n \pi x) introduces higher-frequency oscillations that mimic the overall shape at finer resolutions, scaled by the damping factor a^n. The inclusion of \pi in the argument aligns the periods appropriately with the integer powers of b, ensuring the terms interact to produce the desired fractal-like structure without phase mismatches. This series construction admits generalizations by replacing the cosine with other periodic or self-similar building blocks while preserving uniform convergence under similar damping conditions. Similar infinite series constructions yield other continuous nowhere-differentiable functions, such as the , defined as t(x) = \sum_{n=0}^{\infty} 2^{-n} \phi(2^n x), where \phi(y) = \min_{k \in \mathbb{Z}} |y - k| is the distance-to-nearest-integer function (or equivalently, a symmetric tent map on [0,1]). This piecewise linear base yields another continuous nowhere-differentiable function with uniform convergence via the M-test using M_n = 2^{-n} and \sum M_n < \infty.

Parameter Choices and Convergence

The Weierstrass function is defined via an infinite series involving parameters a and b, where $0 < a < 1 ensures the amplitudes of successive terms decay geometrically, and b is chosen as a positive odd integer greater than 1 to control the increasing frequency of oscillations. The parameter a governs the vertical scaling, reducing the contribution of higher-frequency terms, while b dictates the horizontal compression, producing increasingly rapid wiggles that accumulate to create the function's pathological behavior. These choices are essential for balancing decay and growth to achieve continuity without differentiability. To guarantee nowhere differentiability, the parameters must satisfy ab > 1 + \frac{3\pi}{2}, a condition originally established by Weierstrass in his 1872 to ensure that the oscillations prevent the existence of a finite at any point. Later analysis by in 1916 showed that the weaker condition ab \geq 1 suffices for non-differentiability, provided b is an odd , broadening the family of such functions. If instead ab < 1, the series term's growth is insufficient to disrupt smoothness, rendering the resulting function differentiable everywhere. Regarding convergence, the series converges uniformly on \mathbb{R} by the Weierstrass M-test, as the absolute value of each term is bounded by a^n, and \sum a^n < \infty since $0 < a < 1. This uniform convergence implies the limit function is continuous everywhere, inheriting the continuity of the smooth partial sums. Moreover, the absolute convergence of the series ensures the function is well-defined pointwise on \mathbb{R}, with no issues from conditional convergence.

Core Properties

Everywhere Continuous

The Weierstrass function w(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x), defined with parameters $0 < a < 1 and b an odd integer greater than 1, is continuous at every point on the real line. This continuity arises from the uniform convergence of the series on \mathbb{R}. Each partial sum w_N(x) = \sum_{n=0}^N a^n \cos(b^n \pi x) is a finite sum of continuous functions and thus continuous everywhere. To establish uniform convergence, apply the Weierstrass M-test: the absolute value of each term satisfies |a^n \cos(b^n \pi x)| \leq a^n, and the series \sum_{n=0}^\infty a^n converges to $1/(1-a) since $0 < a < 1. Therefore, the original series converges absolutely and uniformly on \mathbb{R}. By the theorem on uniform limits of continuous functions, the limit w(x) is continuous on \mathbb{R}. The function's periodicity with period 2 further supports its continuity across the entire real line. Specifically, w(x+2) = \sum_{n=0}^\infty a^n \cos(b^n \pi (x+2)) = \sum_{n=0}^\infty a^n \cos(b^n \pi x + 2 b^n \pi) = w(x), since b is an integer and cosine is $2\pi-periodic. Combined with the boundedness of w(x), where |w(x)| \leq \sum_{n=0}^\infty a^n = 1/(1-a), this periodicity ensures the function remains well-behaved and continuous without discontinuities at any points. In contrast to standard Fourier series, which typically involve dense sets of frequencies to represent smooth functions, the Weierstrass series is lacunary, featuring frequencies b^n \pi that exhibit exponentially growing gaps due to the integer b \geq 3. This sparse structure contributes to the function's irregularity but preserves its continuity, as the uniform convergence depends solely on the decay parameter a, not the frequency spacing. The continuity property holds for any parameter choices ensuring series convergence, particularly under the conditions $0 < a < 1 and ab > 1 with b an odd integer, as these guarantee the necessary without altering the continuous nature of the limit.

