Fact-checked by Grok 2 weeks ago

Almost all

In mathematics, the phrase "almost all" describes a property that holds for every element of a set except those in a negligible subset, where negligibility is determined by the ambient structure, such as a measure or density. In measure theory, a property holds almost everywhere (often synonymous with "almost all") with respect to a measure \mu on a space (\Omega, \Sigma, \mu) if the set N \subseteq \Omega where the property fails satisfies \mu(N) = 0; two functions are then considered equivalent if they agree almost everywhere. This concept is fundamental to integration and convergence theorems, allowing mathematicians to disregard exceptional sets of measure zero without affecting results, as seen in the Lebesgue integral where functions differing on null sets yield the same integral value. In and , "almost all" natural numbers n satisfy a property if the exceptional set has zero, defined as \lim_{N \to \infty} \frac{|S \cap [1,N]|}{N} = 0 for the subset S of exceptions, implying the proportion approaches zero asymptotically. For instance, almost all positive integers are composite, since the primes have zero by the , which states that the number of primes up to N is asymptotically \frac{N}{\log N}. This usage extends to probabilistic statements about integers, where "almost all" captures behaviors typical for large n, excluding rare outliers like powers or highly composite numbers. The term also appears in other areas, such as and random graph theory, where it aligns with probability measures approaching 1, but always emphasizes the irrelevance of the exceptional set for the 's validity. These interpretations ensure precision in proofs involving infinite sets, avoiding absolute universality while maintaining rigor.

Conceptual Foundations

Prevalent Meaning

In , the phrase "almost all" informally describes a that holds for every of a set except those belonging to a negligible or exceptional subset, where negligibility is interpreted relative to the ambient and the chosen notion of size. This allows for exceptions while emphasizing that they are insignificant in the broader context, such as finite subsets of infinite sets or subsets with zero asymptotic density. The term is prevalent in everyday mathematical to convey that the exceptional cases do not affect the general behavior or conclusions drawn about the set. A classic example is the statement that almost all real numbers are : the rational numbers, being countable, form a negligible subset of the uncountable real line, leaving the irrationals as the overwhelming majority in terms of . Similarly, in the natural numbers, almost all positive integers are composite, as the primes, though , constitute a set of zero and thus represent an exceptional minority. This usage distinguishes "almost all" from "all," which permits no exceptions whatsoever, and from "most," which merely implies a simple majority without regard to the scale of exceptions. In asymptotic analysis and limiting processes, "almost all" underscores typical behaviors that emerge as parameters tend to infinity, prioritizing structural dominance over isolated anomalies. This intuitive sense paves the way for rigorous formalizations, such as in measure theory where exceptions form a set of measure zero.

Historical Development

The concept of "almost all," rooted in early 20th-century , derives from the notion of "" introduced by in his 1902 doctoral thesis Intégrale, longueur, aire, where properties holding except on sets of measure zero were formalized using the French term presque partout. This framework emerged from efforts to extend integration beyond Riemann's method, emphasizing negligible exceptions in continuous settings. Early applications appeared in David Hilbert's 1900 address on unsolved problems, particularly the seventh problem concerning , which implicitly relied on density arguments akin to later measure-theoretic ideas for "most" real numbers, though without explicit terminology. By 1935, Antoni Zygmund's Trigonometric Series marked one of the first major English-language texts employing "almost everywhere" convergence for , building directly on Lebesgue's foundations to analyze behavior except on null sets. The term expanded into during the mid-20th century through probabilistic , with and Mark Kac's 1940 paper "The Gaussian Law of Errors in the Theory of Additive Number Theoretic Functions" popularizing "almost all" integers—meaning all but a proportion approaching zero—as a tool for studying additive functions like the number of prime factors. This shift adapted measure-theoretic notions to asymptotic densities in integers, influencing results on and distribution laws. From the 1960s onward, "almost all" standardized in and random structures, drawing on Andrey Kolmogorov's 1933 axiomatization of probability—which underpinned measure in dynamical systems—and Norbert Wiener's 1939 proving for processes. These works facilitated applications in theory, such as the (1959–1960), where properties hold for graphs on n vertices as n grows. The terminology evolved from presque partout in texts to "almost everywhere/all" in English by the 1930s, reflecting translations and adoption in American and British literature like Zygmund's.

