Fact-checked by Grok 2 weeks ago

List of numbers

A list of numbers in refers to the systematic classification of numerical values and structures that form the basis of , , and other branches of the field, encompassing sets such as natural numbers, , integers, rational numbers, irrational numbers, real numbers, and complex numbers. These categories build upon one another hierarchically, starting with the simplest counting elements and extending to more abstract systems that enable solutions to equations involving roots and imaginaries. Natural numbers are the foundational set used for counting and ordering, commonly consisting of positive integers beginning with 1 (i.e., 1, 2, 3, ...), excluding zero and negatives, though some definitions (such as in set theory) include zero. Whole numbers extend this set by including zero, forming {0, 1, 2, 3, ...}, which supports basic operations without fractions. Integers broaden the scope further to include negative values, yielding the complete set {..., -3, -2, -1, 0, 1, 2, 3, ...}, essential for representing differences and debts in real-world modeling. Rational numbers comprise all integers and fractions where the numerator and denominator are integers (with denominator nonzero), such as 1/2 or -3/4, and they are dense on the , meaning any two can be separated by another rational. In contrast, irrational numbers cannot be expressed as such fractions, including constants like √2 or π, and they fill the gaps in to complete the real numbers, which include all points on the infinite and support continuous functions. Finally, complex numbers incorporate imaginary units, expressed as a + bi where i^2 = -1, allowing solutions to equations like x^2 + 1 = 0 and extending applications to fields like and . This progression from basic counting to advanced structures mirrors the historical evolution of mathematical systems.

Natural Numbers

Small Natural Numbers

Small natural numbers, ranging from to 10, serve as the foundational elements of , , and numerical representation in and daily . These numbers emerged in ancient civilizations through early systems, where tallies or notches on bones and stones represented quantities starting from , as seen in Mesopotamian records dating back over 5,000 years and around 3000 BCE. Over time, these systems evolved into structured numeral notations, with base-10 likely influenced by human finger anatomy in many cultures. The following table summarizes the unique properties of these small natural numbers, highlighting their basic roles in arithmetic and number theory:
NumberKey Properties
1Multiplicative identity (a \times 1 = a for any a); the multiplicative unit, neither prime nor composite.
2Smallest prime number; smallest even number; the only even prime.
3Smallest odd prime; sum of first two natural numbers (1+2).
4Smallest composite number ($2 \times 2); first perfect square after 1 ($2^2).
5Prime number; forms the basis for the pentagon in geometry.
6First perfect number (sum of proper divisors 1+2+3=6); smallest number divisible by 1, 2, and 3.
7Prime number; generates the heptagon in geometry.
8First power of 2 greater than 4 ($2^3); highly composite with divisors 1, 2, 4.
9First perfect square after 4 ($3^2); sum of first three odd numbers (1+3+5).
10Basis of the decimal system; sum of first four natural numbers (1+2+3+4).

Mathematical Properties

The natural numbers \mathbb{N} are rigorously defined through the , a set of five foundational postulates introduced by Italian mathematician in his 1889 publication Arithmetices principia. These axioms establish the structure of \mathbb{N} (starting from 1) as follows:
  • One axiom: 1 is a natural number.
  • Successor axiom: For every natural number n, there exists a unique successor S(n) that is also a natural number.
  • No predecessor for one: There is no natural number n such that S(n) = 1.
  • Injectivity of successor: If S(m) = S(n), then m = n.
  • Induction axiom: If a set K \subseteq \mathbb{N} contains 1 and, whenever it contains n, it also contains S(n), then K = \mathbb{N}.
These axioms capture the intuitive notion of natural numbers as generated successively from 1, ensuring a discrete, infinite, and well-ordered set without gaps or cycles. A key consequence of the Peano axioms is the principle of , which provides a method to prove properties holding for all natural numbers. It states: If a property P(n) holds for n = 1 (base case) and, whenever it holds for some n, it also holds for S(n) (inductive step), then P(n) holds for all n \in \mathbb{N}. This principle is equivalent to the induction axiom in first-order formulations and underpins proofs in and beyond. The describes the multiplicative structure of natural numbers greater than 1: every such number can be uniquely factored into a product of prime numbers, disregarding order. Formally, for any integer n > 1, n = p_1^{e_1} p_2^{e_2} \cdots p_k^{e_k}, where p_1 < p_2 < \cdots < p_k are distinct primes and e_i \geq 1 are positive integers. This uniqueness (up to associates in more general rings) was first rigorously proved by Carl Friedrich Gauss in his seminal 1801 work Disquisitiones Arithmeticae, building on earlier ideas from Euclid. The theorem implies that prime factorization is a canonical representation, essential for algorithms in cryptography and computational number theory. Divisibility rules offer practical tests for whether one natural number divides another, derived from properties of base-10 representation and modular arithmetic. For small divisors, these rules simplify checks without full division: a number is divisible by 2 if its last digit is even (0, 2, 4, 6, or 8); by 3 if the sum of its digits is divisible by 3; by 5 if its last digit is 0 or 5; by 9 if the sum of its digits is divisible by 9; and by 10 if its last digit is 0. For example, 1236 is divisible by 3 because $1 + 2 + 3 + 6 = 12, and 12 is divisible by 3. These rules stem from congruences like $10 \equiv 1 \pmod{9}, enabling efficient verification of factors.

Cultural and Practical Uses

Natural numbers play a prominent role in cultural calendars and symbolic traditions worldwide. The number 7, for instance, structures the seven-day week, a cycle rooted in ancient Mesopotamian astronomy and religious practices, where it symbolized completeness and was used in rituals like temple offerings and exorcisms. This septenary division influenced Jewish, Christian, and Islamic calendars, reflecting the seven classical planets known to antiquity. Similarly, the number 12 divides the year into twelve months in the Gregorian calendar, derived from the Roman lunar-solar system that approximated the solar year with twelve lunar cycles. In astrology, the twelve zodiac signs, originating from Babylonian astronomy around the 5th century BCE, partition the ecliptic into equal segments for horoscopic purposes. In practical applications, natural numbers underpin measurement and computational systems essential to commerce and technology. The base-10 (decimal) system, likely developed independently in ancient Egypt and Mesopotamia due to human finger counting, facilitated trade by enabling efficient recording of quantities on clay tablets and papyrus for accounting and inventory. This positional notation, refined by Indian mathematicians around 500 CE and transmitted via Arab scholars, became the global standard for economic transactions. In contrast, the binary system, a base-2 numeral system using only 0 and 1, was formalized by Gottfried Wilhelm Leibniz in the 1690s, inspired by ancient Chinese I Ching hexagrams, and later adopted in the 20th century for digital computing due to its compatibility with electronic switches. Leibniz envisioned binary arithmetic as a foundation for mechanical calculators, paving the way for modern computers. Cultural taboos associated with certain natural numbers reveal symbolic fears tied to language and history. In Western cultures, 13 is deemed unlucky, tracing to Norse mythology where Loki, the 13th guest at a divine banquet, caused the death of Balder, and reinforced by the Christian Last Supper with 13 attendees including Judas the betrayer. This triskaidekaphobia leads to omissions like skipping the 13th floor in buildings. In East Asian societies, particularly China, Japan, and Korea, the number 4 evokes tetraphobia because its pronunciation ("shi") homophonically resembles "death," prompting avoidance in addresses, phone numbers, and hospital rooms. In ancient philosophy, perfect numbers such as 6 and 28, equal to the sum of their proper divisors, symbolized cosmic harmony for Pythagoreans, who viewed them as embodiments of divine order.

Special Classes of Natural Numbers

Prime Numbers

A prime number is a natural number greater than 1 that has no positive divisors other than 1 and itself. This definition originates from ancient Greek mathematics, where Euclid described such numbers as those "measured by a unit alone." Primes form the foundational building blocks of all natural numbers through multiplication, underscoring their central role in number theory as the atoms of the integer system. The first few prime numbers are 2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31, 37, 41, 43, 47, 53, 59, 61, 67, 71, 73, 79, 83, 89, and 97. Note that 2 is the only even prime number, with all subsequent primes being odd. These examples illustrate the distribution of primes among natural numbers, becoming sparser as numbers increase, yet remaining essential for understanding divisibility and factorization in number theory. Euclid proved that there are infinitely many prime numbers in his Elements, a result known as . The proof proceeds by contradiction: assume there are finitely many primes p_1, p_2, \dots, p_k, and construct the number N = p_1 p_2 \cdots p_k + 1. This N is greater than 1 and not divisible by any p_i, so it must either be prime itself or divisible by some prime larger than p_k, contradicting the assumption of finiteness. This theorem highlights the unbounded nature of primes, influencing countless results in analytic and algebraic . Twin primes are pairs of primes that differ by 2, such as (3, 5), (5, 7), (11, 13), and (17, 19). The twin prime conjecture posits that there are infinitely many such pairs, a problem dating back to ancient times but formally stated in the 19th century. Similarly, , proposed in 1742, states that every even integer greater than 2 can be expressed as the sum of two primes, such as 4 = 2 + 2 and 10 = 3 + 7. These unsolved conjectures underscore the enduring mysteries surrounding primes and their additive properties in number theory.