Nowhere Differentiable

The Weierstrass function exhibits nowhere differentiability, a property first demonstrated by in 1872, meaning the limit defining the fails to exist at every point in its domain. The core proof relies on analyzing the \frac{w(x + h) - w(x)}{h} for the function w(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x), where $0 < a < 1 and b is an odd integer greater than 1. As h \to 0, higher-order terms in the series dominate due to their rapidly increasing frequencies, causing the quotient to oscillate with unbounded amplitude and prevent convergence to any finite value. To establish this rigorously, fix a point x and consider a sequence h_m \to 0 chosen such that h_m = \frac{1}{b^m} or a slight perturbation to align the cosine phases for maximum variation. The difference w(x + h_m) - w(x) can then be decomposed into the sum up to m-1, the m-th term, and the remainder tail. The initial partial sum is bounded by a geometric series summing to less than \frac{a}{1 - a}, while the m-th term contributes up to 2 (ab)^m in magnitude to the quotient when the phase shift is π (optimized cosine difference of 2). The low terms (n < m) contribute at most \sum_{n=0}^{m-1} \pi a^n b^n \leq \frac{\pi (ab)^m}{ab - 1}. The tail from n = m+1}^\infty satisfies |tail difference| ≤ \sum_{n=m+1}^\infty 2 a^n = \frac{2 a^{m+1}}{1 - a}, so the tail contribution to the quotient is at most \frac{2 a}{1 - a} (ab)^m. Under the condition ab > 1 + \frac{3\pi}{2}, the m-th term dominates the low and tail contributions, ensuring the quotient's magnitude exceeds a positive constant times (ab)^m (which tends to infinity as m → ∞). By choosing phases to alternate the sign of the dominant term (possible due to b odd), the quotient oscillates between large positive and negative values, ensuring no limit exists. The parameter condition ab > 1 + \frac{3\pi}{2} is crucial, as it guarantees the dominant m-th term overwhelms both the partial sum (controlled by ab > 1) and the tail (where the \frac{3\pi}{2} factor arises from tightening the bound via refined estimates in the geometric tail sum). This condition ensures the remainder's contribution is less than half the main term's , preserving the large . Intuitively, the nowhere differentiability stems from the self-similar structure of the series: each successive term introduces wiggles of scaled by a^n but scaled by b^n, resulting in slopes steepening by factor ab > 1. As one zooms in toward any point, finer-scale iterations reveal increasingly jagged features without smoothing to a line, akin to a boundary that defies local linearity. In 1916, G. H. Hardy simplified the original proof by employing integration techniques to bound the difference quotient more directly, relaxing the condition to ab > 1 while confirming nowhere differentiability. His approach integrates over small intervals to estimate variations, avoiding phase alignments and providing a cleaner demonstration of oscillation through average slopes that fail to settle.

Hölder Continuity

The Weierstrass function w(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x), with parameters satisfying $0 < a < 1 and ab > 1, exhibits Hölder continuity with exponent \alpha = -\frac{\log a}{\log b}, where $0 < \alpha < 1. This means there exists a constant C > 0 such that |w(x) - w(y)| \leq C |x - y|^\alpha for all x, y in the domain. The exponent \alpha arises naturally from the scaling in the series, balancing the contraction factor a against the frequency growth b. To establish this bound, consider the difference |w(x) - w(y)| = \left| \sum_{n=0}^\infty a^n [\cos(b^n \pi x) - \cos(b^n \pi y)] \right| \leq \sum_{n=0}^\infty a^n |\cos(b^n \pi x) - \cos(b^n \pi y)|. Using the trigonometric identity, |\cos u - \cos v| = 2 \left| \sin \frac{u+v}{2} \sin \frac{u-v}{2} \right| \leq |u - v| since |\sin \theta| \leq |\theta| for real \theta, it follows that |\cos(b^n \pi x) - \cos(b^n \pi y)| \leq b^n \pi |x - y|. However, for the Hölder estimate, a refined bound is applied: for small arguments, |\cos u - \cos v| \leq 2 (b^n |x - y| \pi / 2)^\alpha, leveraging the fact that |\sin \theta| \leq \theta^\alpha adjusted for the exponent. Splitting the sum at the term where b^n |x - y| \approx 1, the partial sums form with common ratio a b^\alpha = 1, but strict inequality ensures convergence to yield the desired C |x - y|^\alpha. The function is not Lipschitz continuous, which would require \alpha = 1, because the condition ab > 1 implies a > b^{-1}, so -\log a < \log b and thus \alpha < 1. This sub-Lipschitz behavior reflects a "rougher" form of continuity, where the function's oscillations prevent a uniform linear bound on increments. The Hölder exponent \alpha < 1 contributes to the fractal-like appearance of the graph, as the self-similar structure at finer scales amplifies variations in a non-smooth manner. This property also underpins the nowhere differentiability of the function.