In Continuous Mathematics

Measure Theory Definition

In measure theory, a property holds for almost all elements of a (X, \mathcal{A}, \mu) if the set of elements where the property fails is a , meaning it belongs to the \sigma-algebra \mathcal{A} and has measure zero, \mu(E) = 0. A consists of a set X, a \sigma-algebra \mathcal{A} of subsets of X (closed under countable unions, intersections, and complements), and a measure \mu: \mathcal{A} \to [0, \infty] that is countably additive on disjoint measurable sets. The Lebesgue measure on \mathbb{R}^n, for instance, is a specific complete measure on the \sigma-algebra of Lebesgue measurable sets, assigning to each Borel set its intuitive volume while extending to a larger class via Carathéodory's criterion. To handle null sets more flexibly, the completion of a measure space (X, \mathcal{A}, \mu) is often considered, yielding a larger \sigma-algebra \overline{\mathcal{A}} that includes all subsets of null sets from \mathcal{A}, with the extended measure \overline{\mu} defined such that \overline{\mu}(A \cup N) = \mu(A) for measurable A and null N. In this completed space, which is complete by construction (every subset of a null set is measurable and null), the notion of "almost all" equates two functions or properties if they differ only on a null set, forming equivalence classes modulo null sets. This invariance under measure-zero modifications ensures that statements about almost all elements are robust to negligible perturbations. A key property in , which builds on measure theory with \mu a , is : a sequence of random variables X_n converges to X if \mu(\{\omega : \lim_{n \to \infty} X_n(\omega) \neq X(\omega)\}) = 0. This is a direct application of the almost all framework, where the exceptional set has probability zero. A fundamental result illustrating the power of this concept is Lebesgue's differentiation theorem, which asserts that for a locally integrable f on \mathbb{R}^n with respect to \mu, the limit \lim_{r \to 0} \frac{1}{\mu(B(x,r))} \int_{B(x,r)} f(y) \, d\mu(y) = f(x) holds for x \in \mathbb{R}^n, where B(x,r) is the of radius r centered at x. Equivalently, the set where the average deviation fails to vanish, \left\{ x : \lim_{r \to 0} \frac{1}{\mu(B(x,r))} \int_{B(x,r)} |f(y) - f(x)| \, d\mu(y) \neq 0 \right\}, has measure zero, underscoring how recovers the function value despite potential singularities on null sets.

Topological Variants

In topological spaces, the notion of "almost all" points satisfying a property is defined in terms of Baire category, where the exceptional set is meager, meaning it can be expressed as a countable of nowhere dense subsets. A subset is nowhere dense if the interior of its is empty, ensuring that such sets lack in any open region. This categorical approach provides a qualitative measure of "largeness" that is independent of any or measure structure, making it applicable to a wide range of topological settings, including non- spaces. The foundational result underpinning this perspective is the , which states that in a , the intersection of countably many dense s is itself dense. Consequently, a comeager set—whose complement is meager—is not only dense but also "prevalent" in the topological sense, as it intersects every non-empty in a substantial way. This theorem, originally established by René Baire in his 1899 doctoral thesis, highlights how comeager sets represent the generic case in complete spaces, contrasting with the quantitative "" from measure theory by emphasizing structural density over size. A classic example occurs in the real line \mathbb{R} with the standard topology, where the rational numbers \mathbb{Q} form a meager set as a countable union of singletons, each of which is nowhere dense. Their complement, the irrationals \mathbb{R} \setminus \mathbb{Q}, is therefore comeager, expressed as the countable intersection of dense open sets \mathbb{R} \setminus \{q_n\}, where \{q_n\} enumerates \mathbb{Q}. In dynamical systems, this framework identifies generic properties, such as the prevalence of minimal actions or scrambled sets under homeomorphisms of the , where comeager subsets of the space of transformations exhibit or transitive behavior for almost all initial points. Related concepts appear in fractal geometry through porous sets, where a set is porous at a point if there exist arbitrarily small balls centered nearby that are mostly disjoint from the set, quantifying "holes" in a topological manner. A \sigma-porous set, as a countable of porous sets, extends the meager idea to irregular structures, allowing "almost all" points (in the comeager sense) to exhibit positive , which bounds dimensions and aids in analyzing self-similar fractals like boundaries of Mandelbrot sets. This topological variant of "almost all" is orthogonal to measure-theoretic notions, as meager sets can carry full Lebesgue measure; for instance, the Smith-Volterra-Cantor set is a closed nowhere dense subset of [0,1] with measure $1/2, serving as a basic meager set of positive measure that underscores the independence of category and measure. Measure zero sets, by contrast, represent smallness in a probabilistic sense but may be comeager, illustrating how the two frameworks complement rather than overlap in assessing prevalence.

