Fact-checked by Grok 2 weeks ago

Rational number

In , a rational number is any number that can be expressed as the of two s p and q, where p is the numerator and q (the denominator) is a non-zero . This representation allows rational numbers to include all s (as \frac{p}{1}), proper and improper fractions (such as \frac{3}{4} or \frac{5}{3}), and their equivalents in decimal form, which either terminate (like 0.75 for \frac{3}{4}) or repeat periodically (like 0.204545... for \frac{9}{44}). The set of all rational numbers is denoted by \mathbb{[Q](/page/Q)}. Rational numbers form a under the standard operations of and , meaning they are closed under these operations, support additive and multiplicative inverses (except zero has no multiplicative inverse), and satisfy the distributive, associative, and commutative . For example, addition of two rationals \frac{a}{b} + \frac{c}{d} yields \frac{ad + bc}{bd}, multiplication gives \frac{a}{b} \times \frac{c}{d} = \frac{ac}{bd}, and division (by a non-zero rational) is handled via the . Subtraction follows from addition by inverses, ensuring all basic arithmetic is well-defined within \mathbb{Q}. These make rational numbers for exact computations in fractions and proportions, extending the integers \mathbb{Z} to handle division comprehensively. As a subset of the real numbers \mathbb{R}, the rational numbers are dense in \mathbb{R}, meaning that between any two distinct reals, there exists a rational number, which underscores their utility in approximations and limits. However, \mathbb{Q} is countable, unlike the uncountable reals, allowing it to be listed in a despite its infinite extent. Rational numbers exclude numbers like \sqrt{2} or \pi, which cannot be expressed as finite or repeating decimals or integer ratios. This distinction highlights the foundational role of rationals in and , where they serve as the building blocks for more complex structures.

Definition and Terminology

Definition

A rational number is any number that can be expressed as the quotient or fraction \frac{p}{q}, where p and q are integers and q \neq 0. This definition presupposes the existence of the integers, which form the building blocks for constructing through division. Rationals are commonly denoted using fractional notation \frac{a}{b}, where the fraction is typically reduced to lowest terms by dividing both numerator and denominator by their , ensuring \gcd(a, b) = 1 and b > 0. Examples include \frac{1}{2} (half), -\frac{3}{4} (negative three-quarters), and \frac{5}{1} = 5 (showing that all integers are rational numbers). The set of rational numbers forms a proper of the real numbers, excluding irrational numbers that cannot be expressed as such ratios. The concept of ratios underlying rational numbers was recognized in ancient civilizations, such as the who represented fractions as sums of unit fractions for practical computations.

Terminology and Etymology

The term "rational number" derives from the Latin adjective rationalis, rooted in ratio, which means "reason," "," or "proportion." This etymology underscores the conceptual link to proportional reasoning, as rational numbers are fundamentally ratios of two integers, a idea central to ancient mathematical discourse on commensurability. In the evolution of mathematical terminology, "rational" emerged to differentiate numbers expressible as ratios from those that are not, such as roots. Ancient mathematicians like , in his Elements (circa 300 BCE), developed the theory of ratios in Book V without using the modern term "rational," but his treatment of commensurable magnitudes—ratios of integers—provided the foundation for later distinctions between rational and quantities. The explicit use of "rational" in this context appeared in translations of , with "" first recorded in English mathematical texts around 1551 to contrast with expressible ratios. The key term "fraction," denoting a rational number less than one in its basic form, originates from the Latin fractio (a breaking), derived from the verb frangere (to break), evoking the division of a whole into parts. Within fractional representation, the "numerator" stems from Late Latin numerator (a counter), from numerus (number), signifying the quantity of parts selected, while the "denominator" comes from Medieval Latin denominator (one that names), from denominare (to name), indicating the total units into which the whole is divided. Distinctions like "proper fraction" (numerator smaller than denominator) and "improper fraction" (numerator greater than or equal to denominator) developed in the , with "improper" first appearing in English in 1542 in Robert Recorde's The Ground of Artes, reflecting the notion that such exceed the "proper" piece of a whole. A "mixed number," combining an with a proper (e.g., $3 \frac{1}{2}), arose as a practical notation for values greater than one, its terminology descriptively indicating a of whole and fractional components without a specific ancient Latin root. Fractional notation itself has ancient origins, with systematic use appearing in around the 7th century ; Brahmagupta's Brahmasphutasiddhanta (628 CE) presented fractions as one number above another, separated by a space, marking an early precursor to modern vinculum-barred forms. This convention spread through Islamic scholars like al-Hassâr in the 12th century, who introduced the , before full Western adoption in medieval via Fibonacci's works.

Representations

Fractional Representation

A rational number is commonly expressed in fractional form as \frac{p}{q}, where p and q are integers with q > 0 and the \gcd(p, q) = 1. This representation, known as the irreducible or lowest terms form, ensures the fraction is in its simplest state by eliminating any common factors between the numerator and denominator other than 1. To obtain this standard form from any given , the simplification process involves computing the GCD of the absolute values of the numerator and denominator using the and then dividing both by this value. For instance, consider \frac{4}{8}: the GCD of 4 and 8 is 4, so dividing yields \frac{4 \div 4}{8 \div 4} = \frac{1}{2}. This method guarantees the fraction is reduced, preserving the value while minimizing the integers involved. Sign conventions in fractional representations allow the negative sign to appear in either the numerator or denominator, but the standard practice is to place it in the numerator with a positive denominator to maintain . For example, \frac{-2}{3}, \frac{2}{-3}, and -\frac{2}{3} all represent the same rational number, but the form with q > 0 is preferred. Rational numbers greater than 1 in can also be written as mixed numbers, which combine a and a proper , such as $2 \frac{1}{2} for \frac{5}{2}. To convert a mixed number to an improper , multiply the by the denominator, add the numerator, and place the result over the denominator: $2 \frac{1}{2} = \frac{2 \times 2 + 1}{2} = \frac{5}{2}. Conversely, to convert an improper like \frac{5}{2} to mixed form, divide the numerator by the denominator to get the and . However, improper fractions are generally preferred in advanced for their compactness and ease in algebraic manipulations. Every nonzero rational number possesses a unique irreducible fractional representation with a positive denominator, providing a way to identify and compare them unambiguously.

Decimal Representation

The of a rational number \frac{p}{q} in lowest terms, where p and q are integers with q > 0, is either terminating or eventually repeating. A terminating decimal ends after a finite number of digits after the decimal point, while an eventually repeating decimal consists of a (possibly empty) non-repeating prefix followed by a repeating sequence of digits that continues indefinitely. This dichotomy distinguishes rational numbers from irrationals, whose decimal expansions are non-terminating and non-repeating. A decimal expansion terminates if and only if the denominator q, after simplification, has no prime factors other than 2 and/or 5. In such cases, the can be rewritten with a denominator that is a by multiplying numerator and denominator by appropriate powers of 2 or 5 to balance the factors. For example, \frac{1}{2} = 0.5 terminates after one digit, as $2 = 2^1 requires multiplying by $5^1 to get denominator 10; similarly, \frac{1}{4} = \frac{1}{2^2} = 0.25, where multiplying by $5^2 = 25 yields denominator $100 = 10^2. The number of decimal places equals the maximum of the exponents of 2 and 5 in q. For all other rational numbers, where q has at least one prime factor different from 2 or 5, the decimal expansion is eventually repeating. The repeating part, known as the repetend, begins after a non-repeating prefix whose length is determined by the highest power of 2 or 5 dividing q. If q is coprime to 10 (no factors of 2 or 5), the expansion is purely repeating with no non-repeating part; otherwise, it is mixed or eventually repeating. For instance, \frac{1}{3} = 0.\overline{3} is purely repeating with period 1, while \frac{1}{6} = 0.1\overline{6} has a non-repeating digit "1" (length 1, matching the power of 2 in 6=2·3) followed by a repeating "6". Another example is \frac{1}{7} = 0.\overline{142857}, a pure repetend of length 6, and \frac{1}{12} = 0.08\overline{3}, mixed with non-repeating "08" (length 2, from 12=2^2·3) and repeating "3". To obtain the decimal expansion, perform of p by q, which generates digits sequentially; the process terminates if a remainder of occurs, or repeats when a repeats, signaling the start of the repetend. The length of the repetend, or period, for \frac{1}{q} (with q coprime to 10) is the multiplicative order of 10 q, the smallest positive k such that $10^k \equiv 1 \pmod{q}. This order divides \phi(q), , and for general \frac{p}{q}, the period is the order the part of q coprime to 10. For example, the period of \frac{1}{7} is 6, as 6 is the smallest k with $10^6 \equiv 1 \pmod{7}. Except for terminating decimals, every rational number has a unique decimal expansion. Terminating decimals admit a dual representation: the finite form and an equivalent repeating form ending in 9s. For example, $0.5 = 0.4\overline{9}, since both equal \frac{1}{2}. This non-uniqueness arises because the repeating 9s expansion corresponds to the of approaching the terminating value from below. In standard convention, the terminating form is preferred for simplicity.