Advanced Characteristics

Fractal Dimension

The graph of the Weierstrass function, defined as w(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x) with parameters $0 < a < 1 and b > 1 such that ab > 1, possesses a box-counting d = 2 - \alpha = 2 + \frac{\log a}{\log b}, where \alpha = -\frac{\log a}{\log b} is the Hölder exponent ensuring $1 < d < 2. This quantifies the fractal complexity of the curve, reflecting its intricate structure at all scales without being space-filling. The value of d arises from the self-similar nature of the graph, where each iteration of the series introduces scaled copies: the x-coordinate scales by a factor of $1/b, while the y-coordinate scales by a, yielding the similarity dimension formula d = 2 + \frac{\log a}{\log b}, which equals the for this construction. This approach leverages the iterative refinement in the infinite series to estimate the number of boxes needed to cover the graph at successively finer resolutions. In 2025 review articles, researchers have consolidated results showing that the Hausdorff dimension matches the box-counting dimension $2 - \alpha for Weierstrass-type functions under specific conditions, such as when the phases in generalized cosine terms are chosen randomly with probability 1. These developments extend classical bounds and confirm the fractal dimensionality across parameter regimes where convergence holds. Visually, the graph's fractal character manifests in its infinite arc length—due to the dimension exceeding 1—contrasted with the finite area it encloses over a compact interval, as the function remains bounded and continuous. This duality underscores the function's role as a paradigmatic example in .

Absence of Bounded Variation

A function f: [a, b] \to \mathbb{R} is of bounded variation if its total variation V_a^b(f) = \sup\left\{ \sum_{i=0}^{n-1} |f(x_{i+1}) - f(x_i)| \right\}, where the supremum is taken over all finite partitions a = x_0 < x_1 < \dots < x_n = b, is finite. The Weierstrass function f(x) = \sum_{n=0}^\infty a^n \cos(b^n \pi x), with $0 < a < 1, integer b \ge 3, and ab > 1, is not of bounded variation on [0,1]. To establish this, consider the partial sums s_N(x) = \sum_{n=0}^N a^n \cos(b^n \pi x). The total variation of f is infinite if we can find partitions where the sum of absolute differences grows without bound. For a suitable partition aligned with the frequencies of the terms, the variation contributed by the n-th term is at least c a^n b^n for some constant c > 0, since the term a^n \cos(b^n \pi x) completes approximately b^n / 2 full oscillations in [0,1], each contributing a variation of roughly $2 a^n. Thus, a lower bound for the variation on such partitions is proportional to \sum_{n=0}^N a^n b^n, which diverges as N \to \infty because ab > 1. This infinite total variation implies that the graph of the Weierstrass function over [0,1] is non-rectifiable, meaning it has infinite arc length. This property aligns with the fractal dimension of the graph exceeding 1. In contrast, differentiable functions on compact intervals with integrable derivatives are of bounded variation, as their total variation equals the integral of the absolute value of the derivative; the Weierstrass function's failure in this regard underscores its nowhere differentiability.