Highly Composite Numbers

A highly composite number is a positive integer n such that the number of its positive divisors, denoted by the divisor function d(n), is greater than d(m) for every positive integer m < n. This property makes these numbers particularly rich in factors relative to their magnitude, with the sequence beginning 1, 2, 4, 6, 12, 24, 36, 48, 60, 120, and continuing to larger values like 840. For instance, 12 has d(12) = 6 divisors (1, 2, 3, 4, 6, 12), exceeding the divisor counts of all smaller positives; similarly, 60 has d(60) = 12 divisors, and 840 has d(840) = 32. In 1915, Srinivasa Ramanujan extended this concept by defining superior highly composite numbers, a subclass where, for some \epsilon > 0, the ratio d(n)/n^\epsilon is at least as large as d(k)/k^\epsilon for all positive integers k. These numbers satisfy a stricter condition on density and form a of highly composite numbers, including 1, 2, 6, 12, , 840, , and 2520, among others. Ramanujan's formulation emphasizes their optimal structure in terms of prime factorization, where exponents in the prime power decomposition are non-increasing. Highly composite numbers find applications in factorization efficiency due to their composition from small primes with balanced exponents, allowing rapid decomposition compared to numbers with fewer or larger factors. In geometry, they enable constructions of highly symmetric polygons; for example, a regular 60-sided polygon benefits from the 12 divisors of , supporting diverse rotational and reflectional symmetries in its . Additionally, their abundance of divisors makes them practical in engineering and measurement systems, such as using 12 for inches in a foot or for seconds in a minute, simplifying fractional calculations and subdivisions.

Perfect and Amicable Numbers

A perfect number is a positive integer equal to the sum of its proper divisors, excluding the number itself. For example, 6 is perfect because its proper divisors are 1, 2, and 3, and $1 + 2 + 3 = 6; similarly, 28 is perfect as $1 + 2 + 4 + 7 + 14 = 28. The ancient Greeks, including , studied these numbers, recognizing the first four: 6, 28, 496, and 8128. All known perfect numbers are even, and proved in the BCE that every number of the form $2^{p-1}(2^p - 1), where $2^p - 1 is prime, is perfect. In the , Leonhard Euler demonstrated that this form generates all even perfect numbers, establishing the Euclid-Euler theorem. Here, $2^p - 1 must be a , linking perfect numbers to the search for such primes. As of 2025, 52 even perfect numbers are known, each corresponding to a discovered . Whether odd perfect numbers exist remains an unsolved problem in , dating back over two millennia and considered one of the oldest open questions in . Extensive searches and theoretical bounds, such as requiring at least 101 prime factors and exceeding $10^{1500} in size, have failed to find any, but no proof of nonexistence exists. Amicable numbers form pairs of distinct positive integers where the sum of the proper divisors of each equals the other number. The smallest such pair is 220 and 284: the proper divisors of 220 sum to 284 ($1 + 2 + 4 + 5 + 10 + 11 + 20 + 22 + 44 + 55 + 110 = 284), and those of 284 sum to 220 ($1 + 2 + 4 + 71 + 142 = 220). Discovered by ancient mathematicians like Iamblichus around 200 CE, amicable pairs extend the perfect number concept to mutual divisor sums. Over 1.2 billion such pairs are known as of 2023, with no general formula for generating all pairs, though specific constructions exist for certain forms.

Other Integers

Zero

Zero occupies a unique position in the number system as neither positive nor negative, serving as the between the two and embodying the concept of nothingness in quantitative terms. This neutrality distinguishes it from all other integers, enabling its foundational role in arithmetic and . Unlike positive or negative numbers, zero does not contribute to in or but is indispensable for representing absence or in equations and measurements. The concept of zero as a number originated in ancient , where it evolved from a placeholder in to a full-fledged with arithmetic properties. By the CE, the mathematician formalized its use in his treatise Brahmasphuṭasiddhānta (628 CE), providing the first explicit rules for operations involving zero, such as stating that zero added to or subtracted from any number yields that number unchanged. This innovation built on earlier Indian developments, including the from around the 3rd–4th century CE, which used a dot as a zero in decimal calculations. 's work treated zero as an independent number, revolutionizing by allowing consistent handling of voids in numerical systems. These Indian advancements were adopted into the Arabic numeral system during the Islamic Golden Age, with scholars like Muhammad ibn Musa al-Khwarizmi incorporating zero into his 9th-century texts on algebra and arithmetic, which used the term sifr (meaning "empty") for zero. This integration facilitated the spread of the decimal place-value system to Europe via translations and the work of Fibonacci in the 13th century, establishing zero as a core element of modern numerals. The adoption transformed global computation, enabling efficient representation of large and small quantities. Mathematically, zero functions as the additive identity in the ring of integers, satisfying the property that for any integer a, a + 0 = 0 + a = a. This ensures that adding zero preserves the value of any expression, forming the basis for many algebraic identities and proofs. However, division by zero remains undefined, as no integer q satisfies $0 \cdot q = a for a \neq 0, and $0/0 leads to indeterminacy since any number multiplied by zero yields zero. This undefined nature prevents inconsistencies in the field axioms and underscores zero's exceptional status in arithmetic operations./07:_The_Properties_of_Real_Numbers/7.05:_Properties_of_Identity_Inverses_and_Zero) In , zero represents the of the \emptyset, which contains no elements and thus has measure zero. This correspondence defines the natural number 0 in axiomatic constructions like von Neumann's ordinals, where \emptyset is the with zero elements. The 's establishes zero as the foundational in .

Negative Integers

Negative integers are the less than zero, expressed as -1, -2, -3, and continuing indefinitely. They form the set { ..., -3, -2, -1 }, symmetric to the positive integers through the of additive inversion, where for any positive n, the negative -n satisfies n + (-n) = 0, formally defined as -n = 0 - n. The historical development of negative integers traces back to ancient financial practices, particularly in 7th-century , where mathematician (c. 598–668 ) formalized their use in his treatise Brahmasphutasiddhanta (628 ), interpreting positive numbers as fortunes or assets and negative numbers as s or deficiencies to balance accounts. This practical application in systems provided an early conceptual foundation, enabling operations like and involving negatives. In , acceptance grew in the 17th century through ' introduction of in (1637), where the coordinate plane naturally extended the number line to include negative directions, despite Descartes' initial reluctance to treat negative roots of equations as valid solutions, labeling them "false." A key property of negative integers is the absolute value function, which for a negative integer -n (with n > 0) yields |-n| = n, measuring the distance from zero on the number line irrespective of sign. This function underscores their magnitude equivalence to positives while preserving directional distinction. On the real number line, negative integers occupy positions to the left of zero, ordered such that \dots < -4 < -3 < -2 < -1 < 0, with greater negativity indicating smaller values and increasing distance from the origin. Zero acts as the neutral boundary separating negative integers from their positive counterparts.

Powers of Ten and SI Prefixes

Powers of ten provide a systematic way to denote large integers by expressing them as 10 raised to an integer exponent, where 10^n equals 1 followed by n zeros in decimal notation. This base-10 system is foundational in mathematics and science for scaling numbers efficiently, starting from 10^0 = 1 and increasing to extremely large values such as 10^3 = 1,000 or 10^6 = 1,000,000. The International System of Units (SI) formalizes the naming of these powers through prefixes, which attach to base units to indicate multiples by factors of 10^3 or higher (and submultiples for smaller scales, though the focus here is on large integers). These prefixes, established by the International Bureau of Weights and Measures (BIPM), extend up to 10^30 as of the 2022 revision, enabling concise expression of vast quantities in fields like physics and computing. For powers below 10^3, prefixes like deca (10^1) and hecto (10^2) exist but are less commonly applied to standalone integers. The following table lists the official SI prefixes for positive powers of ten, including the most recent additions of ronna (10^27) and quetta (10^30), which address growing needs in data storage and cosmology.
FactorNameSymbol
10^0(none)
10^1decada
10^2hectoh
10^3kilok
10^6megaM
10^9gigaG
10^12teraT
10^15petaP
10^18exaE
10^21zettaZ
10^24yottaY
10^27ronnaR
10^30quettaQ
Beyond the SI framework, informal mathematical names describe even larger powers of ten, such as the googol, defined as 10^100 and coined in 1940 by mathematician Edward Kasner to illustrate immense scales in popular science. This number, 1 followed by 100 zeros, vastly exceeds practical SI applications but highlights the conceptual extension of powers of ten. In practice, powers of ten underpin scientific notation, a compact method for writing numbers as a coefficient between 1 and 10 multiplied by 10 raised to an exponent, essential for handling extreme scales in astronomy, particle physics, and engineering. For instance, the observable universe's diameter is approximately 8.8 × 10^26 meters, demonstrating how these powers quantify cosmic dimensions without cumbersome digit strings.

Rational Numbers

Unit Fractions

A unit fraction is a positive rational number expressed as \frac{1}{n}, where n is a positive integer. These fractions formed the basis of fractional notation in ancient , where all rational numbers were represented as sums of distinct unit fractions, a practice documented in surviving papyri. This system avoided repeated denominators and non-unit numerators except for the special case of \frac{2}{3}, reflecting the Egyptians' preference for additive decompositions in arithmetic problems related to division, area, and volume. The Rhind Mathematical Papyrus, dating to approximately 1650 BC and attributed to the scribe Ahmes, contains extensive tables and examples of Egyptian fraction expansions, including a 2/n table for odd n from 5 to 101. For instance, the fraction \frac{2}{5} is decomposed as \frac{1}{3} + \frac{1}{15}, illustrating how Egyptians solved practical problems like dividing loaves or measuring fields by summing unit fractions. Such representations ensured exactness without relying on a general fractional notation, influencing later mathematical traditions. Among notable unit fractions, \frac{1}{2} equals 0.5 in decimal form and serves as the foundational halving operation in many ancient calculations. The fraction \frac{1}{3} has the repeating decimal expansion $0.\overline{3}, embodying the concept of thirds central to Egyptian divisions of resources. Similarly, \frac{1}{7} yields the six-digit repeating decimal $0.\overline{142857}, a cyclic pattern arising from the long period of its denominator in base 10. Unit fractions are fundamental to the harmonic series, whose partial sums are the harmonic numbers defined by H_n = \sum_{k=1}^n \frac{1}{k} = 1 + \frac{1}{2} + \frac{1}{3} + \dots + \frac{1}{n}. This sum approximates \ln n + \gamma, where \gamma \approx 0.5772156649 is the , providing insight into the divergent yet slowly growing nature of the series. The difference H_n - \ln n converges to the irrational \gamma, linking unit fractions to transcendental analysis.