Continued Fraction Representation

A continued fraction provides an alternative representation for rational numbers as a finite expression of the form [a_0; a_1, a_2, \dots, a_n], where a_0 is a non-negative and a_1, \dots, a_n are positive . This notation denotes the value a_0 + \cfrac{1}{a_1 + \cfrac{1}{a_2 + \cfrac{1}{\ddots + \cfrac{1}{a_n}}}}, offering a way to express rationals through nested fractions with terms. To construct the continued fraction expansion of a rational number p/q in lowest terms (with p, q > 0), apply the Euclidean algorithm: divide p by q to get quotient a_0 and remainder r_1, so p = a_0 q + r_1 with $0 \leq r_1 < q; then divide q by r_1 to get a_1 and r_2, continuing until a remainder of zero is reached, yielding the coefficients a_0, a_1, \dots, a_n. For example, $3/2 gives $3 = 1 \cdot 2 + 1, then $2 = 2 \cdot 1 + 0, so [1; 2]. Similarly, $22/7 yields $22 = 3 \cdot 7 + 1, then $7 = 7 \cdot 1 + 0, so [3; 7]. Every rational number has a unique finite continued fraction expansion under the convention that partial quotients are positive integers greater than or equal to 1, except that expansions ending with a_n = 1 admit an equivalent form [a_0; a_1, \dots, a_{n-1}, a_n] = [a_0; a_1, \dots, a_{n-1} + 1], allowing two representations in such cases. The partial quotients, or convergents, h_k / k_k (computed recursively via h_{-1} = 1, h_0 = a_0, h_k = a_k h_{k-1} + h_{k-2}; similarly for k_k with k_{-1} = 0, k_0 = 1), provide successive rational approximations to the number, with each improving upon the previous and the final convergent h_n / k_n equaling the exact rational. Unlike the infinite expansions for irrational numbers, rational numbers yield finite continued fractions, making this representation exact and terminating. Continued fractions are particularly valuable in Diophantine approximation, where the convergents yield the best rational approximations to a number by denominators up to a given size, though for rationals, the process terminates exactly rather than providing ongoing approximations.

Other Representations

Rational numbers can be represented in any integer base b \geq 2 using positional notation, where the expansion is either finite or eventually periodic, analogous to their decimal representations. For example, in binary (base 2), \frac{1}{2} = 0.1_2, which terminates, while \frac{1}{3} = 0.\overline{01}_2, which repeats. This property holds because the denominator of a reduced rational, when factored into primes, determines whether the expansion terminates (if all prime factors divide some power of b) or repeats otherwise. Another representation is the Egyptian fraction, where a positive rational \frac{p}{q} (with p, q positive integers) is expressed as a sum of distinct unit fractions, i.e., \frac{p}{q} = \sum_{i=1}^k \frac{1}{d_i} with distinct positive integers d_i. Every positive rational admits such a decomposition, as proven in modern times though known empirically to ancient Egyptians. The greedy algorithm computes one such expansion by repeatedly subtracting the largest possible unit fraction less than or equal to the remainder; for instance, \frac{2}{3} = \frac{1}{2} + \frac{1}{6}, obtained by taking \frac{1}{2} (the largest unit fraction ≤ \frac{2}{3}) and then applying the method to the remainder \frac{1}{3}. This method always terminates but may not yield the representation with the fewest terms. The Stern-Brocot tree provides an enumerative representation of all positive rationals, structured as an infinite binary tree where each level consists of reduced fractions in lowest terms, generated by starting with \frac{0}{1} and \frac{1}{0} (representing 0 and ∞) and iteratively inserting mediants \frac{a+c}{b+d} between adjacent fractions \frac{a}{b} and \frac{c}{d}. Every positive rational appears exactly once in reduced form in this tree, with no repetitions, offering a systematic way to list them without duplicates. Independently discovered by in 1858 and in 1861, the tree also encodes continued fraction expansions via left and right paths. Similarly, the Calkin-Wilf tree enumerates all positive rationals exactly once through a binary tree where the root is \frac{1}{1}, the left child of \frac{a}{b} is \frac{a}{a+b}, and the right child is \frac{a+b}{b}, with each node in reduced terms. This structure, introduced by and in 2000, corresponds bijectively to positive integers via breadth-first traversal, demonstrating the countability of the rationals. Rationals also appear in modular representations, such as on the projective line over the rationals \mathbb{P}^1(\mathbb{Q}), which consists of equivalence classes of points [x:y] in \mathbb{Q}^2 \setminus \{0\} under scalar multiplication by nonzero rationals, identifying the line with \mathbb{Q} \cup \{\infty\}. This compactifies the affine line of rationals by adding a point at infinity, useful in algebraic geometry for studying rational points and maps.

Arithmetic Operations

Equality and Canonical Forms

Two rational numbers expressed as fractions \frac{p_1}{q_1} and \frac{p_2}{q_2}, where p_1, p_2, q_1, q_2 are integers and q_1, q_2 \neq 0, are equal if and only if p_1 q_2 = p_2 q_1. This condition arises from the definition of rational numbers as equivalence classes of integer pairs, where equality holds when the cross products match. To see why this holds, suppose \frac{p_1}{q_1} = \frac{p_2}{q_2}. Multiplying both sides by q_1 q_2 yields p_1 q_2 = p_2 q_1, preserving equality since q_1 q_2 \neq 0. Conversely, if p_1 q_2 = p_2 q_1, then \frac{p_1}{q_1} - \frac{p_2}{q_2} = \frac{p_1 q_2 - p_2 q_1}{q_1 q_2} = 0, confirming the fractions represent the same rational. This cross-multiplication approach provides an exact test without requiring simplification of either fraction beforehand. The canonical form of a rational number is its unique irreducible fraction \frac{p}{q}, where p and q are integers with \gcd(p, q) = 1 and q > 0. To obtain this form, divide both numerator and denominator by their , then adjust the sign to ensure a positive denominator. For instance, \frac{2}{4} simplifies to \frac{1}{2} since \gcd(2, 4) = 2, while -\frac{2}{4} becomes -\frac{1}{2} to maintain the positive denominator. This representation ensures uniqueness within the of fractions denoting the same rational. To check equality algorithmically without full simplification, compute the integer products p_1 q_2 and p_2 q_1 and them directly, leveraging integer arithmetic. This method avoids the precision loss inherent in converting fractions to approximations for comparison, providing reliable results for exact rational identity as long as the products fit within the representation limits.