Riemann's Variant

Bernhard Riemann proposed an early example of a continuous but nowhere during his lectures in 1861, predating Karl Weierstrass's published construction by over a decade. This unpublished example, drawn from Riemann's notes on , takes the form f(x) = \sum_{n=1}^{\infty} \frac{\sin(n^2 x)}{n^2}, which Riemann asserted was continuous everywhere yet lacked a at any point, though he offered no formal proof and relied on insights from theory. The function's series converges uniformly due to the $1/n^2 coefficients, ensuring on \mathbb{R}, but its quadratic phase terms n^2 x introduce rapid oscillations that Riemann believed prevented differentiability everywhere. Unlike Weierstrass's later function, which incorporates adjustable parameters a and b (with $0 < a < 1 and ab > 1 + 3\pi/2) to guarantee differentiability, Riemann's version employs fixed frequencies without such tunable elements for precise control over the . Riemann's example remained obscure and unpublished until after his death in 1866, but Weierstrass was aware of it through his pupils and mentioned it in his 1872 lecture; both targeted the same counterintuitive phenomenon of continuous functions defying differentiability. Modern scrutiny confirms Riemann's function as Hölder continuous with exponent $1/2, meaning |f(x) - f(y)| \leq C |x - y|^{1/2} for some constant C > 0, but it falls short of being differentiable, as it possesses a (equal to -1/2) at certain rational multiples of \pi. Subsequent mathematicians provided the rigorous analysis absent in Riemann's lifetime, establishing the function's continuity and partial differentiability properties. In 1916, G. H. Hardy showed that the function fails to be differentiable at irrational multiples of \pi. Joseph Gerver demonstrated in 1970 that it is differentiable precisely at certain rational multiples of \pi, with derivative equal to -1/2 there, revealing it as almost nowhere differentiable rather than strictly nowhere so. This work positioned Riemann's construction as a seminal precursor, highlighting early 19th-century intuitions about pathological functions despite incomplete proofs at the time.

Density in Function Spaces

A fundamental result in establishes that the set of nowhere-differentiable continuous functions on [0,1] is comeager, or residual, in the space C[0,1] equipped with the uniform topology. This means that the set of functions differentiable at at least one point is meager, consisting of a countable union of closed nowhere dense sets, as demonstrated using the . The theorem implies not only the existence of such pathological functions but also their ubiquity in the topological sense within the space of all continuous functions. The proof leverages the by partitioning C[0,1] into sets where functions fail to be differentiable at specific points. For each point x \in [0,1] and rational \lambda, the set of functions differentiable at x with derivative \lambda is closed and has empty interior, hence nowhere dense; the union over all such sets is meager. To explicitly construct approximations, any can be perturbed by adding a Weierstrass-like series with sufficiently small amplitude, preserving while ensuring the perturbation destroys potential differentiability at every point without introducing discontinuities. Weierstrass's original example plays a pivotal role in these constructions, providing the foundational nowhere-differentiable building block that enables the perturbation technique to approximate arbitrary continuous functions while maintaining the pathological property across the entire space in the Baire category sense. Extensions of this density result appear in more refined function spaces, such as the Hölder spaces C^{0,\alpha}[0,1] for $0 < \alpha < 1, where the set of nowhere-differentiable functions remains comeager under the respective norms. Similar genericity holds for subclasses with additional constraints, including strictly increasing (monotonic) functions, where constructions analogous to Weierstrass's yield dense sets of nowhere-differentiable monotone continuous functions.

Applications and Extensions

In Optimization Problems

The Weierstrass function is widely employed as a benchmark test function in global optimization problems owing to its highly multimodal nature, featuring infinitely many local minima in the continuous infinite series formulation, which poses significant challenges for algorithms seeking the global minimum. This property makes it an ideal candidate for assessing the performance of population-based metaheuristic methods, such as and , where the ability to explore rugged landscapes and avoid premature convergence to suboptimal solutions is critical. A common finite-sum approximation used in these benchmarks is given by W(x) = \sum_{k=0}^{m} a^k \cos(2\pi b^k x), where $0 < a < 1, b > 1 is typically an , and m is chosen large enough to approximate the infinite series while maintaining computational feasibility; parameters are often set as a = 0.5 and b = 3 to ensure the function's pathological , such as local extrema within any . The version underscores the optimization difficulties by exhibiting fractal-like oscillations that intensify with higher frequencies, complicating for standard local search techniques. Given its nowhere differentiability, the Weierstrass function is particularly valuable for evaluating algorithms, which must rely on function evaluations alone without information. Implementations of this test function are available in suites for languages like , compatible with SciPy's optimization ecosystem, and in through user-contributed toolboxes for testing. Its non-convexity and self-similar structure further exacerbate the risk of entrapment in local minima, serving as a rigorous evaluator for algorithms' global search capabilities in high-dimensional settings.

