Fact-checked by Grok 2 weeks ago

Construction of the real numbers

The construction of the real numbers is a foundational process in mathematics that rigorously defines the set \mathbb{R} as the completion of the rational numbers \mathbb{Q}, ensuring it forms a complete ordered field where every nonempty subset bounded above has a least upper bound. This construction addresses the incompleteness of \mathbb{Q}, which lacks certain limits like \sqrt{2}, by extending it to include all limits of Cauchy sequences or partitions known as Dedekind cuts. Developed independently in the late 19th century, primarily by Richard Dedekind and Georg Cantor in 1872, it provides the arithmetic, order, and completeness properties essential for calculus and real analysis. Historically, the need for a precise construction arose during the 19th century as mathematicians like Augustin-Louis Cauchy and Karl Weierstrass sought to rigorize analysis, revealing gaps in \mathbb{Q} through concepts like irrational numbers and continuity. Dedekind, working in the 1850s but publishing in 1872, introduced Dedekind cuts as partitions of \mathbb{Q} into two nonempty subsets A and B such that all elements of A are less than those in B, A has no greatest element, and every rational is in exactly one set; each cut corresponds to a real number, with rational reals represented by cuts where the upper set has a least element. For instance, the cut for \sqrt{2} places in A all rationals q with q < 0 or q^2 < 2, and in B the rest. Cantor, also in 1872, used Cauchy sequences—sequences of rationals where terms get arbitrarily close—to define reals as equivalence classes under the relation where two sequences differ by a null sequence (one converging to 0), yielding the same structure up to isomorphism. These constructions establish \mathbb{R} as an Archimedean ordered field, meaning for any positive reals x and y, there exists a natural number n such that nx > y, and \mathbb{Q} is dense in \mathbb{R} (between any two reals lies a rational). The completeness axiom—via the least upper bound property—distinguishes \mathbb{R} from \mathbb{Q}, enabling theorems like the intermediate value theorem in calculus. Both methods build hierarchically from natural numbers (via Peano axioms) to integers and rationals using equivalence classes, ensuring arithmetic operations and order are well-defined and compatible. Up to isomorphism, \mathbb{R} is the unique complete ordered field, a result proven using the constructions themselves.

Motivations for Constructing the Reals

Gaps in the Rational Numbers

The rational numbers \mathbb{Q}, consisting of all fractions a/b where a, b \in \mathbb{Z} and b \neq 0, form a field under the standard operations of addition and multiplication. This structure satisfies the field axioms, including closure under addition and multiplication, associativity, commutativity, the existence of additive and multiplicative identities (0 and 1, respectively), additive inverses, and multiplicative inverses for nonzero elements, with distributivity holding throughout. However, despite these algebraic properties, \mathbb{Q} exhibits significant deficiencies that prevent it from serving as a foundation for analysis. One key algebraic gap is that \mathbb{Q} is not algebraically closed, meaning not every non-constant polynomial with coefficients in \mathbb{Q} has a root in \mathbb{Q}. A classic example is the equation x^2 = 2, which has no rational solution. To see this, suppose \sqrt{2} = p/q where p, q are positive integers with no common factors and q \neq 0. Then p^2 = 2q^2, implying p^2 is even, so p must be even (as the square of an odd integer is odd). Let p = 2k; substituting gives $4k^2 = 2q^2, or q^2 = 2k^2, so q^2 is even and q is even. This contradicts the assumption that p/q is in lowest terms. Thus, \sqrt{2} is irrational, highlighting how simple quadratic equations escape the rationals. Analytically, the rationals suffer from incompleteness, lacking a least upper bound property for all bounded subsets. For instance, the decimal expansion of \pi \approx 3.14159\ldots is infinite and non-repeating, a direct consequence of \pi's irrationality. The Leibniz series \pi/4 = 1 - 1/3 + 1/5 - 1/7 + \cdots produces rational partial sums that approximate \pi but never equal it, as convergence in \mathbb{Q} would imply \pi \in \mathbb{Q}, which it is not. Johann Lambert proved \pi irrational in 1761 using continued fractions, showing that assuming \pi = a/b leads to a contradiction via infinite descent in the denominator. This gap means that sequences of rationals, such as those from decimal approximations or series, may "converge" to points outside \mathbb{Q}, underscoring the need for an extension to handle limits properly. The intermediate value theorem also fails over the rationals. Consider the continuous function f(x) = x^2 on the rational interval \mathbb{Q} \cap [2, 3]. Here, f(2) = 4 and f(3) = 9, so the rational value 5 lies between the endpoint images. Yet no x \in \mathbb{Q} satisfies x^2 = 5, since \sqrt{5} is irrational (provable similarly to \sqrt{2}). This example demonstrates how continuous rational functions on rational domains can skip rational intermediate values due to the "holes" in \mathbb{Q}. Historical examples like Zeno's paradoxes further illustrate these suprema issues in the rationals. In the dichotomy paradox, to traverse a distance of 1, one must first cover 1/2, then 1/4, and so on infinitely; the partial sums of this geometric series (all rational) approach 1, requiring the concept of a limit to resolve the infinite process. Modern resolutions via real numbers provide the completeness needed for such limits in general.