Ordering

The rational numbers form a totally ordered set under the standard ordering, where for any two rationals a and b, exactly one of a < b, a = b, or a > b holds, and the order is compatible with the field operations. Specifically, a < b if and only if b - a > [0](/page/0), where positive elements are those greater than . For fractions in with positive denominators, the order can be determined without by : for \frac{p_1}{q_1} and \frac{p_2}{q_2}, \frac{p_1}{q_1} < \frac{p_2}{q_2} if and only if p_1 q_2 < p_2 q_1. For example, to compare \frac{2}{3} and \frac{3}{4}, compute $2 \cdot 4 = 8 and $3 \cdot 3 = 9; since $8 < 9, it follows that \frac{2}{3} < \frac{3}{4}. The rational numbers satisfy the density property: between any two distinct rationals a < b, there exists another rational c such that a < c < b. One such c is the mediant \frac{p_1 + p_2}{q_1 + q_2} of \frac{p_1}{q_1} = a and \frac{p_2}{q_2} = b (with positive denominators), which lies strictly between them. The rationals also possess the Archimedean property: for any positive rationals a and b, there exists a positive integer n such that n a > b. This follows from the structure of the rationals, where the natural numbers are unbounded above. Unlike the real numbers, the rationals lack the least upper bound property: there exist nonempty subsets of the rationals that are bounded above but have no least upper bound in the rationals. For instance, the set of rationals whose square is less than 2 has upper bounds like 2 but no smallest such bound within the rationals.

Addition and Subtraction

To add two rational numbers expressed as fractions \frac{p_1}{q_1} and \frac{p_2}{q_2}, where p_1, q_1, p_2, q_2 \in \mathbb{Z} and q_1, q_2 \neq 0, compute \frac{p_1}{q_1} + \frac{p_2}{q_2} = \frac{p_1 q_2 + p_2 q_1}{q_1 q_2}. The resulting fraction should then be reduced to lowest terms by dividing numerator and denominator by their greatest common divisor (GCD). For example, \frac{1}{2} + \frac{1}{3} = \frac{1 \cdot 3 + 1 \cdot 2}{2 \cdot 3} = \frac{5}{6}, which is already in lowest terms. Subtraction follows analogously: \frac{p_1}{q_1} - \frac{p_2}{q_2} = \frac{p_1 q_2 - p_2 q_1}{q_1 q_2}, again reducing the result. For instance, \frac{3}{4} - \frac{1}{2} = \frac{3 \cdot 2 - 1 \cdot 4}{4 \cdot 2} = \frac{2}{8} = \frac{1}{4}. These operations preserve rationality, as the or of two is always rational, since the numerator and denominator remain integers. To perform addition or subtraction efficiently, especially with unequal denominators, convert the fractions to a common denominator using the least common multiple (LCM) of q_1 and q_2. The LCM can be computed via the formula \operatorname{LCM}(q_1, q_2) = \frac{|q_1 q_2|}{\gcd(q_1, q_2)}, where \gcd is the greatest common divisor found using the Euclidean algorithm: repeatedly replace the larger number by its remainder modulo the smaller until the remainder is zero, with the last non-zero remainder being the GCD. This minimizes intermediate values compared to using the product q_1 q_2 directly. Addition of rationals satisfies commutativity (\frac{p_1}{q_1} + \frac{p_2}{q_2} = \frac{p_2}{q_2} + \frac{p_1}{q_1}) and associativity (\left( \frac{p_1}{q_1} + \frac{p_2}{q_2} \right) + \frac{p_3}{q_3} = \frac{p_1}{q_1} + \left( \frac{p_2}{q_2} + \frac{p_3}{q_3} \right)). Every rational \frac{p}{q} has an -\frac{p}{q}, such that \frac{p}{q} + \left( -\frac{p}{q} \right) = 0. is defined as adding the additive inverse. In computational implementations using fixed-precision integers, and of rationals may lead to in the numerator or denominator during intermediate steps, such as ; however, using arbitrary-precision integers avoids this issue and ensures exact results.

Multiplication and Division

of two rational numbers \frac{p_1}{q_1} and \frac{p_2}{q_2}, where q_1 \neq 0 and q_2 \neq 0, is defined as \frac{p_1}{q_1} \times \frac{p_2}{q_2} = \frac{p_1 p_2}{q_1 q_2}. For example, \frac{2}{3} \times \frac{3}{4} = \frac{2 \times 3}{3 \times 4} = \frac{6}{12} = \frac{1}{2}, where the is simplified by dividing numerator and denominator by their . This operation satisfies several key properties: it is commutative, meaning \frac{p_1}{q_1} \times \frac{p_2}{q_2} = \frac{p_2}{q_2} \times \frac{p_1}{q_1}; associative, so (\frac{p_1}{q_1} \times \frac{p_2}{q_2}) \times \frac{p_3}{q_3} = \frac{p_1}{q_1} \times (\frac{p_2}{q_2} \times \frac{p_3}{q_3}); and distributive over , with \frac{p_1}{q_1} \times (\frac{p_2}{q_2} + \frac{p_3}{q_3}) = \frac{p_1}{q_1} \times \frac{p_2}{q_2} + \frac{p_1}{q_1} \times \frac{p_3}{q_3}. The multiplicative is \frac{1}{1}, since \frac{p}{q} \times \frac{1}{1} = \frac{p}{q}. Additionally, the rational numbers have no zero divisors other than zero itself: if \frac{p_1}{q_1} \times \frac{p_2}{q_2} = \frac{0}{1}, then either \frac{p_1}{q_1} = \frac{0}{1} or \frac{p_2}{q_2} = \frac{0}{1}. Division of rational numbers is defined as multiplication by the reciprocal: \frac{p_1}{q_1} \div \frac{p_2}{q_2} = \frac{p_1}{q_1} \times \frac{q_2}{p_2} = \frac{p_1 q_2}{q_1 p_2}, provided p_2 \neq 0 to ensure the reciprocal exists. For instance, \frac{3}{5} \div \frac{2}{7} = \frac{3}{5} \times \frac{7}{2} = \frac{21}{10}. To minimize computational errors, especially with large numerators and denominators, it is advisable to simplify by canceling common factors between the numerators and denominators before performing the multiplication, as this reduces the size of the intermediate products. In the earlier example, canceling the common factor of 3 in \frac{2}{3} \times \frac{3}{4} yields \frac{2}{4} = \frac{1}{2} directly.

Exponentiation to Integer Powers

For a rational number expressed in lowest terms as \frac{p}{q} where p and q are with q \neq 0, raising it to a positive power n > 0 is defined by repeated , yielding \left( \frac{p}{q} \right)^n = \frac{p^n}{q^n}. For example, \left( \frac{3}{2} \right)^2 = \frac{3^2}{2^2} = \frac{9}{4}. Any non-zero rational number raised to the power of zero equals 1, consistent with the general rule for exponentiation where the base is non-zero. For negative integer exponents, \left( \frac{p}{q} \right)^{-n} = \frac{1}{\left( \frac{p}{q} \right)^n} = \left( \frac{q}{p} \right)^n = \frac{q^n}{p^n} where n > 0 and p \neq 0. For instance, \left( \frac{2}{3} \right)^{-1} = \frac{3}{2}. The operation preserves rationality: the result of raising a non-zero rational to any integer power is always rational, as the numerator and denominator remain integers. Exponentiation to integer powers on the rationals obeys standard rules, such as (a^b)^c = a^{b c} and a^b \cdot a^c = a^{b+c} for compatible integer exponents b and c, where the base a is a non-zero rational.