Famous Rational Approximations

One of the most renowned rational approximations is \frac{22}{7} \approx 3.142857, which serves as an upper bound for \pi. This fraction was established by Archimedes in the 3rd century BCE through his method of inscribing and circumscribing polygons around a circle to bound the value of \pi, proving \frac{223}{71} < \pi < \frac{22}{7}. A superior approximation to \pi is \frac{355}{113} \approx 3.14159292, accurate to six decimal places and remaining the best rational approximation with a denominator up to four digits until the 16th century. This fraction was independently discovered by the Chinese mathematician Zu Chongzhi around 480 CE using an iterative polygon method extending Archimedes' approach with a 24,576-sided polygon. For the base of the natural logarithm e, continued fraction convergents provide notable rational approximations, such as \frac{19}{7} \approx 2.714286, which is the fifth convergent in the expansion e = [2; \overline{1, 2k, 1}] for increasing integers k. This arises from the series definition of e and yields an error of less than 0.004. Rational approximations to the golden ratio \phi = \frac{1 + \sqrt{5}}{2} \approx 1.618034 are generated by ratios of consecutive Fibonacci numbers, with \frac{8}{5} = 1.6 as an early convergent accurate to two decimal places. These convergents stem from the continued fraction \phi = [1; \overline{1}], where the error decreases as | \phi - \frac{F_{n+1}}{F_n} | < \frac{1}{F_n F_{n+1}}, with F_n denoting the nth Fibonacci number. Simple rationals like \frac{3}{4} = 0.75 appear in basic decimal approximations or quadratic equation solutions, illustrating everyday uses beyond advanced constants.

Real Numbers

Algebraic Real Numbers

Algebraic real numbers are the real roots of non-zero polynomial equations with integer (or equivalently, rational) coefficients. These numbers form a field extension of the rational numbers \mathbb{Q}, and every algebraic real number \alpha has a minimal polynomial over \mathbb{Q} of some finite degree n \geq 1, meaning [\mathbb{Q}(\alpha) : \mathbb{Q}] = n. Rationals themselves are algebraic of degree 1, while irrationals like \sqrt{2} have higher degrees. A classic example is \sqrt{2}, the positive real solution to the quadratic equation x^2 - 2 = 0, which has minimal degree 2 over \mathbb{Q} and approximates 1.414213562. Another prominent example is the golden ratio \phi = \frac{1 + \sqrt{5}}{2} \approx 1.618033989, which satisfies the minimal polynomial x^2 - x - 1 = 0 and also has degree 2; it arises in geometry, such as the ratio of successive . These examples illustrate how algebraic reals can be expressed explicitly using radicals in some cases, though not always for higher degrees. Among algebraic real numbers, constructible numbers form an important subclass, defined as those obtainable from the integers via a finite sequence of field operations (addition, subtraction, multiplication, division) and square root extractions. The field extension generated by a constructible number over \mathbb{Q} has degree $2^k for some non-negative integer k, reflecting the quadratic nature of each adjoining step; for instance, \sqrt{2} is constructible (degree 2), as is \sqrt{2 + \sqrt{2}} (degree 4). This property underpins classical compass-and-straightedge constructions in Euclidean geometry, limiting which regular polygons (e.g., pentagon, but not heptagon) can be constructed.

Transcendental Real Numbers

Transcendental real numbers are real numbers that are not algebraic, meaning they are not roots of any non-zero polynomial equation with rational coefficients. This class of numbers arises naturally in analysis and geometry, distinguishing them from algebraic reals that satisfy polynomial relations. The number e, the base of the natural logarithm, is a fundamental example of a transcendental real number. It was proved transcendental by Charles Hermite in 1873 using integral representations and properties of exponential functions. Hermite's proof relies on showing that assuming e algebraic leads to a contradiction via continued fraction approximations and factorial growth rates. The number e admits the infinite series expansion e = \sum_{n=0}^\infty \frac{1}{n!}, which converges rapidly and defines e approximately as 2.71828. Another prominent example is \pi, the ratio of a circle's circumference to its diameter, proved transcendental by in 1882. Lindemann's proof extends techniques, establishing the Lindemann-Weierstrass theorem that e^{i\alpha} is transcendental for non-zero algebraic \alpha, implying \pi (as -i \ln(-1)) cannot be algebraic. This result resolved the ancient problem of , showing it impossible with straightedge and compass. The provides a series for \pi: \pi = 4 \sum_{n=0}^\infty \frac{(-1)^n}{2n+1}, which alternates and converges to approximately 3.14159, though slowly. The existence of transcendental numbers was first demonstrated in 1844 by through an explicit construction using . bounds how well algebraic irrationals can be approximated by rationals, allowing him to build a number \lambda = \sum_{k=1}^\infty 10^{-k!} that violates this bound for any algebraic degree, proving it transcendental. Such are highly approximable by rationals and form a dense set in the reals, though they have Lebesgue measure zero.

Numbers of Uncertain Irrationality

Numbers of uncertain irrationality encompass real constants that mathematicians widely believe to be irrational or even transcendental, yet definitive proofs remain elusive despite centuries of investigation. These open problems highlight the challenges in transcendental number theory, where established techniques like continued fractions or modular forms have succeeded for some constants but falter for others. Key examples include linear and multiplicative combinations of fundamental constants such as e and \pi, as well as values of the at odd integers. A classic case is e + \pi, whose irrationality has been an unresolved question since the 18th century, when explored properties of e and \pi. Although it is proven that at least one of e + \pi and e\pi must be irrational—via the non-vanishing of the minimal polynomial for algebraic dependence—the specific irrationality of e + \pi is unknown. Similarly, \pi / e is suspected to be irrational, but no proof exists, representing another longstanding open problem in the field. Apéry's constant, denoted \zeta(3) = \sum_{n=1}^\infty \frac{1}{n^3}, provides a notable instance where irrationality is established but higher-degree properties are not. In 1979, Roger Apéry demonstrated its irrationality through a clever construction of continued fraction convergents satisfying a specific recurrence relation, marking a breakthrough after prior failures for odd zeta values beyond \zeta(2). However, whether \zeta(3) is transcendental—meaning not algebraic over the rationals—remains unknown, with conjectures suggesting it is, based on patterns in zeta function values. Other examples include $2^e, where irrationality is anticipated but unproven, illustrating the difficulty in handling exponential forms with transcendental bases or exponents. These cases underscore ongoing research directions, such as improving irrationality measures or leveraging arithmetic geometry to resolve such statuses.

Imprecisely Known Real Constants

Imprecisely known real constants are mathematical entities whose numerical values can only be approximated within wide bounds due to prohibitive computational demands, often stemming from exponential growth in required resources as precision increases. Unlike computable constants with algorithmic paths to arbitrary accuracy, these arise in number theory and combinatorics where exhaustive search or simulation becomes infeasible beyond certain scales. This imprecision highlights the boundaries of current computational capabilities, even as supercomputers push limits in other domains. A classic example is Skewes' number, an upper bound for the smallest integer where the prime counting function π(n) exceeds the logarithmic integral li(n), assuming the Riemann hypothesis holds. Stanley Skewes originally derived a bound of approximately $10^{10^{10^{34}}} in 1933, but refinements over decades have reduced it significantly; as of 2005, the bound stands below about $1.4 \times 10^{316}. The difficulty arises from the need to analyze oscillatory behaviors in prime distribution, which requires evaluating complex integrals and zeta function zeros over vast ranges, rendering exact determination practically impossible with current methods. Ramsey numbers exemplify combinatorial constants with imprecise values due to exponential verification times. The diagonal Ramsey number R(5,5), defined as the minimal vertices ensuring a monochromatic K_5 in any 2-edge-coloring of the complete graph, is bounded by 43 ≤ R(5,5) ≤ 46. The lower bound of 43 was established via a explicit coloring construction in 1989, while the upper bound of 46 was proven in 2024 by showing no critical colorings exist at that scale. Computing exact values demands checking an astronomically large number of graph colorings, with the search space growing factorially, limiting progress to incremental bound improvements. In dynamical systems like the Collatz conjecture, bounds on sequence behaviors illustrate similar challenges. The conjecture posits that iterating the rule (3n+1 if odd, n/2 if even) reaches 1 for any positive integer n, verified computationally up to approximately 4 × 10^{21} as of 2025, but establishing global bounds on stopping times or maximum heights requires simulating paths that can explode to exponential sizes relative to n. This leads to imprecise estimates, as full exploration for larger n demands resources scaling exponentially with bit length. In contrast, transcendental constants like π, detailed elsewhere, have been computed to 300 trillion digits as of April 2025 using efficient series accelerations, underscoring how algorithmic tractability varies across real constants.