In Fractal and Geometric Analysis

The Weierstrass function serves as a foundational model for curves due to its self-similar structure and pathological smoothness properties, exhibiting fractal-like irregularity that mimics natural jagged forms. In higher dimensions, products or iterative constructions of generate graphs with space-filling-like behavior, where the approaches the ambient space , enabling the study of complex geometric objects beyond simple shapes. Weierstrass-type functions act as deterministic analogues to sample paths, sharing comparable Hölder regularity—specifically, when the scaling parameter corresponds to exponent 1/2, the functions display the same almost-sure Hölder continuity properties as Brownian trajectories, facilitating approximations in rough path theory and stochastic analysis. Recent advancements in 2025 have extended these ideas to generalized within , computing precise Hausdorff and box dimensions for variants and applying them to random fractals, which model irregular sets in probabilistic geometric settings. Extensions to the Weierstrass-Mandelbrot function in two dimensions provide a surface analogue, defined via sums of scaled sines, and have been employed to simulate terrains such as coastlines, capturing their self-affine roughness through parameter-tuned dimensions between 1 and 2. Similarly, random variants of this function model scalar in atmospheric flows, replicating high-Reynolds-number irregularities in carbon dioxide dispersion.

References

  1. [1]
    [PDF] Weierstrass's non-differentiable function
    Dec 12, 2014 · In the nineteenth century, many mathematicians held the belief that a continuous function must be differentiable at a large set of points.<|control11|><|separator|>
  2. [2]
    [PDF] CONTINUOUS, NOWHERE DIFFERENTIABLE FUNCTIONS
    We begin with a proof that Weierstrass' famous nowhere differentiable function is in fact everywhere continuous and nowhere differentiable. We then address the ...Missing: source | Show results with:source
  3. [3]
    The Jagged, Monstrous Function That Broke Calculus
    Jan 23, 2025 · But Weierstrass had proved beyond doubt that, though his function had no discontinuities, it was never differentiable. To show this, he first ...Missing: paper | Show results with:paper<|control11|><|separator|>
  4. [4]
    [PDF] THE WEIERSTRASS PATHOLOGICAL FUNCTION - UCSD Math
    Define the function W : R → R by. W(x) = ∞. X n=0 an cos(bnπx). Then W is uniformly continuous on R, but is differentiable at no point. REMARK 0.2. Notice ...
  5. [5]
    Weierstrass Function - UMD Physics
    Weierstrass Function. The Weierstrass function is defined as: f(x)=∞∑n=1ancos(bnπx)(1) (1) f ( x ) = ∑ n = 1 ∞ a n cos ⁡ ( b n π x ) where 0<a<1 0 < a < 1 and ...
  6. [6]
    weierstrass functions with random phases
    Mar 19, 2003 · where g is a 1-periodic Lipschitz function, 1 < b < +∞ and 0 < α < 1. Such a function will be called a Weierstrass function, and it is easy to ...
  7. [7]
    the hausdorff dimension of graphs of weierstrass functions
    Brief discussions of the problem of the Hausdorff dimension of the graph of the. Weierstrass function, along with additional references, can be found in the ...
  8. [8]
    Properties of the Weierstrass function in the time and frequency ...
    The Weierstrass function is continuous everywhere and differentiable nowhere. It is used advantageously for modeling fractally coarse media.
  9. [9]
    WEIERSTRASS'S NON-DIFFERENTIABLE FUNCTION
    The chief results which I prove here concerning Weierstrass's function, and the corresponding function defined by a series of sines, may be sum- marized as ...
  10. [10]
    [PDF] Continuous Nowhere Differentiable Functions
    Proof. The continuity of C follows exactly like in the proof for Weierstrass' function (Theorem 3.4). That C is nowhere differentiable follows from Hardy.Missing: primary | Show results with:primary
  11. [11]
    Cauchy's Calculus - MacTutor History of Mathematics