Algebraic Properties

Field Structure

The rational numbers \mathbb{Q}, defined as the set of all fractions \frac{m}{n} where m, n \in \mathbb{Z} and n \neq 0, with equivalence \frac{m}{n} = \frac{p}{q} if mq = np, form a field under the standard operations of addition and multiplication inherited from the integers. These operations are defined as \frac{m}{n} + \frac{p}{q} = \frac{mq + np}{nq} and \frac{m}{n} \cdot \frac{p}{q} = \frac{mp}{nq}, and the field axioms are verified by reducing computations to integer arithmetic and leveraging the ring structure of \mathbb{Z}. More precisely, \mathbb{Q} is the field of fractions of the ring \mathbb{Z}, obtained by adjoining multiplicative inverses for all nonzero integers. The field axioms include closure under addition and multiplication (ensured by the definitions above), associativity and commutativity of both operations (inherited from \mathbb{Z}), the existence of additive and multiplicative identities (0 and 1, respectively), additive inverses ( -\frac{m}{n} = \frac{-m}{n} ), multiplicative inverses for nonzero elements ( \left( \frac{m}{n} \right)^{-1} = \frac{n}{m} if m \neq 0 ), and distributivity of multiplication over addition. All these properties hold without exception in \mathbb{Q}, distinguishing it from finite fields or rings without full inverses. The characteristic of \mathbb{Q} is 0, as the multiple n \cdot 1 = 1 + 1 + \cdots + 1 (n times) equals the integer n, which is nonzero for any positive integer n. \mathbb{Q} is also an ordered field, with the standard order \frac{m}{n} < \frac{p}{q} if mq < np (assuming positive denominators for simplicity), compatible with the operations: addition preserves order, and multiplication preserves order for positive elements. This makes \mathbb{Q} the smallest ordered field, as every ordered field contains a unique subfield isomorphic to \mathbb{Q}. The integers \mathbb{Z} embed as a of \mathbb{Q}, but \mathbb{Q} extends this by including multiplicative inverses for all nonzero integers, yielding a full structure that is unique up to among fields of characteristic 0. Specifically, \mathbb{Q} is the prime field of characteristic 0, the smallest containing an isomorphic copy of \mathbb{Z} and generated by 1 under operations.

Closure and Embeddings

The set of rational numbers \mathbb{Q} is closed under the operations of addition, subtraction, and multiplication, meaning that the result of any such operation on two rational numbers is again a rational number. It is also closed under division, provided the divisor is non-zero, as the quotient of two non-zero rationals can always be expressed as a ratio of integers. However, \mathbb{Q} is not closed under the extraction of roots; for instance, the square root of 2 is irrational and thus not in \mathbb{Q}. The integers \mathbb{Z} embed into \mathbb{Q} via the canonical map \phi: \mathbb{Z} \to \mathbb{Q} defined by \phi(n) = \frac{n}{1} for each n. This embedding is an injective ring homomorphism that preserves and , making \mathbb{Z} a of \mathbb{Q}. Consequently, every is a rational number, but the converse does not hold, as there exist rationals like \frac{1}{2} that are not . The identity map provides a natural embedding of \mathbb{Q} into itself, which is trivially bijective and preserves all field operations. \mathbb{Q} itself is not algebraically closed, as it lacks roots for certain polynomials with rational coefficients, such as x^2 - 2 = 0.

Countability

The set of rational numbers \mathbb{Q} is countable, meaning there exists a between \mathbb{Q} and the natural numbers \mathbb{N}, and thus |\mathbb{Q}| = \aleph_0. This was first established by in 1873. To prove the countability of \mathbb{Q}, first consider the positive rationals \mathbb{Q}^+. Each element can be uniquely represented as a reduced p/q where p and q are positive integers with \gcd(p, q) = 1. Group these fractions by the "height" k = p + q, starting from k = 2. For each fixed k, there are finitely many such pairs (p, q) with p + q = k and \gcd(p, q) = 1. Enumerate the fractions within each group in order of increasing p (or q), and traverse the groups in a zigzag pattern across the infinite grid of positive integers, skipping non-reduced fractions to ensure each rational appears exactly once. This process yields a from \mathbb{Q}^+ to \mathbb{N}. To extend to all of \mathbb{Q}, map zero to 0 (or include it separately), and handle negative rationals by pairing each positive rational with its negative counterpart, preserving countability since the union of two countable sets is countable. A formal relies on Cantor's , which establishes a \pi: \mathbb{N} \times \mathbb{N} \to \mathbb{N} given by \pi(m, n) = \frac{(m + n)(m + n + 1)}{2} + m. Since \mathbb{Q}^+ is in with the set of reduced pairs (p, q), which is a of \mathbb{N} \times \mathbb{N}, and subsets of countable sets are countable, \mathbb{Q}^+ is countable; the extension to \mathbb{Q} follows similarly. An explicit enumeration of the positive rationals can be obtained via the Calkin-Wilf tree, a where each node a/b (in lowest terms) has left child (a + b)/b and right child a/(a + b), with root $1/1; breadth-first traversal lists every positive rational exactly once. This structure, introduced by Neil Calkin and Herbert Wilf in , provides a recursive way to generate the without reducing fractions explicitly. The countability of \mathbb{Q} implies that it forms a countable dense of the real numbers \mathbb{R}, as every in \mathbb{R} contains infinitely many rationals, yet the total set remains enumerable.

Relations to Other Number Systems

Embedding in Real Numbers

The real numbers \mathbb{R} are constructed as the of the rational numbers \mathbb{Q} with respect to the , ensuring that every of rationals converges to a . One standard method defines \mathbb{R} as the set of classes of s of rationals, where two sequences are equivalent if their difference converges to zero. Another approach uses Dedekind cuts, partitioning the rationals into two nonempty subsets A and B such that all elements of A are less than all elements of B, no greatest element exists in A, and every corresponds to such a cut. These constructions embed \mathbb{Q} naturally into \mathbb{R} as a subfield, preserving the field operations and . This embedding is dense: for any two distinct real numbers x < y, there exists a rational q such that x < q < y. Consequently, every real number is the limit of a sequence of rationals, allowing \mathbb{Q} to approximate any element of \mathbb{R} arbitrarily closely. Algebraically, \mathbb{R} is a real closed field containing \mathbb{Q} as a subfield, where every positive element has a square root and every odd-degree polynomial over \mathbb{R} has a root. For instance, the irrational \sqrt{2} lies outside \mathbb{Q} but can be approximated by rationals via the convergents of its continued fraction expansion [1; 2, 2, 2, \dots], such as $3/2, $7/5, and $17/12, which satisfy |\sqrt{2} - p_n/q_n| < 1/(q_n q_{n+1}). Historically, these constructions addressed the incompleteness of \mathbb{Q}, exemplified by gaps like the absence of a supremum for the set \{q \in \mathbb{Q} \mid q^2 < 2\}, prompting 19th-century mathematicians to formalize \mathbb{R} for a complete ordered field.