Complex and Hypercomplex Numbers

Standard Complex Numbers

Complex numbers extend the real numbers by incorporating the imaginary unit i, defined such that i^2 = -1. A standard complex number is expressed as z = a + bi, where a and b are real numbers representing the real and imaginary parts, respectively. This form allows solutions to equations like x^2 + 1 = 0, which have no real roots, and underpins applications in fields such as electrical engineering and quantum mechanics./03%3A_Polynomial_and_Rational_Functions/3.01%3A_Complex_Numbers) The modulus of a complex number z = a + bi, denoted |z|, measures its distance from the origin in the complex plane and is given by |z| = \sqrt{a^2 + b^2}. This Euclidean norm facilitates geometric interpretations, such as representing complex numbers as points or vectors in a two-dimensional plane with real and imaginary axes. For instance, the modulus of z = 3 + 4i is \sqrt{3^2 + 4^2} = 5, illustrating the in this context./03%3A_Polynomial_and_Rational_Functions/3.01%3A_Complex_Numbers) Gaussian integers form a significant subset of complex numbers, consisting of those where both a and b are integers, denoted as a + bi with a, b \in \mathbb{Z}. Introduced by in 1832, they constitute the ring \mathbb{Z} and enable unique factorization in a manner analogous to ordinary integers, though with different primes; for example, $5 = (1 + 2i)(1 - 2i). A representative Gaussian integer is $1 + i, whose modulus is \sqrt{2}. These integers are fundamental in algebraic number theory for studying quadratic fields. A profound relation in complex analysis is Euler's identity, which states e^{i\pi} + 1 = 0, connecting five fundamental constants: e, i, \pi, 1, and 0. Derived from the exponential form of complex numbers, e^{i\theta} = \cos \theta + i \sin \theta, it emerges when \theta = \pi, highlighting the deep interplay between exponential, trigonometric, and imaginary functions. First published by Leonhard Euler in 1748, this identity exemplifies the elegance of complex numbers./01%3A_Complex_Algebra_and_the_Complex_Plane/1.06%3A_Euler%27s_Formula)

Quaternions and Octonions

Quaternions represent a four-dimensional extension of the complex numbers, introduced by in 1843 to handle three-dimensional rotations and geometric transformations. A quaternion is expressed as q = a + bi + cj + dk, where a, b, c, d are real numbers, and i, j, k are imaginary units satisfying the relations i^2 = j^2 = k^2 = -1 and ijk = -1. These rules ensure that quaternion multiplication is non-commutative, distinguishing them from complex numbers, which are commutative in two dimensions. Hamilton's original formulation appeared in his 1843 communication to the , with a systematic treatment published later in his 1853 lectures. Quaternions find significant applications in three-dimensional rotations, particularly in computer graphics, robotics, and aerospace engineering, where they provide a compact and numerically stable representation that avoids singularities like gimbal lock associated with Euler angles. Unit quaternions, those with norm 1, form the group SU(2), which double-covers the rotation group SO(3), enabling efficient interpolation and composition of rotations. This computational utility was highlighted in early work on quaternion-based rotation algorithms, which demonstrated their efficiency over matrix methods for storage and computation. Octonions extend this progression to eight dimensions, forming an eight-dimensional algebra discovered independently by John T. Graves in 1843 and Arthur Cayley in 1845. An octonion can be written as o = a_0 + a_1 e_1 + \cdots + a_7 e_7, where the e_i are basis elements with specific multiplication rules that render octonions non-commutative and non-associative, though they remain alternative algebras. Unlike quaternions, the lack of associativity complicates their use, but octonions appear in advanced theoretical physics, including string theory, where their structure relates to exceptional Lie groups and the geometry of higher-dimensional spacetime. For instance, octonionic formulations have been proposed to unify aspects of superstring theory and the standard model of particle physics. The quaternions and octonions are linked through the , a recursive process that doubles the dimension of a normed algebra while preserving certain properties like the norm but progressively losing commutativity and associativity. Starting from the , one application yields the , a second the quaternions, and a third the octonions; further iterations produce sedenions and beyond, though only up to octonions retain division algebra properties. This construction, formalized by in the early 20th century but rooted in 19th-century discoveries, underscores the unique sequence of real division algebras in dimensions 1, 2, 4, and 8.

Transfinite Numbers

Cardinal Numbers

Cardinal numbers, also known as cardinals, provide a measure of the size of a set, extending the intuitive counting of elements in finite collections to infinite ones through the concept of bijections between sets. Two sets have the same cardinality if there exists a one-to-one correspondence between their elements, allowing sets of vastly different appearances to be deemed equally large. This framework, developed by in the late 19th century, revolutionized the understanding of infinity by revealing that not all infinite sets are of the same size. For finite sets, cardinal numbers coincide with the natural numbers, where the cardinality of a set is simply the number of distinct elements it contains; for example, a set with three elements has cardinality 3. This aligns directly with everyday counting, and any two finite sets of the same size can be paired bijectively without leftovers. In contrast, infinite cardinalities begin with \aleph_0 (aleph-null), the smallest infinite cardinal, which is the size of the set of natural numbers (including 0 in the set-theoretic sense) \mathbb{N} = \{0, 1, 2, \dots \}. Cantor established that the set of integers \mathbb{Z}, including negatives and zero, also has cardinality \aleph_0, as it can be bijectively mapped to \mathbb{N} via a simple enumeration that alternates positive and negative values. Similarly, the rational numbers \mathbb{Q} are countable, possessing the same cardinality \aleph_0, demonstrating that some infinite sets can be listed in a sequence despite their apparent density. A fundamentally larger infinite cardinal is the cardinality of the continuum, denoted $2^{\aleph_0} or \mathfrak{c}, which is the size of the set of real numbers \mathbb{R}. Cantor proved this in 1891 using his diagonal argument: assuming a bijection between \mathbb{N} and \mathbb{R} leads to a contradiction, as one can construct a real number differing from every listed number in at least one decimal place, corresponding to the diagonal of the enumeration matrix. This shows that the power set of \mathbb{N}, whose elements are all subsets of the naturals, has cardinality $2^{\aleph_0}, and since \mathbb{R} can be injectively mapped into this power set (via binary expansions), \mathfrak{c} = 2^{\aleph_0} > \aleph_0. The argument highlights the existence of uncountably sets, establishing a hierarchy of infinities. The () posits that no exists strictly between \aleph_0 and $2^{\aleph_0}, implying \mathfrak{c} = \aleph_1, the next after \aleph_0 in the aleph hierarchy. In 1938, demonstrated that is consistent with Zermelo-Fraenkel including the (ZFC), by constructing a model (the constructible L) where holds, assuming ZFC is consistent. This relative consistency result showed cannot be disproved within ZFC. Complementing this, Paul Cohen proved in 1963 that the of is also consistent with ZFC, using his forcing technique to build a model where $2^{\aleph_0} > \aleph_1. Together, these results establish the independence of from ZFC, meaning neither nor its can be proved or disproved using standard set-theoretic axioms.

Ordinal Numbers

Ordinal numbers, or ordinals, generalize the concept of counting positions in a to well-ordered sets, extending beyond finite lengths to transfinite ones. They represent the of such sets, where each ordinal α is the set of all ordinals strictly less than it, ensuring a transitive and well-ordered structure under membership. The finite ordinals correspond precisely to the natural numbers in set-theoretic constructions, such as the ordinals, where 0 is the ∅, 1 is {∅}, 2 is {∅, {∅}}, and so on, up to any finite n built by successive successors. These form the initial segment of all ordinals and capture the order types of finite well-orderings. The first infinite ordinal is denoted ω, which is the least upper bound of all finite ordinals and represents the of the natural numbers under their usual ordering. Successor ordinals beyond ω include ω + 1, obtained by adjoining a single element after the ω-sequence, and limit ordinals like ω + 2 or higher. Further examples involve and : ω ⋅ 2 is the order type of two copies of ω concatenated, while ω² = sup{ω ⋅ n | n < ω} denotes the limit of ω ⋅ n for finite n, illustrating polynomial-like growth in ordinal hierarchy. Transfinite induction provides a foundational proof principle for ordinals, analogous to on naturals but extended to all ordinals. To prove a property P(α) holds for every ordinal α, it suffices to show that if P(β) holds for all β < α, then P(α) holds; the well-ordering ensures no infinite descending chains, validating the argument across transfinite structures. This principle underpins many results in , such as the existence of operations. Ordinal arithmetic defines operations on s, preserving well-orderings. Addition α + β is the of the of sets with order types α and β, where the β portion follows the α portion entirely; this operation is associative but not commutative. For instance, + ω equals ω, as adjoining a single element before an infinite absorbs into the limit, whereas ω + strictly exceeds ω by placing the extra element after the infinite . α ⋅ β concatenates β copies of α, and α^β builds iterated multiplications, yielding a non-commutative algebra distinct from cardinal arithmetic. Among countable ordinals, the Church-Kleene ordinal, denoted ω₁^{CK}, is the smallest ordinal greater than ω that is inadmissible, meaning it lacks a definable well-ordering relative to recursive functions and fails to satisfy the replacement schema for recursive notations. It serves as the supremum of all recursive ordinals, marking the boundary of computably representable well-orderings in the hyperarithmetic hierarchy.