    Cauchy wrote Cours d'Analyse (1821) based on his lecture course at the École Polytechnique. In it he attempted to make calculus rigorous.Missing: non- | Show results with:non-
  12. [12]
    [PDF] MATHEMATISCH CENTRUM
    WEIERSTRASS' function is the first example that has been published. I. P. du BOIS-REYMOND [8] describes in 1875 this function which Weierstrass has send to ...Missing: 1821 Poincaré<|control11|><|separator|>
  13. [13]
    [PDF] bilyk-weierstrass.pdf
    Weierstrass function (continuous, nowhere differentiable). A lunar crater and ... , b an odd integer (Weierstrass, 1872) if ab > 1 (du Bois-Reymond, 1875).Missing: history 1834 Cauchy 1821 Poincaré
  14. [14]
    [PDF] Weierstrass and Approximation Theory
    (published in 1895) there appears the paper Weierstrass [1872] wherein Weierstrass proves that the function f(x) = ∞. X n=0 bn cos(anxπ). Page 4. 4. Allan ...
  15. [15]
    The origins of fractals | plus.maths.org
    Sep 1, 1998 · Hermite and his pupil Poincaré in particular described Weierstrass' new creations as "deplorable evil"! By way of an introduction, we start with ...Missing: 1886 | Show results with:1886
  16. [16]
    Weierstrass Function -- from Wolfram MathWorld
    The pathological function f_a(x)=sum_(k=1)^infty(sin(pik^ax))/(pik^a) (originally defined for a=2 ) that is continuous but differentiable only on a set of ...Missing: parameters | Show results with:parameters
  17. [17]
    [PDF] The Takagi function: a survey - arXiv
    Nov 7, 2011 · A variant of Takagi's function, using the base 10, was discovered in 1930 by Van der Waerden, who credits Dr. A.
  18. [18]
    Weierstrass's Non-Differentiable Function - jstor
    (1.21) ab > 1 +- 3( - a). These conditions forbid the existence of a differential coefficient fnite or infinite. For the non-existence of a finite difVerential ...Missing: parameters | Show results with:parameters
  19. [19]
    [PDF] The Weierstrass Function
    The function appearing in the above theorem is called the Weierstrass function. Before we prove the theorem, we require the following lemma: Lemma (The ...<|control11|><|separator|>
  20. [20]
    [PDF] Weierstrass's Non-Differentiable Function Author(s): G. H. Hardy ...
    Weierstrass's Non-Differentiable Function. Author(s): G. H. Hardy. Source: Transactions of the American Mathematical Society, Vol. 17, No. 3, (Jul., 1916), pp.Missing: ab | Show results with:ab
  21. [21]
  22. [22]
    [PDF] Fractals And The Weierstrass-Mandelbrot Function
    Here, we first review some basic ideas from measure theory and fractal geometry, focusing on the Hausdorff, box counting, packing, and similarity dimensions.
  23. [23]
    SOLVING (I S)g = / WHEN S IS A GENERALIZED SHIFT OPERATOR
    Jun 15, 1992 · sin(2πt) yields the Weierstrass function g(t) = И j>¼ aj ... It is well known that g is not of bounded variation on any subinterval in this.
  24. [24]
    [PDF] On Hölder continuity and pth-variation function of Weierstrass-type ...
    Jul 12, 2024 · In this paper, we investigate Hölder continuity (see Definition 1.1), pth-variation function. (see Definition 1.2) and Riesz variation (see ...Missing: "seminal | Show results with:"seminal
  25. [25]
  26. [26]
    Weierstrass Function - an overview | ScienceDirect Topics
    An elliptic function is a meromorphic function on the complex plane with two (independent) periods, ω1 and ω2, that is, f(z + ω1 = f(z), f(z + ω2) = f(z), and ...
  27. [27]
    On Weierstrass' Monsters and lineability - Project Euclid
    There are no infinite dimensional closed subspaces of C[0, 1] composed by just functions of bounded variation. ... Weierstrass function, since 9a > 7 > 1 + 3.
  28. [28]
    Some geometric properties of Riemann's non-differentiable function
    Riemann's non-differentiable function R ( x ) = ∑ n = 1 ∞ sin ⁡ ( n 2 x ) n 2 , x ∈ R , is a classic example of a continuous but almost nowhere differentiable ...
  29. [29]
    [PDF] Riemann's non-differentiable function is intermittent - HAL