Topological Properties

The rational numbers \mathbb{Q} inherit the subspace topology from the real numbers \mathbb{R}, meaning that a subset U \subseteq \mathbb{Q} is open if and only if there exists an open set V \subseteq \mathbb{R} such that U = V \cap \mathbb{Q}. This topology is metrizable, induced by the standard Euclidean metric d(a, b) = |a - b| for a, b \in \mathbb{Q}. A basis for the topology on \mathbb{Q} consists of sets of the form (p, q) \cap \mathbb{Q}, where p, q \in \mathbb{Q} and p < q. The metric space (\mathbb{Q}, d) is incomplete, as it contains Cauchy sequences that do not converge within \mathbb{Q}. For instance, the sequence of partial decimal approximations to \pi—3, 3.1, 3.14, 3.141, 3.1415, ...—is Cauchy in \mathbb{Q} because the terms become arbitrarily close, yet its limit \pi lies outside \mathbb{Q}. This incompleteness highlights the "gaps" in \mathbb{Q}, distinguishing it topologically from the complete space \mathbb{R}. The space \mathbb{Q} is totally disconnected, possessing no connected subsets with more than one point. For any two distinct points r_1, r_2 \in \mathbb{Q} with r_1 < r_2, an c \in (r_1, r_2) exists, and the sets (-\infty, c) \cap \mathbb{Q} and (c, \infty) \cap \mathbb{Q} are nonempty, disjoint, open in \mathbb{Q}, and their union contains \{r_1, r_2\}. Every open interval in \mathbb{Q} is itself disconnected, reflecting the dense interspersion of irrationals. As a countable metric space without isolated points—since every neighborhood of a rational contains infinitely many others due to the density of \mathbb{Q} in \mathbb{R}—\mathbb{Q} is not locally compact. Compact subsets of \mathbb{Q} must be finite, as any infinite subset has a limit point outside \mathbb{Q} or violates sequential compactness.

p-adic Numbers

In the context of rational numbers, the p-adic numbers arise from a different notion of "size" or valuation, distinct from the usual absolute value. For a fixed prime number p, the p-adic valuation v_p on the nonzero rationals \mathbb{Q}^\times is defined by writing a rational x = a/b in lowest terms and setting v_p(x) = v_p(a) - v_p(b), where v_p(n) for a nonzero integer n is the highest power of p dividing n, i.e., the exponent of p in its prime factorization. By convention, v_p(0) = +\infty. This valuation measures the extent to which p divides the numerator relative to the denominator, providing a way to quantify divisibility by powers of p. The p-adic metric is then derived from this valuation: for rationals x and y, the distance is d_p(x, y) = p^{-v_p(x - y)} if x \neq y, and d_p(x, x) = 0. This metric satisfies the ultrametric inequality d_p(x, z) \leq \max\{ d_p(x, y), d_p(y, z) \}, making the rational numbers \mathbb{Q} a metric space with a non-Archimedean topology where "closeness" emphasizes congruence modulo high powers of p. The p-adic numbers \mathbb{Q}_p form the completion of \mathbb{Q} with respect to this metric, obtained by adjoining limits of all Cauchy sequences in \mathbb{Q} under d_p; the rationals are dense in \mathbb{Q}_p, ensuring that \mathbb{Q} embeds naturally as a subfield. Elements of \mathbb{Q}_p can be represented via p-adic expansions, analogous to decimal expansions but in base p and extending infinitely to the left. Any x \in \mathbb{Q}_p admits a unique of the form x = \sum_{n = k}^\infty d_n p^n, where k \in [\mathbb{Z}](/page/Z) is sufficiently negative, each d_n is an with $0 \leq d_n < p, and the series converges in the p-adic metric. This representation is written as \dots d_2 d_1 d_0 . d_{-1} d_{-2} \dots_p, with the "decimal" point separating non-negative and negative powers of p. The field \mathbb{Q}_p is locally compact and complete under the p-adic metric, differing fundamentally from the real numbers [\mathbb{R}](/page/R) in its non-Archimedean nature, where there are infinitesimally small nonzero elements relative to the valuation. It plays a crucial role in , for instance, in solving systems of congruences via , which lifts solutions modulo p to solutions in \mathbb{Z}_p, the ring of p-adic integers.

Formal Constructions

As Equivalence Classes of Pairs

One formal way to construct the rational numbers \mathbb{Q} is as the set of equivalence classes of ordered pairs of integers, where each pair (p, q) consists of an integer p \in \mathbb{Z} (the numerator) and a nonzero integer q \in \mathbb{Z} \setminus \{0\} (the denominator). The equivalence relation \sim on \mathbb{Z} \times (\mathbb{Z} \setminus \{0\}) is defined by (p, q) \sim (r, s) if and only if p s = q r. This relation is reflexive, symmetric, and transitive, partitioning the set of pairs into equivalence classes, each denoted [(p, q)], which represent the rational numbers. Addition and multiplication on these equivalence classes are defined componentwise to mimic fractional arithmetic while remaining within integers. Specifically, for equivalence classes [(p, q)] and [(r, s)], [(p, q)] + [(r, s)] = [(p s + q r, q s)], [(p, q)] \cdot [(r, s)] = [(p r, q s)]. These operations are associative, commutative, and distributive, with additive identity [(0, 1)] and multiplicative identity [(1, 1)]. Additive inverses exist as [- (p, q)] = [(-p, q)], and nonzero elements have multiplicative inverses [(q, p)] provided p \neq 0. To ensure the operations are well-defined on equivalence classes rather than specific representatives, one verifies that if (p, q) \sim (p', q') and (r, s) \sim (r', s'), then the sums and products of the pairs are equivalent: for , (p s + q r, q s) \sim (p' s' + q' r', q' s'), and similarly for . This construction yields a structure on \mathbb{Q}, satisfying all field axioms, including the existence of inverses for nonzero elements. This pair-based approach avoids direct use of fractions, operating solely with integers and their products, yet establishes a natural isomorphism to the intuitive notion of rationals via the mapping [(p, q)] \mapsto \frac{p}{q}, where equivalent pairs yield the same fraction. The integers \mathbb{Z} embed into \mathbb{Q} via the injective homomorphism n \mapsto [(n, 1)], preserving addition and multiplication.

Quotient Field of Integers

In , the rational numbers \mathbb{Q} arise as the quotient field of the \mathbb{Z}, which is an . More abstractly, \mathbb{Q} is obtained by localizing \mathbb{Z} at the multiplicative set S = \mathbb{Z} \setminus \{0\}, yielding the ring of fractions S^{-1}\mathbb{Z}. Since \mathbb{Z} has no zero divisors, every element of S is invertible in this localization, making S^{-1}\mathbb{Z} a ; concretely, its elements can be represented as equivalence classes of pairs (a, b) with a \in \mathbb{Z}, b \in S, under the (a, b) \sim (c, d) if there exists s \in S such that s(ad - bc) = 0, though the pair representation aligns with the explicit construction in equivalence classes of pairs. A key feature of this construction is its universal property: for any injective ring homomorphism \phi: \mathbb{Z} \to F into a field F, there exists a unique ring homomorphism \overline{\phi}: \mathbb{Q} \to F such that \overline{\phi} \circ \iota = \phi, where \iota: \mathbb{Z} \hookrightarrow \mathbb{Q} is the natural embedding sending n \mapsto (n, 1). To see this, define \overline{\phi}((a, b)) = \phi(a) \phi(b)^{-1}; since \phi(b) \neq 0 for b \in S (as F has no zero divisors and \phi is nonzero on S), this is well-defined and preserves addition and multiplication because the operations in \mathbb{Q} are induced componentwise from \mathbb{Z}. Uniqueness follows from the fact that every element of \mathbb{Q} is generated by elements of \mathbb{Z} under inversion. This implies that \mathbb{Q} is the smallest containing \mathbb{Z} (up to the \iota), in the sense that any containing an isomorphic copy of \mathbb{Z} must contain a subfield isomorphic to \mathbb{Q}. Moreover, any two such fields of \mathbb{Z} are isomorphic via a unique compatible with the embeddings from \mathbb{Z}. One advantage of viewing \mathbb{Q} through this lens is its generalization: for any R, the localization at R \setminus \{0\} yields the field of fractions of R, providing a uniform way to adjoin inverses for nonzero elements while preserving the ring structure.