Numbers in Physical Contexts

Fundamental Physical Constants

Fundamental physical constants are fixed numerical values that underpin the laws of physics, appearing in equations describing electromagnetic, gravitational, quantum, and other phenomena. These constants are either defined exactly through international agreements, such as the revision of the units, or determined through high-precision measurements. The Committee on Data for Science and Technology (CODATA) periodically recommends values based on least-squares adjustments of experimental data to ensure consistency across scientific fields. The in , denoted c, is a of and , representing the maximum speed at which information can propagate. It is defined exactly as c = 299792458 m/s, establishing the meter as the distance travels in in $1/299792458 of a second. This exact value has been in place since the 17th General Conference on Weights and Measures in 1983. The , h, quantifies the scale of quantum mechanical effects and is central to the energy-frequency relation E = h\nu for photons. It is now defined exactly as h = 6.62607015 \times 10^{-34} J s, fixing the in terms of the , , and . This definition was adopted in 2019 to replace the previous artifact-based mass standard, improving measurement precision in quantum technologies. The , \alpha, is a that characterizes the strength of the electromagnetic interaction between elementary charged particles. It appears in the of spectral lines and is approximately \alpha \approx 7.2973525643 \times 10^{-3}, or equivalently \alpha^{-1} \approx 137.035999177(21). Unlike the others, \alpha is not exactly defined but determined from measurements such as and anomalous magnetic moment of the , with its value refined in the 2022 CODATA adjustment to reflect improved experimental accuracy. Avogadro's constant, N_A, links the microscopic scale of atoms to the macroscopic scale of moles in chemistry, defining the number of constituent particles in one mole of substance. It is exactly N_A = 6.02214076 \times 10^{23} mol^{-1} , established in the 2019 SI revision to fix the mole independently of the kilogram. This exact value facilitates precise stoichiometry and supports advancements in and .
ConstantSymbolValueUnitsStatusRole
c299792458 s⁻¹ (defined )Basis for and meter
h$6.62607015 \times 10^{-34}J (defined 2019)Quantum scale and kilogram
\alpha$7.2973525643(11) \times 10^{-3}DimensionlessMeasured (2022 CODATA)Electromagnetic coupling strength
N_A$6.02214076 \times 10^{23}mol⁻¹ (defined 2019)Atomic scale to molar scale

Quantities in Measurements

In measurements at the atomic scale, the represents a fundamental associated with the , defined as the most probable distance between the proton and the in its . This quantity is approximately a_0 \approx 5.29 \times 10^{-11} m. It serves as a characteristic scale for atomic dimensions, influencing models in and where electron orbits or wavefunctions are approximated. At human scales, everyday measurements reflect biological norms and physical limits. The average adult is approximately 1.7 m, varying by population but providing a baseline for , , and health assessments. In terms of speed, the peak human running velocity record stands at about 12.4 m/s, achieved by during his 100-meter sprint in 2009, highlighting the biomechanical constraints of muscle power and . Energy measurements often employ convenient units derived from fundamental constants, such as the (), which quantifies energy on the scale. One equals $1.602 \times 10^{-19} J, corresponding to the gained by an accelerated through a potential difference of 1 volt. This unit is essential in and for expressing binding energies, wavelengths, and potentials without resorting to SI joules. Derived dimensionless quantities like the further illustrate how measurements integrate multiple physical parameters to characterize phenomena. In , the Re = \frac{\rho v L}{\mu}, where \rho is fluid , v is , L is a , and \mu is dynamic viscosity, predicts flow regimes—laminar for low Re (e.g., below 2000 in pipes) and turbulent for high Re (e.g., above 4000). For instance, blood flow in arteries typically yields Re around 1000–4000, aiding designs.

Numbers in Geographical and Astronomical Contexts

Planetary and Stellar Distances

In astronomy, planetary and stellar distances provide essential scales for understanding the vastness of space, often expressed in units like the (AU) for solar system measurements and light-years for interstellar and cosmic expanses. These distances highlight the progression from nearby celestial bodies to the structure of galaxies and the itself. The average distance from to defines the astronomical unit, which is approximately 149.6 million kilometers (or 92.96 million miles). This unit serves as a fundamental baseline for measuring distances within our solar system, such as the orbit of planets like Mars at about 1.52 AU or at 5.2 AU. Beyond the solar system, the light-year measures interstellar distances as the distance light travels in one Julian year, approximately $9.461 \times 10^{12} kilometers (or 5.879 × 10^{12} miles). The nearest star to our Sun, Proxima Centauri, lies at about 4.24 light-years away, underscoring the isolation of our stellar neighborhood. On galactic scales, the Milky Way galaxy has a diameter of roughly 100,000 light-years, encompassing hundreds of billions of stars within its spiral arms. The observable universe, limited by the speed of light and the age of the cosmos, extends to a radius of approximately 46.5 billion light-years from Earth, containing an estimated 2 trillion galaxies.

Earthly Geographical Scales

Earth's equatorial circumference measures approximately 40,075 kilometers, reflecting its oblate spheroid shape due to rotational forces, while the meridional circumference is slightly shorter at about 40,008 kilometers. This dimension establishes the baseline for terrestrial navigation and geospatial calculations, encompassing the planet's vast surface traversable by land and sea routes. In terms of vertical extremes, Earth's topography features profound contrasts between elevations and depressions. , the highest peak above , reaches 8,848.86 meters in the , a height confirmed through joint surveys by and using advanced GPS and leveling techniques. Conversely, the in the western plunges to the , the deepest known point at 10,935 meters below , measured via submersible transects and acoustic methods in recent expeditions. These extremes highlight the dynamic geological processes shaping , from tectonic uplift to zones. Continental landmasses further illustrate Earth's geographical scale, with dominating as the largest at approximately 44.5 million square kilometers, accounting for nearly 30% of global land area and encompassing diverse terrains from deserts to plateaus. Other continents, such as at 30.37 million square kilometers, provide comparative scales for understanding distribution and resource allocation. These areas, derived from satellite mapping and boundary delineations, underscore the uneven of land across the planet's surface.
FeatureApproximate ValueUnitNotes
Equatorial Circumference40,075kmBaseline for global circumnavigation
Mount Everest Height8,848.86mHighest point above sea level
Challenger Deep Depth10,935mDeepest ocean point
Asia Area44,500,000km²Largest continent

Named and Undefined Numbers

Named Large Finite Numbers

Named large finite numbers are specific integers, often extraordinarily vast, that have been given proper names by mathematicians or in historical contexts to denote scales beyond everyday comprehension. These names facilitate discussion of immense quantities in theoretical , such as in and , without requiring full . Unlike systematic notations like scientific prefixes, these named numbers arise from playful invention, problem-solving bounds, or ancient numeral systems, highlighting humanity's fascination with the within finite bounds. One of the earliest named large numbers is the myriad, derived from the term μυριάς (myrias), which precisely denoted . In classical and literature, such as in works by , the myriad served as a unit for higher multiples, like myriad-myriads for , reflecting the numerical scale of Hellenistic computations. This term entered Latin as myrias and later influenced European languages, though its strict meaning of 10,000 has largely faded in favor of denoting a vast, indefinite quantity. In the , the googol was coined as $10^{100}, a 1 followed by 100 zeros, by nine-year-old Milton Sirotta at the suggestion of his uncle, Edward , during a walk in 1920. Kasner popularized the term in his 1940 book Mathematics and the Imagination, co-authored with James R. Newman, to illustrate the concept of very large but finite numbers accessible to non-experts. Building on this, the googolplex extends the scale to $10^{\text{googol}}, or 1 followed by a googol zeros, which Kasner described as so immense that writing it out would exceed the observable universe's capacity to contain the digits. These names underscore the googol's role in demystifying for popular audiences. Far surpassing these, emerged as an upper bound in , a branch of studying order in large structures. Defined by mathematician in a 1971 paper co-authored with Bruce Rothschild, it addresses the minimal dimensions needed to ensure certain monochromatic subgraphs in hypercube colorings. The number is constructed using , starting with g_1 = 3 \uparrow\uparrow\uparrow\uparrow 3, defined as $3 \uparrow\uparrow\uparrow (3 \uparrow\uparrow\uparrow 3), where $3 \uparrow\uparrow\uparrow 3 is a power tower of 3's of height $3 \uparrow\uparrow 3 \approx 7.6 \times 10^{12}, and iterating g_{n+1} = 3 \uparrow^{g_n} 3 up to g_{64}, resulting in an incomprehensibly vast integer that dwarfs the . Graham's number, while an overestimate, provided a finite solution to the problem and popularized up-arrow notation for expressing hyper-operations. Even larger is TREE(3), a value from derived from a finite form of , which posits that sequences of labeled trees must contain embeddable subsequences. Introduced by logician Harvey Friedman in the early 1980s as part of studies, TREE(n) measures the longest sequence of n-colored trees where no later tree embeds an earlier one under homeomorphic embedding. For n=3, TREE(3) yields a finite but extraordinarily huge number, vastly exceeding , with growth so rapid that its exact value remains uncomputable in practice, though proofs of finiteness rely on up to the Bachmann-Howard ordinal. This function exemplifies the explosive growth in fast-growing hierarchies and has implications for proof theory's strength. While SI prefixes like yotta (10^24) name smaller large scales in scientific contexts, named numbers like these emphasize conceptual extremes in .