    Oct 29, 2019 · sin n2t n2. , t ∈ [0,2π]. is a celebrated example of a continuous but almost nowhere differentiable function. It was introduced by the ...
  30. [30]
    [PDF] Weierstrass' Response to Riemann - arXiv
    His discovery of both continuous nowhere differentiable functions and series having natural boundaries strengthened him in this view. “All difficulties vanish”, ...
  31. [31]
    [PDF] About the Uniform Hölder Continuity of Generalized Riemann Function
    Apr 1, 2014 · Hölder continuous with exponent 1/2 and gave some results about its Hölder continuity at some particular points. With some similar ...
  32. [32]
    THE DIFFERENTIABILITY OF THE RIEMANN FUNCTION AT ...
    It is shown that a continuous function which Riemann is said to have believed to be nowhere differentiable is in fact differentiable at certain points.Missing: 1861 manuscript discovered 1892
  33. [33]
    The differentiability of Riemann's functions - Semantic Scholar
    Feb 1, 1972 · It is shown that a continuous function which Riemann is said to have believed to be nowhere differentiable is in fact differentiable at certain ...Missing: manuscript discovered 1892
  34. [34]
    [PDF] Continuous functions that are nowhere differentiable
    It is shown that the existence of continuous functions on the interval [0,1] that are nowhere differentiable can be deduced from the Baire category theorem.
  35. [35]
    [PDF] Most Continuous Functions are Nowhere Differentiable
    Let PL(K) ⊆ C(K) be the set of piecewise linear continuous functions on K. Theorem 10 PL(K) is dense in C(K). Proof Suppose f ∈ C(K) and ² > 0 ...
  36. [36]
    Proof that the nowhere differentiable functions are dense in $C_b ...
    Jul 16, 2016 · Proof that the nowhere differentiable functions are dense in Cb(R). ... I tried to make a proof, where I use a Weierstrass function. I was ...
  37. [37]
    Topological genericity of nowhere differentiable functions in the disc ...
    The statement of Theorem 2.1 can be essentially strengthened using another result of Hardy. In [9] he treated continuous functions which satisfy no α-Hölder ...<|control11|><|separator|>
  38. [38]
    [PDF] Continuous Maps Admitting No Tangent Lines
    May 27, 2022 · One of the most influential examples in analysis is a Weierstrass function from R to R that is continuous but differentiable at no point.
  39. [39]
    A Novel Flexible Inertia Weight Particle Swarm Optimization Algorithm
    Aug 25, 2016 · This paper proposes a new nonlinear strategy for selecting inertia weight which is named Flexible Exponential Inertia Weight (FEIW) strategy.
  40. [40]
    A Collection of 30 Multidimensional Functions for Global ... - MDPI
    A collection of thirty mathematical functions that can be used for optimization purposes is presented and investigated in detail.<|control11|><|separator|>
  41. [41]
    [PDF] THE FRACTAL DIMENSION OF THE WEIERSTRASS TYPE ... - CORE
    Our aim in the thesis is to estimate the fractal dimension of certain graphs related to the Weierstrass functions. The Weierstrass function is the best-known ...
  42. [42]
    Weierstrass' function and Brownian motion - MathOverflow
    Mar 21, 2011 · Specifically, when α=1/2, the Weierstrass function has same Holder continuity peoperties that Brownain sample paths do. Some crude Matlab ...
  43. [43]
    Progress on Fractal Dimensions of the Weierstrass Function and ...
    The box dimension of Weierstrass-type functions has been proved to be 2 − α . Moreover, in certain conditions, the Packing dimension of Weierstrass-type ...
  44. [44]
    Progress on Fractal Dimensions of the Weierstrass Function and ...
    Feb 25, 2025 · The Weierstrass function W(x)=∑n=1∞ancos(2πbnx) is a function that is continuous everywhere and differentiable nowhere.
  45. [45]
    On the Weierstrass-Mandelbrot fractal function - Journals
    The properties of IF are illustrated by computed graphs for several values of D (including some 'marginal' cases = 1 where the series for W converges) and ...
  46. [46]