History

Ancient Origins

The earliest known uses of rational numbers appear in ancient around 2000 BCE, where Babylonian mathematicians employed a (base-60) system to represent fractions, particularly in astronomical calculations. This system allowed for precise divisions of circles into 360 degrees and time into hours, minutes, and seconds, with tablets from sites like Senkerah containing tables of squares and cubes up to 59 and 32, respectively, to facilitate computations involving rational proportions. These fractions were essential for predicting celestial events, demonstrating an advanced practical grasp of rational quantities without symbolic algebraic notation. In , rational numbers were expressed as sums of distinct unit fractions (fractions with numerator 1), a method evident in the Rhind Papyrus, dated to approximately 1650 BCE and copied by the scribe . The papyrus includes a table decomposing fractions of the form 2/n (for odd n from 5 to 101) into such sums, for example, 2/5 = 1/3 + 1/15, applied to problems in , area measurement, and . This approach reflected a preference for unit fractions in practical , avoiding more general forms. Babylonian scribes also solved linear equations involving rational coefficients without formal algebraic notation, relying instead on reciprocal tables and step-by-step procedures during the Old Babylonian period (c. 2000–1900 BCE). For instance, they addressed problems like finding x such that (2/3) × (2/3) × x + 100 = x, yielding x = 180 through tabular lookups and proportional reasoning. In ancient Greece, around the 5th century BCE, the Pythagorean school initially viewed all quantities as rational ratios of integers, but the discovery of irrational numbers—such as √2, proven via contradiction using the on an isosceles —challenged this belief and highlighted the distinction between rationals and irrationals. By c. 300 BCE, formalized in his Elements, Book V, defining a as the between two magnitudes of the same kind and proportion as equality of such ratios, applicable to both commensurable (rational) and incommensurable cases, providing a rigorous framework for geometric and arithmetic manipulations akin to rational numbers. In , during the 5th century , incorporated fractions into trigonometric computations in his Āryabhaṭīya (c. 499 ), using them to construct sine tables via the and approximate π as 62,832/20,000 (≈3.1416), advancing and geometric projections.

Development in Modern Mathematics

In the , the notation for rational numbers advanced significantly with the introduction of fractions by in his 1585 pamphlet La Thiende. Stevin proposed representing fractions using powers of ten, allowing for a positional system that extended beyond integers to include terminating and repeating , thereby simplifying arithmetic operations on rationals and facilitating their use in engineering and commerce. The formalization of rational numbers within progressed in the through Richard Dedekind's axiomatization of in 1871. In his supplements to Dirichlet's Vorlesungen über Zahlentheorie, Dedekind defined a as a with unity where every nonzero element has a , positioning the rationals \mathbb{Q} as the prime of characteristic zero and the foundational example of an . This abstraction enabled the study of extensions and structures beyond the rationals, influencing . Set-theoretic constructions of the rationals emerged in the late , building on Giuseppe Peano's 1889 axioms for the natural numbers, which provided a rigorous foundation for from which integers and then rationals could be derived as equivalence classes of pairs. Concurrently, proved in 1874 that the rationals are countable, demonstrating a bijection between \mathbb{Q} and the natural numbers via a enumeration of positive fractions in lowest terms, highlighting their "small" compared to the reals. In the 20th century, rational numbers played a central role in , where \mathbb{Q} serves as the prime model of the of ordered fields, embedding elementarily into every countable model due to its countable dense linear without endpoints. This structure exemplifies a complete, decidable with after naming constants. Additionally, Hilbert's 1900 problems encompassed foundational questions in logic, including aspects of decidability for arithmetic systems; the of the rationals as a dense linear without endpoints was affirmatively resolved as decidable, with algorithms existing via its axiomatization and .