In Discrete Mathematics

Number Theory Applications

In , the concept of "almost all" natural numbers is typically formalized using . A E \subseteq \mathbb{N} has natural density zero if \lim_{n \to \infty} \frac{|E \cap [1,n]|}{n} = 0; thus, a property holds for almost all natural numbers if the exceptional set has natural density zero. This notion captures asymptotic behaviors where exceptions become negligible relative to the total count up to n. The provides a foundational example: it states that the number of primes up to x is asymptotically \sim \frac{x}{\log x}, implying that the primes have zero and hence almost all positive integers are composite. A more refined result is the , which describes the distribution of the number of distinct prime factors \omega(n). Specifically, for almost all n \leq x as x \to \infty, the normalized quantity \frac{\omega(n) - \log \log n}{\sqrt{\log \log n}} converges in distribution to the N(0,1). This theorem highlights that \omega(n) fluctuates around its mean \log \log n with standard deviation \sqrt{\log \log n} for almost all n. Variants of are crucial for progressions. The Dirichlet density of a set of primes E is defined as \lim_{s \to 1^+} \frac{\sum_{p \in E} p^{-s}}{-\log(s-1)} (if the limit exists), which equals density when both are defined. The logarithmic density, given by \lim_{n \to \infty} \frac{1}{\log n} \sum_{k=1}^n \frac{1_{E}(k)}{k} = 0 for the exceptional set, often coincides with the Dirichlet density and is particularly useful for weighted counts in progressions. These measures ensure that primes are equidistributed among coprime residue classes modulo q, with density $1/\phi(q) in each class, as per Dirichlet's theorem. Applications extend to the typical prime factorization: almost all integers n have approximately \log \log n distinct prime factors, a result building on the and earlier work showing this as the normal order. In sieve theory, such asymptotic densities help quantify exceptions; for instance, sieves bound the size of sets avoiding certain prime factors, revealing that while almost all integers satisfy typical factorization properties, exceptional sets (like those with unusually few or many factors) can be controlled to have density zero.

Graph Theory Contexts

In , the notion of "almost all" graphs possessing a certain is typically analyzed within the framework of random graph models, particularly the Erdős–Rényi binomial model G(n, p), where n vertices are fixed and each possible is included independently with probability p = p(n). A holds for almost all graphs if, in the limit as n \to \infty, the probability that a random G(n, p) satisfies the approaches 1 (denoted "" or "with high probability"). This probabilistic perspective allows for the study of asymptotic behavior in sequences of graphs G_n, where structural features emerge reliably above specific thresholds in p. The binomial random graph G(n, p) evolves through distinct phases as p increases from 0 to 1, reflecting the growth and merging of connected components. Initially, for p = o(1/n), the graph consists almost surely of isolated vertices and small tree components, with no cycles. As p reaches $1/n, a phase transition occurs: a giant component of size \Theta(n) emerges almost surely when the expected degree np > 1, while smaller components remain tree-like or contain few cycles. For p = c/n with c > 1, the giant component dominates, encompassing a positive fraction of vertices, and its size is given asymptotically by solving $1 - u = e^{c(u-1)} where u is the survival probability in a branching process approximation. Further increase to p \approx (\ln n)/n leads to connectivity, with all components merging into one. This evolution models phenomena like network formation and percolation. A canonical example of threshold phenomena is graph connectivity. The threshold for G(n, p) to be connected almost surely is p = (\ln n + c)/n for constant c \in \mathbb{R}, where the probability tends to e^{-e^{-c}} as n \to \infty. Below this threshold (e.g., p = (\ln n - \omega(1))/n), isolated vertices or small components persist almost surely, disconnecting the graph. For Hamiltonicity, Bollobás established a sharp hitting-time result in the random graph process: almost surely, the moment the minimum degree reaches 2 coincides with the emergence of a Hamiltonian cycle. In the G(n, p) model, this yields a threshold of p = (\ln n + \ln \ln n + \omega(1))/n, above which G(n, p) is almost surely Hamiltonian, while below it, the graph almost surely contains vertices of degree less than 2, precluding cycles through all vertices. In extremal graph theory, extensions of Turán's theorem apply "almost all" to random settings, characterizing subgraphs that avoid forbidden structures. For q \gg (\log n)^4 / n)^{1/(l-1)} and \delta > 0 with $1/(l-1) > \delta, almost every G \in G(n, q) ensures that any subgraph F with at least (1 - 1/(l-1) + \delta) q \binom{n}{2} edges contains a clique K_l, and in fact at least c q \binom{l}{2} n^l copies of K_l for some c = c(\delta, l) > 0. This random analog bounds the edge density in K_l-free subgraphs of random graphs, mirroring the deterministic Turán graph T(n, l-1) but with high probability over the host graph. Generalizations extend to d-degenerate forbidden graphs H with chromatic number \chi(H), where dense subgraphs almost surely contain H.