Constants Without Closed-Form Expressions

Constants without closed-form expressions are real numbers or functions defined through infinite processes, algorithmic definitions, or limiting behaviors that preclude a finite expression using elementary operations, roots, or standard transcendental functions. These constants often arise in , , and , where their values are intrinsically tied to undecidable problems or infinite summations/products over primes. They highlight fundamental limits in and , as their precise computation or simplification is either impossible or grows beyond any recursive bound. Chaitin's constant, denoted Ω, represents the halting probability of a universal prefix-free , specifically the probability that a randomly generated program (under a self-delimiting encoding) halts when executed. Formally, for a universal prefix-free machine U, \Omega = \sum_{p : U(p) \text{ halts}} 2^{-|p|}, where the sum is over all halting programs p, and |p| is the length of p in bits. This constant is uncomputable because determining whether a program halts is undecidable by the , and the first n bits of Ω encode solutions to all halting instances up to size n, making it algorithmically random with no . Its value lies between 0 and 1, but exact digits beyond a few approximations require solving vast numbers of halting problems. Introduced by , Ω exemplifies a whose expansion is incompressible and thus inherently complex. Busy beaver numbers, denoted BB(n), quantify the maximum productivity of an n-state, two-symbol that halts, typically measured as the maximum number of 1s written on the tape (or steps taken) before halting. The function BB(n) is defined as \text{BB}(n) = \max \{ \Sigma(M) \mid M \text{ is an } n\text{-state } 2\text{-symbol TM that halts} \}, where Σ(M) counts the 1s produced by machine M. This function grows faster than any , rendering it uncomputable and without a , as computing BB(n) would solve the for all n-state machines. Known values include BB(1) = 1, BB(2) = 4, BB(3) = 6, BB(4) = 13, and BB(5) = 47,176,870 (determined in 2024), with higher values eluding computation due to the explosion in machine count. Tibor Radó introduced this concept to illustrate non-computable functions in a concrete, maximal-growth setting. Kolmogorov complexity provides a measure of the intrinsic of an object x, defined as , the of the shortest program (in a fixed ) that outputs x. For a x, K(x) = \min \{ |p| \mid U(p) = x \}, where U is a prefix machine and p is the program. This measure lacks a because it depends on the undecidable to verify program correctness, and it is semi-computable but not computable. Unlike Shannon entropy, which averages over ensembles, Kolmogorov complexity applies to individual objects, capturing absolute randomness without probabilistic assumptions. For example, K of a random of n is approximately n bits, while for structured like primes, it can be much smaller. formalized this in his algorithmic approach to information, bridging and probability. Landau's totient constant arises in the asymptotic analysis of the summatory function of φ(n), which counts integers up to n coprime to n. Landau established that \sum_{k=1}^x \phi(k) = \frac{3}{\pi^2} x^2 + O(x \log x), where the leading constant \frac{3}{\pi^2} \approx 0.30396355 = \frac{1}{2 \zeta(2)}, with \zeta the and \zeta(2) = \prod_p \left(1 - \frac{1}{p^2}\right)^{-1} = \frac{\pi^2}{6}. This constant has no simpler closed form beyond the product, as it encodes the global distribution of primes via the Euler product for \zeta(2). The result quantifies the average density of totients, with the error term reflecting prime gaps. derived this in his work on multiplicative functions, providing a cornerstone for estimates.