References

  1. [1]
    1.1 What Are Numbers? The Rational Numbers - MIT Mathematics
    Ratios of integers are called rational numbers, and you get one for any pairs of integers, so long as the second integer, called the denominator, is not zero.
  2. [2]
    Real Numbers:Rational - Department of Mathematics at UTSA
    Oct 21, 2021 · In mathematics, a rational number is a number that can be expressed as the quotient or fraction {\displaystyle {\frac {p}{q}}} of two integers.
  3. [3]
    Rational Number -- from Wolfram MathWorld
    A rational number is a number that can be expressed as a fraction p/q, where p and q are integers and q!=0.
  4. [4]
    [PDF] Math 8: Rational Numbers
    We can now define a rational number to be any equivalence class of this relation. ... (In other words, a non-zero rational number can be written in lowest terms.).
  5. [5]
    Real Number -- from Wolfram MathWorld
    Real numbers are the field of all rational and irrational numbers, denoted as R, and also called the continuum, denoted as c.
  6. [6]
    Irrational Number -- from Wolfram MathWorld
    An irrational number is a number that cannot be expressed as a fraction p/q for any integers p and q. Irrational numbers have decimal expansions that ...
  7. [7]
    Egyptian Fractions - Mathematics - Williams College
    Feb 12, 2016 · Ancient Egyptians demanded that every fraction have 1 in the numerator. They wanted to write any rational between 0 and 1 as a sum of such “unit” fractions.
  8. [8]
    Rational - Etymology, Origin & Meaning
    Originating in late 14th-century Latin rationalis from ratio, late 14c. "pertaining to reason," late 15c. "endowed with reason," late meaning "reasonable."
  9. [9]
    rational - Wiktionary, the free dictionary
    Etymology 1​​ From Old French rationel, rational, from Latin rationalis (“of or belonging to reason, rational, reasonable; having a ratio”), from ratio (“reason; ...English · Etymology 1 · Etymology 2 · French<|separator|>
  10. [10]
    Ratio, Rational and Irrational ... History and Etymology of Math Terms
    Oct 21, 2024 · By the Latin reri it had taken on the ideas of "reason", from which comes rational, and ratio for a comparison of two magnitudes. Rate is a ...
  11. [11]
    Euclid - Biography - MacTutor - University of St Andrews
    Euclid was a Greek mathematician best known for his treatise on geometry: The Elements. This influenced the development of Western mathematics for more than ...
  12. [12]
    Fraction - Etymology, Origin & Meaning
    Late 14c. origin from Anglo-French and Latin 'fractionem' meaning "a breaking"; denotes breaking, dividing, or fragmenting, evolving from Latin 'frangere' ...
  13. [13]
    Numerator - Etymology, Origin & Meaning
    Originating in the 1540s from Late Latin numerator, meaning "counter," numerator means the number above the line in a fraction or one who numbers.
  14. [14]
    Denominator - Etymology, Origin & Meaning
    Denominator, from Medieval Latin denominare meaning "to name," originates from PIE root *no-men- "name"; it denotes the term in a fraction indicating the ...
  15. [15]
    Origin and use of the adjective "improper" in mathematics
    Apr 9, 2019 · "Improper" fraction was used in English in 1542 by Robert Recorde in The ground of artes, teachyng the worke and practise of arithmetike.
  16. [16]
    History of Fractions - Basic Mathematics
    The modern fractional notation likely has its roots in the Hindu mathematical tradition. Around A.D. 630, the Indian mathematician Brahmagupta introduced a ...
  17. [17]
    [PDF] chapter 6: the rational numbers q - CSUSM
    In Q we can be picky and insist that the denominator be positive: Theorem 20. If r ∈ Q then there are integers a, b such that r = a b and such that b is ...
  18. [18]
    [PDF] Arithmetic Review - City Tech
    Both forms are correct, but in subsequent courses you will find that improper fractions are preferred to mixed numbers. 17. Write as a mixed number. 5. Divide ...
  19. [19]
    Decimal Expansion -- from Wolfram MathWorld
    A decimal expansion is a base-10 representation of a number, using digits multiplied by powers of 10, decreasing from left to right.
  20. [20]
    Repeating Decimal -- from Wolfram MathWorld
    A repeating decimal, also called a recurring decimal, is a number whose decimal representation eventually becomes periodic.Missing: sources | Show results with:sources
  21. [21]
    [PDF] 1.3 The Real Numbers.
    Proposition 1.3.5. Every repeating decimal is the decimal expansion of some rational number. Proof: Start with a repeating decimal r = q.d1d2.
  22. [22]
  23. [23]
    [PDF] The Winning EQUATION - CSUN
    Also recall that (cd)n = an bn from Module 1. Therefore, 10n = (2 • 5)n = 2n • 5n. This means that the only prime factors of 10n are 2 and 5.
  24. [24]
    Decimal Period -- from Wolfram MathWorld
    The decimal period of a repeating decimal is the number of digits that repeat. For example, 1/3=0.3^_ has decimal period one, 1/11=0.09^_ has decimal period ...
  25. [25]
    [PDF] Section 15. Decimals
    Mar 17, 2022 · We can similarly take any terminating decimal representation and convert it into an infinite repeating decimal representation. For example, we ...
  26. [26]
    [PDF] Contents 2 Modular Arithmetic in Z - Evan Dummit
    begins repeating immediately after the decimal point, and the length of the period is the order of 10 modulo q. In particular, the length of the period divides ...
  27. [27]
    Continued Fraction -- from Wolfram MathWorld
    ### Summary of Continued Fraction Representation for Rational Numbers
  28. [28]
    [PDF] Continued Fractions and the Euclidean Algorithm
    The process of finding the continued fraction expansion of a rational number is essentially identical to the process of applying the Euclidean algorithm to the ...
  29. [29]
    [PDF] Continued Fractions - DSpace@MIT
    Apr 2, 2003 · Summary: Any simple continued fraction represents a rational number, and any rational number can be expressed as. Fraction in exactly two ...
  30. [30]
    [PDF] Math 180B - Notes
    The digits of the continued fraction are precisely the sequence of quotients1 from the Euclidean Algo- rithm. among all fractions with denomina- tor at most 13 ...
  31. [31]
    [PDF] Contents 6 Rational Approximation and Diophantine Equations
    • Proposition (Rational Approximation and Continued Fractions): Suppose α is any irrational real number and p/q is any rational number. Then the following ...
  32. [32]
    Rational Numbers - Duke Physics
    Dividing one out produces a finite number of non-repeating digits, followed by a finite sequence of digits that repeats cyclically forever, for example $1/6 ...
  33. [33]
    Egyptian Fractions - Mathematicians of the African Diaspora
    THEOREM. Every rational number is an egyptian number. The modern proof of the Theorem was discovered in 1880, but European's have known how to compute Egyptian ...
  34. [34]
    Algorithms for Egyptian Fractions - UC Irvine
    We will represent Egyptian fractions as lists of unit fractions. The original rational number represented by such a list can be recovered by Plus@@%. Throughout ...
  35. [35]
    Stern-Brocot Tree - Interactive Mathematics Miscellany and Puzzles
    Continuing this way we get an infinite tree known as the Stern-Brocot tree because it was discovered independently by the German mathematician Moriz Stern (1858) ...
  36. [36]
    Recounting the Rationals - Penn Math - University of Pennsylvania
    No information is available for this page. · Learn whyMissing: original paper
  37. [37]
    [PDF] Construction of Rational Maps on the Projective Line with Given ...
    May 11, 2016 · Definition 2.1. The projective line, denoted by P. 1. , is the rational numbers Q adjoined with the point at infinity c. P1. = Q U 1cl. 2.1.
  38. [38]
    [PDF] Equivalent Fractions When do two fractions represent the same ...
    It's called "cross multiplication." Two fractions a b and c d are equal if and only if ad = bc. For example,. 3. 6. = 4. 8 because 3 • 8 ...
  39. [39]
    [PDF] Fractions, Decimals, and Rational Numbers - UC Berkeley math
    Those fractions whose denominators are all positive powers of 10, e.g.,. 1489 ... negative × negative = positive is that the distributive law holds for rational.
  40. [40]
    [PDF] The Type of Rational Numbers (in OCaml and Mathematics) (10/1)
    The “fraction” 1/2 is a canonical form of a rational number because it can't be further reduced. To put a/b in canonical form, we want to remove common factors, ...Missing: irreducible | Show results with:irreducible
  41. [41]
    [PDF] Real Analysis, Fall 2017–Spring 2018
    Jun 2, 2022 · 6 Definition. An ordered field is a field F equipped with a total order, so a set with a relation ≤ and two operations + and · satisfying axioms ...
  42. [42]
    [PDF] An Approach to Fractions Via Measurement - UCSD Math
    Feb 28, 2008 · If we divide each unit into two equal parts, and use these as new measuring units, the same segment is congruent to 2n new units, so its ...
  43. [43]
    [PDF] 5_3_4 M1 Notes Fall 2010
    We say the set of. Rational Numbers is dense. Density means that for any 2 rational numbers there can always be found a rational number between them. In other ...<|control11|><|separator|>
  44. [44]
    [PDF] Honors Thesis A Bound on the Maximum Coefficient of Gaussian ...
    ... mediant of a/b and c/d to be (a + c)/(b + d). Then the mediant is always between the two fractions. In other words, if a/b ≤ c/d, a b. ≤ a + c b + d. ≤ c d.<|control11|><|separator|>
  45. [45]
    [PDF] The Archimedean Property - Penn Math
    Sep 3, 2014 · An ordered field F has the Archimedean Property if, given any positive x and y in F, there is an integer n > 0 so that nx > y.Missing: rational density
  46. [46]
    [PDF] Math 327: Real numbers and limits