Advanced and Algebraic Settings

Algebraic Interpretations

In , the notion of "almost all" elements satisfying a in a group or over an infinite often means that the exceptional set forms a proper or contained in a subvariety of lower . For instance, in an algebraic group defined over an , properties hold for generic elements outside a Zariski-closed of positive . A central concept is that of generic points in irreducible algebraic varieties. In the spectrum of a ring or an , the corresponds to the (0) in an , representing the entire irreducible component, such that properties true at this point hold at a of points avoiding lower-dimensional loci. Hilbert's irreducibility theorem provides a key result in this context: if f(x, t) \in k[x, t] is irreducible over the function field k(t), where k is an , then for almost all a \in k (in the sense of all but finitely many or a thin set), the f(x, a) remains irreducible over k, and the fiber \{x \mid f(x, a) = 0\} has \deg_x f. f(x, t) \in k[x, t], \quad \text{irreducible over } k(t) \implies \deg_x (f(x, a) = 0) = \deg_x f \quad \text{for almost all } a \in k. This theorem implies that almost all specializations preserve irreducibility and related geometric properties, with applications in number theory and Diophantine geometry. In linear algebra over infinite fields, these ideas manifest concretely. Over an infinite field k, almost all n \times n matrices in M_n(k) are invertible, as the determinant is a nonzero polynomial, and its zero set is a hypersurface of codimension 1, hence a proper subvariety. Similarly, in the general linear group GL(n, \mathbb{C}), almost all matrices are diagonalizable, since the set of nondiagonalizable matrices lies in the zero locus of the discriminant of the characteristic polynomial, a hypersurface excluding a dense open set. The underpins these interpretations, where closed sets are algebraic defined by polynomial ideals, and open sets are complements thereof. Dense open sets in this topology correspond to "almost all" points, as any nonempty Zariski-open subset of an irreducible is dense, avoiding only lower-dimensional exceptional hypersurfaces. Thus, properties failing on a Zariski-closed set of positive hold generically throughout the space.