References

  1. [1]
    [PDF] WHAT KIND OF NUMBER IS IT? ,
    Numbers can be natural, whole, integers, rational, irrational, or real. Natural numbers are for counting, and integers include negative numbers. Irrational  ...
  2. [2]
    [PDF] 1.1 The Real Number System
    A real number is either a rational or irrational number. A real number is positive if it is greater than 0, negative if it is less than 0.
  3. [3]
    [PDF] MA 15300 Lesson 1 Notes I REAL NUMBERS Natural Numbers
    Real numbers are any number that is rational or irrational. Natural numbers are 1, 2, 3, ... and whole numbers are natural numbers plus zero.
  4. [4]
    [PDF] Real Numbers and their Properties
    Real numbers (R) are points on a number line, and can be added, subtracted, multiplied, and divided. They follow closure, commutative, associative, and ...
  5. [5]
    (PDF) The Evolution of Numbers: From Ancient Counting Systems to ...
    Oct 19, 2024 · This paper explores the rich history of numbers, tracing their journey from ancient counting systems in Mesopotamia, Egypt, and India to the critical role they ...
  6. [6]
    Question Corner -- What are the Origins of Number Systems?
    Mar 17, 1997 · The origins of numbers date back to the Egyptians and Babylonians, who had a complete system for arithmetic on the whole numbers (1,2,3,4,. . . ) and the ...
  7. [7]
    COUNTING SYSTEMS AND NUMERALS - HistoryWorld
    Ten has been the basis of most counting systems in history. When any sort of record is needed, notches in a stick or a stone are the natural solution.
  8. [8]
    Zero - MacTutor History of Mathematics - University of St Andrews
    The Indians also used a place-value system and zero was used to denote an empty place. In fact there is evidence of an empty place holder in positional numbers ...
  9. [9]
    What Is the Origin of Zero? | Scientific American
    Jan 16, 2007 · The first evidence we have of zero is from the Sumerian culture in Mesopotamia, some 5,000 years ago. There, a slanted double wedge was inserted ...
  10. [10]
    [PDF] 4 Number Theory I: Prime Numbers - Penn Math
    A natural number larger than 1 is called prime if it can be evenly divided only by 1 and itself; other natural numbers greater than 1 are called composite. To ...
  11. [11]
    [PDF] BASIC PROPERTIES OF NUMBERS 1. Natural Numbers First we ...
    Sep 24, 2016 · Natural numbers have addition and multiplication, closure, identity, commutativity, associativity, distributivity, cancellation, and ordering  ...
  12. [12]
    Perfect numbers - MacTutor History of Mathematics
    A perfect number is defined to be one which is equal to the sum of its aliquot parts. The four perfect numbers 6, 28, 496 and 8128 seem to have been known from ...
  13. [13]
    Peano's axioms in their historical context | Archive for History of ...
    Segre, M. Peano's axioms in their historical context. Arch. Hist. Exact Sci. 48, 201–342 (1994). https://doi.org/10.1007/BF00375085Missing: original | Show results with:original
  14. [14]
    Disquisitiones arithmeticae : Gauss, Carl Friedrich, 1777-1855
    Aug 11, 2018 · Disquisitiones arithmeticae. by: Gauss, Carl Friedrich, 1777-1855. Publication date: 1801. Topics: Number theory. Publisher: Lipsiae : In ...
  15. [15]
    Sources of the Seven-Day Week - NASA ADS
    The Sumerians used seven in exorcising evils and favored seven in temple offerings. Seven-day period's are mentioned also in the Assyro-Babylonian epic of ...
  16. [16]
    The Musical Origin of the Seven-Day Week - Language Log
    May 6, 2022 · The days of the week are named after the seven classical planets and ordered using the musical sequence 4,1,5,2,6,3,7.
  17. [17]
    Calendars and their History - NASA Eclipse
    Oct 14, 2012 · The year is divided into twelve formalized months that were eventually adopted for the Gregorian calendar. 8.2 History of the Julian Calendar.Missing: zodiac | Show results with:zodiac
  18. [18]
    Zodiac • Encyclopaedia Britannica, 1911 - Sir Thomas Browne
    The twelvefold division of the zodiac was evidently suggested by the occurrence of twelve full moons in successive ...
  19. [19]
    Number Math - Berkeley Learning Platform
    The Egyptians employed a base-10 system, similar to our modern decimal system, but with distinct symbols for each power of ten. Their number system played a ...
  20. [20]
  21. [21]
    Who Discovered the Binary System and Arithmetic? Did Leibniz ...
    Gottfried Wilhelm Leibniz (1646-1716) is the self-proclaimed inventor of the binary system and is considered as such by most historians of mathematics.
  22. [22]
    [PDF] leibniz.pdf - UMD Department of Computer Science
    Leibniz designed a calculating machine based on binary. He did not give much detail about his device. However, (1) it was hard to build and never was built ...
  23. [23]
    Why is 13 considered unlucky? - University of South Carolina
    Jan 9, 2023 · The number 13 may be associated with some famous but undesirable dinner guests. In Norse mythology, the god Loki was 13th to arrive at a feast ...Missing: taboo | Show results with:taboo
  24. [24]
    Ideas of luck and superstition vary among cultures around the world
    Jul 12, 2018 · The fear of the number 13 is so pervasive in Western culture that many hotels and offices omit the 13th floor, he noted, while airports in ...Missing: taboo origins
  25. [25]
    The Number Four Kills Again… Or Does It? - McGill University
    Jan 4, 2018 · ... unlucky birth years), all of the ICD codes for heart disease were used. In the Baskerville effect study, however, Phillips only used 5 of ...<|separator|>
  26. [26]
    Euclid's Elements, Book VII, Definitions 11 through 14
    A prime number is that which is measured by a unit alone. 12. Numbers relatively prime are those which are measured by a unit alone as a common measure. 13.
  27. [27]
    Prime Number -- from Wolfram MathWorld
    Euclid's second theorem demonstrated that there are an infinite number of primes. However, it is not known if there are an infinite number of primes of the form ...<|control11|><|separator|>
  28. [28]
    The First 10000 Primes
    The First 10,000 Primes (the 10,000th is 104,729) For more information on primes see https://t5k.org/ 2 3 5 7 11 13 17 19 23 29 31 37 41 43 47 53 59 61 67 ...
  29. [29]
    Euclid's Theorems -- from Wolfram MathWorld
    Euclid's second theorem states that the number of primes is infinite. This theorem, also called the infinitude of primes theorem, was proved by Euclid in ...
  30. [30]
    [PDF] Euclid's Infinity of Primes There are infinitely many prime numbers.
    A prime number is an integer greater than one which cannot be divided exactly by any other integer greater than one. Euclid's proof, well over.
  31. [31]
    Twin Primes -- from Wolfram MathWorld
    It is conjectured that every even number is a sum of a pair of twin primes except a finite number of exceptions whose first few terms are 2, 4, 94, 96, 98, 400, ...
  32. [32]
    Twin Prime Conjecture -- from Wolfram MathWorld
    The twin prime conjecture states that there are an infinite number of pairs of twin primes, though it is not yet proven. There are two related conjectures.
  33. [33]
    Goldbach Conjecture -- from Wolfram MathWorld
    Goldbach's original conjecture (sometimes called the "ternary" Goldbach conjecture), written in a June 7, 1742 letter to Euler, states "at least it seems ...Missing: source | Show results with:source
  34. [34]
    Highly Composite Number -- from Wolfram MathWorld
    Highly composite numbers are numbers such that divisor function d(n)=sigma_0(n) (i.e., the number of divisors of n) is greater than for any smaller n.
  35. [35]
    [PDF] Highly composite numbers - Srinivasa Ramanujan
    The numbers marked with the asterisk in the table are called superior highly composite numbers. Their definition and properties will be found in §§ 32, 33. dd(N).
  36. [36]
    Superior Highly Composite Number -- from Wolfram MathWorld
    A superior highly composite number is a positive integer n for which there is an e>0 such that (d(n))/(n^e)>=(d(k))/(k^e) for all k>1.
  37. [37]
    Polygonal and Figurate Numbers - Dr Ron Knott
    The series above is part of the series of highly composite numbers, where each has more factors than any smaller number: 2, 4, 6, 12, 24, 36, 48, 60, 120, ...
  38. [38]
    [PDF] highly-composite-numbers.pdf - tom rocks maths
    A composite number is simply a positive integer with at least one divisor other than 1 and itself. This means that it is possible for a composite number to have ...
  39. [39]
    Perfect Number -- from Wolfram MathWorld
    Perfect numbers are positive integers n such that n=s(n), (1) where s(n) is the restricted divisor function (i.e., the sum of proper divisors of n), ...
  40. [40]
    Mathematicians Open a New Front on an Ancient Number Problem
    Sep 10, 2020 · “And if odd perfect numbers don't exist, the study of odd spoof numbers might lead to a proof of their nonexistence.” Other OPN experts, ...
  41. [41]
    Odd Perfect Number -- from Wolfram MathWorld
    Norton (1960) showed that odd perfect numbers not divisible by 3 or 5, it must have at least 15 distinct prime factors.
  42. [42]
    Amicable Pair -- from Wolfram MathWorld
    An amicable pair (m,n) consists of two integers m,n for which the sum of proper divisors (the divisors excluding the number itself) of one number equals the ...
  43. [43]
    (PDF) BRAHMAGUPTA AND THE CONCEPT OF ZERO
    This paper attempts to reconstruct the possible reasoning process that led the Indian mathematician Brahmagupta in 628 AD to the formulation of two ...
  44. [44]
    The Origin of Zero | Scientific American
    Aug 21, 2009 · Initially, zero functioned as a mere placeholder—a way to tell 1 from 10 from 100, to give an example using Arabic numerals. "That's not a full ...
  45. [45]
    Division by Zero -- from Wolfram MathWorld
    Division by zero is the operation of taking the quotient of any number x and 0, ie, x/0. The uniqueness of division breaks down when dividing by zero.
  46. [46]
  47. [47]
    [PDF] What's so Baffling About Negative Numbers? - Applied Mathematics
    An explicit interpretation of negatives as coordinates on a number line occurs later in the work of the 12th century Bhaskara II (so-called to distinguish him ...
  48. [48]
    History of Negative Numbers
    Descartes partially accepted negative numbers. He rejected negative roots of equations as "false", since they represented numbers less than nothing. However, he ...
  49. [49]
    The History of Negative Numbers | NRICH
    Feb 1, 2011 · Negative numbers did not appear until about 620 CE in the work of Brahmagupta (598 - 670) who used the ideas of 'fortunes' and 'debts' for positive and ...
  50. [50]
    Absolute Value - Ximera - The Ohio State University
    To do this, the absolute value function just returns nonnegative numbers unharmed, and makes negative numbers turn positive. Algebraically, this is ...
  51. [51]
    1.7 Negative numbers and the use of brackets - The Open University
    Numbers smaller than zero (shown to the left of zero on the number line in the figure below) are called negative numbers. We indicate they are negative by ...
  52. [52]
    [PDF] The International System of Units (SI)
    These SI prefixes refer strictly to powers of 10. They should not be used to indicate powers of 2 (for example, one kilobit represents 1000 bits.
  53. [53]
    SI prefixes - BIPM
    Decimal multiples and submultiples of SI units ; tera. T · 10 ; giga. G · 10 ; mega. M · 10 ; kilo. k. 10 ...
  54. [54]
    2022-12-19-si-prefixes - BIPM
    Dec 19, 2022 · The prefixes for multiples of the units have been extended from 1024, yotta, by adding ronna (symbol R) for 1027 and quetta (symbol Q) for 1030.
  55. [55]
    GOOGOL Definition & Meaning - Merriam-Webster
    Sep 26, 2025 · The meaning of GOOGOL is the figure 1 followed by 100 zeros equal to 10100. Google vs. Googol Did you know?
  56. [56]
    googol - PlanetMath.org
    Mar 22, 2013 · The googol was created by the American mathematician Edward Kasner (1878-1955) [4] in [2] to illustrate the difference between an ...Missing: origin | Show results with:origin
  57. [57]
    Egyptian Fraction -- from Wolfram MathWorld
    An Egyptian fraction is a sum of positive (usually) distinct unit fractions. The famous Rhind papyrus, dated to around 1650 BC contains a table of ...
  58. [58]
  59. [59]
    Decimal Expansion -- from Wolfram MathWorld
    1/7, 0.142857^_, 1/(17), 0.0588235294117647^_. 1/8, 0.125, 1/(18), 0.05^_. 1/9, 0.1^_, 1/(19), 0.052631578947368421^_. 1/(10), 0.1, 1/(20), 0.05. If r=p/q has a ...
  60. [60]
    One-Seventh Ellipse -- from Wolfram MathWorld
    Consider the decimal expansion of the reciprocal of the number seven, 1/7=0.142857142857...=0.142857^_, which is a repeating decimal.
  61. [61]
    Harmonic Number -- from Wolfram MathWorld
    A harmonic number arises from truncating the harmonic series, and is given by H_n = 1/n + H_(n-1), with H_1 = 1.
  62. [62]
    How Archimedes showed that $π$ is approximately equal to 22/7