    Oct 25, 2010 · Theorem 2.14. The field Q of rational numbers does not have the least upper bound property. Proof. To prove this, we need to use the following ...Missing: lack | Show results with:lack
  47. [47]
    [PDF] Adding and Subtracting Rational Numbers
    To add or subtract rational numbers in fraction form, you must have common denominators. 4. To add or subtract rational numbers in decimal form, you must ...
  48. [48]
    The Rational Numbers
    Additive inverse: Every rational number $x \in \rationals$ has an additive inverse $y \in \rationals$ such that $x + y = 0$. The additive inverse is generally ...
  49. [49]
  50. [50]
    [PDF] Fractions and Rational Expressions
    When adding or subtracting two fractions, we must make sure that the denominator of the fractions match. If the denominators are different, we have to find ...
  51. [51]
    [PDF] 1.2 The Integers and Rational Numbers
    The associative and commutative laws of addition can now be proved for this new definition of addition by the same proof-by-induction strategy we used in §1.1 ( ...
  52. [52]
    Rational Number Library - Brown Computer Science
    If an unlimited precision integer type is available, rational numbers based on it will never overflow and will provide exact calculations in all circumstances.Missing: computation | Show results with:computation
  53. [53]
    [PDF] The Rational Numbers
    According to the definition, a rational number is an equivalence class containing certain pairs of integers. What do some of the equivalence classes look ...
  54. [54]
    [PDF] Multiplying And Dividing Rational Expressions Worksheet 8
    Then, simplify the resulting expression by factoring and canceling common factors. ... easier and reduces the chance of errors. Working with smaller ...
  55. [55]
    Exponents - Algebra Trig Review
    ... pnpm=pn+mpnpm=pn−m=1pm−n(pn)m=pnmp0=1, provided p≠0(pq)n=pnqn(pq)n=pnqn ...<|control11|><|separator|>
  56. [56]
    Basic rules for exponentiation - Math Insight
    If a is any rational number, then it can be written as a=m/n. We can define taking a number to the ath power as taking that number to the mth power and the nth ...
  57. [57]
    [PDF] notes on finite fields
    If no such n ∈ Z>0 exists, then we say K has characteristic 0. Example 3.11. The rational numbers Q has characteristic 0, but the field Fp has characteristic p.
  58. [58]
    The Rational Numbers - Math Foundations
    We are all familiar with the rational numbers: a rational number is any number that can be written as pq, where p and q are integers, and q≠0. A more formal ...
  59. [59]
    Numbers - Math 1010 on-line
    We say that the set of rational numbers is closed under addition, subtraction, multiplication, and division.
  60. [60]
    [PDF] Lecture 1. - University of Southern California
    ... rational numbers, i.e.,. I ⊂ Q. The rational numbers are closed under the operations of addition, subtraction, multiplication and division (except by zero).<|separator|>
  61. [61]
    [PDF] 1 Basic facts - Berkeley Math Circle
    A rational number is an algebraic number; a rational number is an algebraic ... The set of algebraic numbers is closed under addition, subtraction, multiplication ...
  62. [62]
    [PDF] Natural Numbers, Integers, and Rational Numbers (Following ...
    Remark 5.4 We may embed the integers into the rationals in a way similar to the embedding of the naturals into the integers. The function f : Z → Q, i.e. f ...
  63. [63]
    [PDF] Transcendental Numbers
    Dec 3, 2022 · The rational numbers (Q) are incomplete in two different ways. Firstly, Q is not algebraically closed because there exist polynomials with ...<|control11|><|separator|>
  64. [64]
    [PDF] MATH 361: NUMBER THEORY — NINTH LECTURE 1. Algebraic ...
    Every rational number r is algebraic because it satisfies the polynomial x − r, but not every algebraic number is rational, as shown by the examples given just.
  65. [65]
    Georg Cantor (1845 - 1918) - Biography - MacTutor
    ... Cantor's 1872 paper which Cantor had sent him. In 1873 Cantor proved the rational numbers countable, i.e. they may be placed in one-one correspondence with ...
  66. [66]
    4.7 Cardinality and Countability
    We say a set A is countably infinite if N≈A, that is, A has the same cardinality as the natural numbers. We say A is countable if it is finite or countably ...
  67. [67]
    Recounting the Rationals - jstor
    It follows that a list of all positive rational numbers, each appearing once and only once, can be made by writing down 1/1, then the fractions on the level ...
  68. [68]
    [PDF] Chapter 2: Numbers - UC Davis Math
    In this chapter, we describe the properties of the basic number systems. We briefly discuss the integers and rational numbers, and then consider the real num-.Missing: mathematics | Show results with:mathematics
  69. [69]
    [PDF] Cauchy's Construction of R - UCSD Math
    The real numbers will be constructed as equivalence classes of Cauchy sequences. Let CQ denote the set of all Cauchy sequences of rational numbers. We must ...
  70. [70]
    Dedekind's Contributions to the Foundations of Mathematics
    Apr 22, 2008 · As noted, Dedekind starts with the system of rational numbers; he uses a set-theoretic procedure to construct, in a central step, the new system ...
  71. [71]
    [PDF] Density of the Rationals - UC Davis Math
    The density of rationals means there is a rational number strictly between any two distinct real numbers, however close together they may be.
  72. [72]
    Where do the "best" rational approximations come from?
    The "best" rational approximations, as well as most of the theory of rational approximation, arise from continued fraction expansions.
  73. [73]
    Real numbers 2 - MacTutor History of Mathematics
    This was totally new since all other methods built the real numbers from the known rational numbers. Hilbert's numbers were unconnected with any known system.
  74. [74]
    [PDF] Topology of the Real Numbers - UC Davis Math
    This chapter defines topological properties of real numbers, including open sets, which are defined as sets where every point has a neighborhood within the set.
  75. [75]
    [PDF] MATH 411 HOMEWORK 2 SOLUTIONS 2.13.3. Let X be a set, and ...
    (a) Show that the collection B = {(a, b) | a, b ∈ Q, a<b} is a basis for the standard topology on R. The key fact is that for any real numbers a, b, we may find ...
  76. [76]
    [PDF] Tutorial Sheet 6, Topology 2011
    1. Prove that Q, with the subspace topology inherited from R, is totally disconnected, but not discrete. Solution: It is not discrete because {p/q} is not open ...Missing: incompleteness | Show results with:incompleteness
  77. [77]
    [PDF] countable metric spaces without isolated points - Abhijit Dasgupta
    Any countable metrizable space without isolated points is homeomorphic to Q, the rationals with the order topology (same as Q as a subspace of R with usual ...Missing: locally | Show results with:locally
  78. [78]
    [PDF] A first course on 𝑝-adic numbers - UCSD Math
    For a prime 𝑝, we define the 𝑝-adic numbers, denoted 𝐐𝑝, to be the completion of (𝐐,|∙ |𝑝). We also define the 𝑝-adic integers, denoted 𝐙𝑝, to be the valuation ...<|control11|><|separator|>
  79. [79]
    p-adic Number -- from Wolfram MathWorld
    A p-adic number extends rationals, relating congruences modulo powers of a prime p to proximity in a p-adic metric. It's the completion of rationals with ...
  80. [80]
    [PDF] Constructing the Rationals
    a) We define the difference B C œ B Р CСЮ. So subtraction is defined in terms of addition (adding the additive inverse). b) If. , we define the. B Б ! C ƒ ...
  81. [81]
    [PDF] NOTES ON REAL NUMBERS In these notes we will construct the set ...
    The set of rational numbers, denoted by Q, is defined to be Z × Z∗/ ∼. An equivalence class [(a, b)] represents the rational number a b. , and again ...
  82. [82]
    [PDF] Introduction to Commutative Algebra
    vii) Let A be an integral domain with just one non-zero prime ideal p, and let K be the field of fractions of A. Let B (A/p) x K. Define 4: A B by p(x) = (x ...
  83. [83]
    [PDF] 6. Fields I
    That is, up to unique isomorphism, there is only one field of fractions of an integral domain. Proof: First prove that any field map f : Q −→ Q such that ...
  84. [84]
    [PDF] Analysis 1
    Rational numbers. The rational numbers Q are the field of fractions of Z, that is, the smallest field containing Z such that for each n ∈ Z \ {0} the ...
  85. [85]
    Babylonian mathematics - MacTutor - University of St Andrews
    Writing developed and counting was based on a sexagesimal system, that is to say base 60. Around 2300 BC the Akkadians invaded the area and for some time the ...Missing: astronomy | Show results with:astronomy
  86. [86]
    Egyptian Fraction -- from Wolfram MathWorld
    The famous Rhind papyrus, dated to around 1650 BC contains a table of representations of 2/n as Egyptian fractions for odd n between 5 and 101. The reason ...Missing: BCE | Show results with:BCE
  87. [87]
    How a Secret Society Discovered Irrational Numbers
    Jun 13, 2024 · The famed proof of irrational numbers presented by Hippasus—or another Pythagorean—is most easily illustrated with an isosceles right triangle: ...
  88. [88]
    Real numbers 1 - MacTutor History of Mathematics
    In Book VII Euclid studies numbers. He makes a series of definitions. First he defines a unit, then a number is defined as being composed of a multitude of ...
  89. [89]
    Aryabhata | Achievements, Biography, & Facts - Britannica
    Aryabhata (born 476, possibly Ashmaka or Kusumapura, India) was an astronomer and the earliest Indian mathematician whose work and history are available to ...
  90. [90]
    Simon Stevin - Biography - MacTutor - University of St Andrews
    In 1585 he published La Thiende Ⓣ. (The tenth). , a twenty-nine page booklet in which he presented an elementary and thorough account of decimal fractions.