Combinatorial Extensions

In combinatorial contexts involving finite sets, the concept of "almost all" is typically interpreted through asymptotic density. Specifically, for a sequence of finite ground sets = \{1, 2, \dots, n\}, a property holds for almost all subsets of $$ if the proportion of such subsets satisfying the property tends to 1 as n \to \infty. This notion extends to elements within these sets or to more complex structures like families of subsets, where the limiting proportion (or density) approaching 1 quantifies the prevalence of the property. A cornerstone result in this area is , which asserts that any A \subset with positive upper \delta > 0 (meaning |A \cap [1,n]| / n \geq \delta for infinitely many n) contains arithmetic progressions of arbitrary length k. In the finite setting, this implies that almost all subsets of $$ with fixed \delta > 0 contain k-term arithmetic progressions (k-APs) for any fixed k, as the exceptional subsets avoiding them have tending to 0. More quantitatively, the theorem guarantees a lower bound on the number of such progressions: \left| \{ (a, d) \in \times \mathbb{N} : a + id \in A \ \forall i=0,\dots,k-1, \ d \geq 1, \ a+(k-1)d \leq n \} \right| \geq c(\delta, k) n^2 for some constant c(\delta, k) > 0 depending only on \delta and k, where the pairs (a,d) parameterize the k-APs. This framework finds applications in , where almost all r-colorings of the edges of the K_n contain monochromatic copies of any fixed graph H, provided n is sufficiently large relative to the Ramsey number R(H; r). For instance, in 2-colorings, the proportion of colorings avoiding a monochromatic tends to 0 exponentially fast as n \to \infty, ensuring that nearly all such colorings exhibit the desired monochromatic structures. In extremal , analogous results show that almost all k-uniform intersecting families (those where every pair of sets intersects) are trivial, consisting of all k-subsets containing a fixed element, highlighting the structural rigidity of such families under the "almost all" regime. Extensions to hypergraphs further illustrate the breadth of these ideas. In r-uniform hypergraphs on n vertices, almost all such hypergraphs (in the random model where each possible edge is included independently with probability p > 0 fixed) contain every fixed r-uniform subhypergraph as an induced or embedded structure, once n exceeds certain thresholds derived from extremal hypergraph theory. These generalizations underscore how "almost all" properties permeate arbitrary finite structures, linking density-based arguments across combinatorial domains.

Proof Techniques

General Strategies

The , pioneered by , employs random constructions to prove the existence of combinatorial structures or properties holding for almost all objects by demonstrating that the probability of failure tends to zero. By modeling the underlying space as a and computing the expectation of indicator variables for undesirable events, one can apply to show that the expected proportion of exceptions is small, implying the property holds with high probability. For stronger "almost surely" results, variance calculations paired with establish concentration, ensuring deviations from the mean are negligible with overwhelming probability. This approach, originating in Erdős's work on , revolutionized proofs in by bypassing explicit constructions. Analytic techniques leverage Tauberian theorems to infer asymptotic densities from the analytic properties of associated generating functions, particularly in additive and multiplicative number theory. These theorems, such as the Wiener-Ikehara variant, translate singularities or growth rates of Dirichlet series into precise estimates for the summatory functions of sequences, thereby quantifying the density of sets where certain arithmetic properties fail to hold. Generating functions further facilitate this by encoding enumerative data, with asymptotic expansions near dominant singularities yielding the proportion of elements satisfying recursive or structural conditions, applicable to almost all terms in sequences like partitions or permutations. Sieve methods, exemplified by Brun's sieve, utilize finite approximations to the inclusion-exclusion principle to bound the of exceptional sets defined by avoidance of primes or other sieving conditions. Introduced by Viggo Brun around , this combinatorial tool truncates the alternating sum over divisors up to a that balances accuracy and , providing upper bounds on the of integers not divisible by small primes while estimating those divisible by larger ones. In number-theoretic contexts, it effectively demonstrates that almost all integers exhibit typical prime factor distributions by controlling the size of sifted outliers. Applications from , notably Birkhoff's pointwise ergodic theorem of , establish that time averages of integrable functions under an ergodic measure-preserving transformation converge to the corresponding space average with respect to the invariant measure. This result underpins proofs of generic behavior in dynamical systems, where exceptions form a set of measure zero, and extends to infinite measure spaces under additional conditions like conservativity. It provides a foundational tool for showing that almost all points in a follow the equilibrium distribution. Proving "almost all" statements requires vigilance against pitfalls such as non-uniformity in limiting approximations, where error terms varying with parameters can inflate exceptional sets beyond negligible size. In probabilistic settings, overlooking dependencies among random events often leads to flawed independence assumptions and inaccurate tail bounds; dependency graphs or the mitigate this by partitioning variables into limited-interaction clusters.