    Aug 18, 2020 · In this article, we discuss how Archimedes came up with his formula. Archimedes in fact proved that 223/71 < \pi < 22/7.Missing: source | Show results with:source
  63. [63]
    [PDF] zu chongzhi (429 - University of St Andrews
    It was reprinted again in 1084, but all traces of the book were lost in the 12th century. In his book ZU CHONGZHI gave the approximation. 113. 355 for the ...
  64. [64]
    e Continued Fraction -- from Wolfram MathWorld
    This continued fraction is sometimes known as Euler's continued fraction. A plot of the first 256 terms of the continued fraction represented as a sequence of ...
  65. [65]
    Golden Ratio -- from Wolfram MathWorld
    The golden ratio, also known as the divine proportion, golden mean, or golden section, is a number often encountered when taking the ratios of distances in ...
  66. [66]
    Algebraic Number -- from Wolfram MathWorld
    An algebraic number is a root of a nonzero polynomial equation with integer coefficients, and it can be real or complex, like sqrt(2) or i.
  67. [67]
    Minimal Polynomial -- from Wolfram MathWorld
    ... minimal polynomial of the golden ratio phi (as well as the the golden ratio conjugate -phi^(-1)) since, via the quadratic equation, the roots of x^2-x-1 are ...
  68. [68]
    Constructible Number -- from Wolfram MathWorld
    A number which can be represented by a finite number of additions, subtractions, multiplications, divisions, and finite square root extractions of integers.
  69. [69]
    [PDF] Lecture 7 - Math 5111 (Algebra 1)
    Oct 1, 2020 · The set of constructible real numbers is a field, and we can also ... n - 1 over any field has a splitting field extension of degree at.
  70. [70]
    Recounting the History of Math's Transcendental Numbers
    Jun 27, 2023 · Transcendental numbers are non-algebraic values, not roots of polynomials with integer coefficients. They are plentiful, difficult to find, and ...
  71. [71]
    [PDF] TRANSCENDENCE OF e - KEITH CONRAD
    Such sums are not as mathematically interesting as π or e. • In 1873, Hermite [4] proved e is transcendental. • In 1874, Cantor [1] published his first paper on ...
  72. [72]
    The Number e and the Exponential Function - Galileo
    ... in this way, so as n goes to infinity this simple series represents the limiting value of (1+1n)n. Therefore, e must be just the sum of this infinite series.
  73. [73]
    e -- from Wolfram MathWorld
    |e-p/q|>1/((S(q)+1. (6). It is not known if pi+e or pi/e is irrational. It is known that pi+e and pi/e do not satisfy any polynomial equation of degree <=8 ...
  74. [74]
    Apéry's Constant -- from Wolfram MathWorld
    (OEIS A002117) where zeta(z) is the Riemann zeta function. Apéry (1979) proved that zeta(3) is irrational, although it is not known if it is transcendental.Missing: transcendence | Show results with:transcendence
  75. [75]
    Irrational Number -- from Wolfram MathWorld
    Irrational numbers have decimal expansions that neither terminate nor become periodic. Every transcendental number is irrational. There is no standard notation ...
  76. [76]
    Skewes Number -- from Wolfram MathWorld
    The Skewes number (or first Skewes number) is the number Sk_1 above which pi(n)<li(n) must fail (assuming that the Riemann hypothesis is true)
  77. [77]
    A lower bound for r(5, 5) - Wiley Online Library
    A &-free two-coloring of the edges of K42 is described, establishing 43 as a lower bound for R(5,5). The Ramsey Number, R(s, t), is defined as the smallest ...
  78. [78]
    Most accurate value of pi | Guinness World Records
    02 April 2025. The most accurate value of pi is 300,000,000,000,000 digits, and was achieved by Linus Media Group (Canada) and KIOXIA (USA) on 2 April 2025.
  79. [79]
    Complex Number -- from Wolfram MathWorld
    Complex numbers are useful abstract quantities that can be used in calculations and result in physically meaningful solutions.
  80. [80]
    Gaussian Integer -- from Wolfram MathWorld
    A Gaussian integer is a complex number a+bi where a and b are integers. The Gaussian integers are members of the imaginary quadratic field Q(sqrt(-1)) and form ...Missing: authoritative source
  81. [81]
    [PDF] THE GAUSSIAN INTEGERS Since the work of Gauss, number ...
    The example we will look at here is the Gaussian integers, introduced by Gauss in 1832 [1]: Z[i] = {a + bi : a, b ∈ Z}.Missing: authoritative | Show results with:authoritative
  82. [82]
    Euler Formula -- from Wolfram MathWorld
    The Euler formula, sometimes also called the Euler identity (eg, Trott 2004, p. 174), states e^(ix)=cosx+isinx, where i is the imaginary unit.
  83. [83]
    [PDF] ON QUATERNIONS By William Rowan Hamilton
    In the theory which Sir William Hamilton submitted to the Academy in November,. 1843, the name quaternion was employed to denote a certain quadrinomial ...
  84. [84]
    Lectures on quaternions : containing a systematic statement of a ...
    Jun 8, 2009 · A systematic statement of a new mathematical method, of which the principles were communicated in 1843 to the Royal Irish academy.
  85. [85]
    [PDF] Application of Quaternions to Computation with Rotations
    For computation with rotations, quaternions offer the advantage of requiring only 4 numbers of storage, com- pared with 9 numbers for orthogonal matrices.
  86. [86]
    [PDF] Quaternions and Rotations∗
    Sep 10, 2013 · A quaternion is a 4-tuple, a concise representation of rotations in R3, defined as a sum of a scalar and a vector.
  87. [87]
    The Peculiar Math That Could Underlie the Laws of Nature
    Jul 20, 2018 · New findings are fueling an old suspicion that fundamental particles and forces spring from strange eight-part numbers called “octonions.”
  88. [88]
    Cayley-Dickson construction in nLab
    Sep 18, 2024 · The diagram on the right shows a basis of imaginary octonions obtained via Cayley-Dickson doubling from a standard basis { i , j , k } \{i,j,k\} ...Idea · Definition · Properties · Examples
  89. [89]
    Cardinal Number -- from Wolfram MathWorld
    A cardinal number (also called "the cardinality") is a type of number defined in such a way that any method of counting sets using it gives the same result.Missing: source | Show results with:source
  90. [90]
    Aleph-0 -- from Wolfram MathWorld
    The set theory symbol aleph_0 refers to a set having the same cardinal number as the "small" infinite set of integers.
  91. [91]
    Cantor Diagonal Method -- from Wolfram MathWorld
    N is the set of natural numbers, is larger than the cardinality aleph_0 of N . By applying this argument infinitely many times to the same infinite set, it ...Missing: source | Show results with:source
  92. [92]
    Cantors 1891 Diagonal Proof - English Translation - Logic
    Bd. I, S. pp 75-78 (1891). This is the basis for the Diagonal proof and for the Power Set Theorem. The original German text of Cantor's proof is also included ...
  93. [93]
    The Consistency of the Axiom of Choice and of the Generalized ...
    Gödel,. The Consistency of the Axiom of Choice and of the Generalized Continuum-Hypothesis, Proc. Natl. Acad. Sci. U.S.A. 24 (12) 556-557, https://doi.org ...
  94. [94]
    THE INDEPENDENCE OF THE CONTINUUM HYPOTHESIS - PNAS
    This is the first of two notes in which we outline a proof of the fact that the Con- tinuum Hypothesis cannot bederived from the other axioms of set theory, ...
  95. [95]
    [PDF] Ordinal numbers - Cornell Mathematics
    x<y ⇐⇒ x ∈ y. The finite ordinals (i.e., the ordinals that are finite sets) are precisely the natural numbers. The first infinite ordinal is the set N.
  96. [96]
    Transfinite Induction -- from Wolfram MathWorld
    Transfinite induction, like regular induction, is used to show a property P(n) holds for all numbers n. The essential difference is that regular induction ...
  97. [97]
    [PDF] ord-arithmetic.1 Ordinal Addition - Open Logic Project Builds
    Ordinal addition is not commutative; 1 + ω = ω<ω + 1. Proof. Note that 1 + ω = lsubn<ω(1 + n) = ω ∈ ω ∪ {ω} = ω+ = ...
  98. [98]
    [PDF] A zoo of ordinals
    The Church-Kleene ordinal ωCK. 1. : the smallest admissible ordinal > ω. This is the smallest ordinal which is not the order type of a recursive (equivalently ...
  99. [99]
    [PDF] 2022 codata recommended values of the fundamental constants of ...
    2022 CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL. CONSTANTS OF PHYSICS ... ∗speed of light in vacuum c. 299 792 458. m s−1. ∗Planck constant h. 6.626 ...<|separator|>
  100. [100]
    [PDF] CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL ...
    CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL PHYSICAL CONSTANTS: 2022 ... speed of light in vacuum c. 299 792 458 (exact). m s−1 muon g-factor −2(1 + ...
  101. [101]
    Bohr radius - CODATA Value
    Bohr radius $a_0$ ; Numerical value, 5.291 772 105 44 x 10-11 m ; Standard uncertainty, 0.000 000 000 82 x 10-11 m.
  102. [102]
    Human Height - Our World in Data
    The standard deviation was 7.59 cm. This means 68% of men were between 170.8 and 186 cm tall; 95% were between 163.2 and 193.6 cm. Women were smaller on ...
  103. [103]
    Usain Bolt record collection: the sprint king's greatest hits
    Oct 12, 2020 · At the record-winning event, Usain Bolt's average ground speed was 37.58km/h, whilst reaching a top speed of 44.72km/h in the 60-80m stretch ...
  104. [104]
    electron volt - CODATA Value
    Click symbol for equation. electron volt eV. Numerical value, 1.602 176 634 x 10-19 J. Standard uncertainty, (exact). Relative standard uncertainty, (exact).
  105. [105]
    Reynolds Number
    The Reynolds number is the ratio of inertial to viscous forces, and it is a dimensionless number. High values mean viscous forces are small.
  106. [106]
  107. [107]
    How big is Earth? | Space
    Jul 6, 2021 · The equatorial circumference of Earth is 24,901 miles (40,075 km). However, from the North Pole to the South Pole, the so-called meridional ...
  108. [108]
    Mt Everest grows by nearly a metre to new height - BBC
    Dec 8, 2020 · The new height is 8,848.86m (29,032 ft). China's previous official measurement of 8,844.43m had put the mountain nearly four metres lower than ...
  109. [109]
    Revised depth of the Challenger Deep from submersible transects
    For the Challenger Deep, the deepest observed seafloor depth was 10,935 m (±6 m at 95% C.I.) below mean sea level. For this work, the error term is dominated by ...
  110. [110]
    Maps, Facts and Geography of Asia - Worldatlas.com
    The largest continent on Earth, Asia's total size is roughly 17,212,048 square miles (44,579,000 square kilometers), or 30% of the planet's total landmass.
  111. [111]
    Myriad - Etymology, Origin & Meaning
    Myriad originates from Greek myrias meaning "ten thousand" or "countless," later from Latin and French; it means a large, indefinite number or countless ...
  112. [112]
    Edward Kasner (1878 - 1955) - Biography - MacTutor
    The name "googol" was invented by a child (Dr Kasner's nine-year-old nephew) who was asked to think up a name for a very big number, namely, 1 with a hundred ...
  113. [113]
    Graham's Number -- from Wolfram MathWorld
    is the so-called Knuth up-arrow notation. g_(64) is often cited as the largest number that has ever been put to practical use (Exoo 2003). In chained arrow ...Missing: origin | Show results with:origin
  114. [114]
    Too big to write but not too big for Graham | plus.maths.org
    Sep 4, 2014 · Big numbers have always been a part of Ramsey theory but in 1971 mathematician Ronald Graham came up with a number that dwarfed all before it.<|control11|><|separator|>
  115. [115]
    What's so special about Kruskal's theorem and the ordinal Γo? A ...
    This paper consists primarily of a survey of results of Harvey Friedman about some proof-theoretic aspects of various forms of Kruskal's tree theorem.
  116. [116]
    Wrap Your Head Around the Enormity of the Number TREE(3)
    Oct 20, 2017 · For example, American mathematician Joseph Kruskal proved that any TREE(n) will ultimately result in a tree that contains a previous tree, ...
  117. [117]
    [PDF] Randomness and Mathematical Proof - Gwern
    5 (May 1975), pp. 47-52 by Gregory J. Chaitin. Although randomness can be precisely defined and can even be measured, a given number cannot be proved to be ...
  118. [118]
    [PDF] On Non-Computable Functions - Gwern.net
    The construction of non-computable functions used in this paper is based on the principle that a finite, non-empty set of non-negative integers has a.
  119. [119]
    [PDF] Three approaches to the quantitative definition of information
    Dec 21, 2010 · There are two common approaches to the quantitative definition of. "information": combinatorial and probabilistic. The author briefly de-.
  120. [120]
    Lösung des Lehmer'schen Problems - jstor
    Er vermutete nun allgemein, dass der Grenzwert existiert und > 0 ist. 1 "cAsymptotic Evaluation of certain Totient Sums" tAMIERICAN JOURNAL OF MATHEMATICS, Bd.