Illustrative Examples

One illustrative example of the "almost all" concept arises in number theory, where the prime number theorem implies that the set of prime numbers has asymptotic density zero. The prime number theorem states that the number of primes up to n, denoted \pi(n), satisfies \pi(n) \sim n / \log n as n \to \infty. Thus, the density \pi(n)/n \sim 1 / \log n \to 0, so the exceptional set of primes has density zero, meaning almost all positive integers are composite. In theory, the Pósa rotation-extension technique demonstrates that almost all graphs in the G(n, p) with p = (\log n + \log \log n + \omega(1))/n are , where \omega(1) \to \infty arbitrarily slowly. A visits each vertex exactly once and returns to the start. Pósa's method constructs such cycles by iteratively extending paths and rotating endpoints to increase length, leveraging the expander properties of dense s. Komlós and Szemerédi showed that the for Hamiltonicity is around p \sim (\log n + \log \log n)/n, and for the stated p, the probability of Hamiltonicity tends to 1 as n \to \infty. This implies that (with probability approaching 1), a at this edge probability contains a . A foundational application of "almost all" in probability uses the for events. Consider a of events \{A_k\}_{k=1}^\infty on a (\Omega, \mathcal{F}, P). The first states that if \sum_{k=1}^\infty P(A_k) < \infty, then P(\limsup_{k \to \infty} A_k) = 0, where \limsup A_k = \bigcap_{n=1}^\infty \bigcup_{k=n}^\infty A_k is the set of outcomes where infinitely many A_k occur. In equation form: P\left( \limsup_{k \to \infty} A_k \right) = 0 if \sum P(A_k) < \infty. Consequently, (with probability 1), only finitely many A_k occur, meaning almost all \omega \in \Omega belong to only finitely many A_k and avoid infinitely many. This underpins proofs where rare events fail to accumulate, such as in the or recurrence in dynamical systems.

References

  1. [1]
    [PDF] Almost Everywhere - metaphor
    Definition 2. We say that a property holds µ-almost everywhere if it holds on a set. Nc, where µ(N) = 0.
  2. [2]
    [PDF] Estimating the Density of the Abundant Numbers
    Feb 1, 2019 · Definition 1. Given a set S of positive integers, let S(n) be the number of integers x ≤ n with x ∈ S. We define the natural density of S, dS, ...
  3. [3]
    [PDF] Almost all reductions of an elliptic curve have a large exponent
    This shows that, for almost all p, the group of Fp-points of every elliptic curve defined over Fp contains points of order at least p3/4/ log p. 3. Page 4. This ...
  4. [4]
    Negligible Set - an overview | ScienceDirect Topics
    A negligible set is defined as a set that is insignificant in measure or ... almost all points of E. Remarks. (i). In the original definition of ...
  5. [5]
    Almost All -- from Wolfram MathWorld
    P is said to hold true for almost all numbers. For example, almost all positive integers are composite numbers.
  6. [6]
    [PDF] niven.pdf
    This is a spe- cial case of the following result. THEOREM 1.1. Almost all real numbers are irrational. Proof. First we prove that the positive rational num ...
  7. [7]
    Trigonometric Series - Cambridge University Press & Assessment
    Professor Zygmund's Trigonometric Series, first published in Warsaw in 1935, established itself as a classic. It presented a concise account of the main ...
  8. [8]
    [PDF] Norbert Wiener and Probability Theory - Indian Academy of Sciences
    Among the topics we talked about were mostly his own past work in ergodic theory, prediction theory and generalized harmonic analysis. We also worked together ( ...
  9. [9]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    A property which holds for all x ∈ X \ N where N is a set of measure zero is said to hold almost everywhere, or a.e. for short. If we want to emphasize the ...
  10. [10]
    [PDF] 2.2 Measures - Christopher Heil
    Lebesgue measure is a σ-finite measure on Rd that is not a bounded mea- sure (because |Rd| = ∞). When we deal with Lebesgue measure in this vol- ume, we will ...
  11. [11]
    [PDF] 2.4 The Completion of a Measure - Christopher Heil
    A complete measure is one such that every subset A of every null set E is measurable (Definition 2.19). Complete measures are often more convenient to work with ...
  12. [12]
    [PDF] 2.3 Basic Properties of Measures - Christopher Heil
    Given a measure space (X, Σ,µ), a set E ∈ Σ such that µ(E) = 0 is called a µ-null set or a set of µ-measure zero. A property that holds for all x ∈ X except ...
  13. [13]
    [PDF] Differentiation - UC Davis Math
    The Lebesgue differentiation theorem states that (6.5) holds pointwise µ-a.e. for any locally inte- grable function f. To prove the theorem, we will introduce ...
  14. [14]
    Applications of the Baire Category Theorem - Project Euclid
    In his doctoral thesis in 1899, René Baire writes, “The continuum constitutes a set of the second category.” This is one of the the statements for which ...Missing: original | Show results with:original
  15. [15]
    [PDF] The Baire category theorem
    ... René-Louis Baire (1874–1932) in his 1899 doctoral thesis. Baire made a number of important contributions to real analysis in addition to the category theorem.Missing: paper | Show results with:paper
  16. [16]
    [PDF] dense and comeager conjugacy classes in zero-dimensional ...
    Conjugacy classes are orbits of Homeo(o) acting on actions of a group G. A comeager class means generic behavior is determined by a single system.
  17. [17]
    [PDF] POROSITY AND DIMENSION OF SETS AND MEASURES
    Aug 21, 2009 · almost all x ∈ A there are no approximative tangent m-planes for A at x. Moreover, it is known that if we have a purely m-unrectifiable set A ⊂ ...
  18. [18]
    [PDF] The algebraic difference of a Cantor set and its complement - arXiv
    Recall that the classical Cantor ternary C Ç [0, 1] is “small” in both the sense of Baire category and Lebesgue measure, that is, it is meager and has measure.
  19. [19]
    [PDF] Introduction to the Theory of Numbers
    Nov 21, 2014 · THIS book has developed gradually from lectures delivered in a number of universities during the last ten years, and, like many books which have ...
  20. [20]
    [PDF] Math 676. Dirichlet density for global fields
    The main difference between the two concepts is that natural density rests on an ordering of the set of primes, whereas Dirichlet density does not. For a ...
  21. [21]
    [PDF] On the notions of upper and lower density - arXiv
    Abstract. Let P(N) be the power set of N. We say that a function µ* : P(N) → R is an upper density if, for all X, Y ⊆ N and h, k ∈ N+, the following hold: ...
  22. [22]
    [PDF] Dirichlet characters and L-functions Dirichl
    We can use the notion of logarithmic density to state Dirichlet's theorem as follows: Theorem [Dirichlet]. For any positive integer q, and any integer a coprime.
  23. [23]
    [PDF] The normal number of prime factors of a number n
    Theorem A′′. The normal order of the number of prime factors of a quadratfrei number is log log n. III. The normal order of f(n).
  24. [24]
    [PDF] Sieve Methods - cs.wisc.edu
    In this treatise we survey the major sieve methods and their important applications in number theory. We apply sieves to study the distribution of square-free ...
  25. [25]
    [PDF] 4 Random Graphs - CMU School of Computer Science
    Recall that G(n, p). “almost surely does not have the property” means that the probability that it has the property goes to zero in the limit, as n goes to ...
  26. [26]
    [PDF] ON THE EVOLUTION OF RANDOM GRAPHS - SNAP: Stanford
    this range the whole graph is not only almost surely connected, but the orders of all points are almost surely asymp:o:ically equal. Thus the graph becomes ...
  27. [27]
    [PDF] THE EVOLUTION OF RANDOM GRAPHS
    It is easily seen that for every fixed A we have lx ~ (logn)/(e-log(l/(l-e))). ~2(logn)/£ and /i-w(n)/e. < l\ < li+u(n)/e. Hence the assertion follows from ...
  28. [28]
    [PDF] Lectures 3 and 4 Hamiltonicity threshold in random graphs
    The main aim of these two lectures is to establish the Hamiltonicity threshold in the probability space G(n, p), this is the minimum value of the edge ...<|separator|>
  29. [29]
  30. [30]
    [PDF] the tur´an theorem for random graphs
    Abstract. The aim of this paper is to prove a Turán type theorem for random graphs. For 0 < γ ≤ 1 and graphs G and H, write G →γ H if any γ-proportion.
  31. [31]
    [PDF] Algebraic Groups - James Milne
    Dec 20, 2015 · An algebraic group is a matrix group defined by polynomial conditions. More abstractly, it is a group scheme of finite type over a field.
  32. [32]
    33.8 Geometrically irreducible schemes - Stacks Project
    13. Proof. Let T \subset X be an irreducible component of X. Let \eta \in T be its generic point. By Lemmas 33.8.10 and 33.8.11 the generic points of ...
  33. [33]
    [PDF] notes on algebraic geometry math 202a - Brandeis
    Oct 29, 2014 · Conversely, for any irreducible algebraic subset Y ⊆ Ωn, there is an element y ∈ Y. (called a generic point of Y ) so that Y = y. Proof. The ...
  34. [34]
    Hilbert's Irreducibility Theorem - SpringerLink
    Hilbert’s theorem states that if f(t, X) is irreducible, then there exist infinitely many rational t0 such that f(t0, X) is irreducible over Q.
  35. [35]
    [PDF] Hilbert's Proof of His Irreducibility Theorem - arXiv
    Sep 20, 2017 · Hilbert's irreducibility theorem states that an irreducible polynomial with integral coefficients can become irreducible in its variables alone ...
  36. [36]
    [PDF] Invertible matrix
    Jan 24, 2013 · Thus in the language of measure theory, almost all n-by-n matrices are invertible. Furthermore the n-by-n invertible matrices are a dense open ...
  37. [37]
    On approximately simultaneously diagonalizable matrices
    Aug 6, 2025 · We begin our discussion with the fact that almost all complex matrices over complex fields are diagonalizable 52, 54 . Namely, geometric and ...
  38. [38]
    [PDF] On sets of integers containing k elements in arithmetic progression
    ACTA ARITHMETICA. XXVII (1975). [1] P. X. Gallagher, A large sieve density ... It is well known and obvious that neither class must contain an infinite arithmetic ...
  39. [39]
    [PDF] SZEMERÉDI'S PROOF OF SZEMERÉDI'S THEOREM - Terry Tao
    In 1975, Szemerédi famously established that any set of integers of posi- tive upper density contained arbitrarily long arithmetic progressions. The proof was.
  40. [40]
    [PDF] Intersecting families of sets are typically trivial - arXiv
    Oct 22, 2024 · Let I(n, k) denote the number of k- uniform intersecting families in 2[n] and I(n, k, ≥ 1) the number of non-trivial such families. ∗Department ...
  41. [41]
    [PDF] Paul Erd˝os and the Probabilistic Method
    Dec 1, 2013 · The Probabilistic Method is one of the most significant contributions of Paul Erd˝os. Indeed, Paul himself said, during his 80th birthday ...
  42. [42]
    [PDF] The Probabilistic Method (Third edition)
    if the family consists of all subsets of a given set ...<|separator|>
  43. [43]
    [PDF] Tauberian Theorems and Prime Densities. by Klaus Hoechsmann
    Our aim is to derive a version of the so-called Tauberian Theorem by Wiener-Ikehara and to demonstrate some of its uses in analytic number theory.Missing: almost | Show results with:almost
  44. [44]
    [PDF] An Overview of the Sieve Method and its History - arXiv
    Dec 27, 2006 · 1.10 Brun's Sieve yields remarkable assertions not only about those great conjectures but also about fundamental queries in the theory of the ...<|separator|>
  45. [45]
    [PDF] the birkhoff ergodic theorem with applications - UChicago Math
    For f : X ! R measurable, if f T = f almost everywhere, then f is constant almost everywhere. We now use a method to prove ergodicity ...
  46. [46]
    [PDF] Probabilistic Methods in Combinatorics - Yufei Zhao
    Jun 18, 2024 · Above, the vertices are partitioned into three nearly equal sets 𝑉1,𝑉2,𝑉3, and all the edges come in two types: (i) with one vertex in ...
  47. [47]
    A simple proof that the square-free numbers have density 6/(pi^2)
    Abstract. In this note we give a simple proof of the well-known result that the square-free numbers have density 6 ∕ π2.Missing: almost | Show results with:almost
  48. [48]
    Limit distribution for the existence of hamiltonian cycles in a random ...
    Pósa proved that a random graph with cn log n edges is Hamiltonian with probability tending to 1 if c > 3. Korsunov improved this by showing that, if Gn is ...