Historical Development

The concept of real numbers emerged gradually through efforts to address limitations in rational numbers, beginning with ancient approximations of irrational quantities. Around 400 BCE, Eudoxus of Cnidus developed the method of exhaustion, a technique for rigorously calculating areas and volumes by approximating curved figures with inscribed polygons, effectively handling irrational magnitudes without explicitly naming them as such. This approach, later formalized in Euclid's Elements (Books V and XII), allowed comparisons of incommensurable lengths, such as the ratio of 1 to √2, laying early groundwork for a continuous number system. During the medieval and Renaissance periods, Indian and Arab mathematicians advanced numerical systems that facilitated work with decimals and irrationals. Indian scholars in the 4th to 6th centuries developed the decimal place-value system, which Arabs adapted by the 9th century for systematic arithmetic operations. Al-Khwarizmi, around 825 CE, provided the first comprehensive treatment of these operations, including addition, subtraction, multiplication, and division using Hindu-Arabic numerals, though his focus remained on rationals. Successors like Al-Karaji (c. 953–1029) and Al-Samaw'al (c. 1130–1180) extended this to irrationals by developing approximation rules for square and cubic roots via decimal fractions, treating them as limits of rational sequences. In the 17th century, the study of infinite series highlighted the need for numbers beyond algebraic irrationals, pointing toward transcendentals. John Wallis, in 1655, derived an infinite product formula for π/2 as ∏{n=1}^∞ [(2n/(2n-1)) · ((2n)/(2n+1))], approximating the transcendental constant through rational terms and revealing the inadequacy of finite rationals for such values. Isaac Newton, building on this in the 1660s, employed binomial expansions and infinite series to compute π to 15 decimal places and to represent e as ∑{n=0}^∞ 1/n!, demonstrating how series could model transcendental phenomena but underscoring gaps in the rational framework. The 19th century brought a crisis in analysis, driven by physical applications requiring robust convergence properties. Joseph Fourier's work on the heat equation in the 1820s, published in Théorie Analytique de la Chaleur (1822), modeled temperature diffusion via infinite trigonometric series, demanding convergence beyond rational approximations to ensure realistic continuous solutions. This motivated formal constructions of the reals. In 1817, Bernard Bolzano proved the intermediate value theorem for continuous functions, implicitly assuming a complete real line where bounded sequences have limits, though without explicit construction. By the 1850s–1870s, Karl Weierstrass emphasized Cauchy sequences in his lectures to rigorize limits and continuity, influencing the arithmetization of analysis. Charles Méray published a construction in 1869, and Georg Cantor formalized this in 1872 by defining reals as equivalence classes of Cauchy sequences of rationals, providing a sequential construction. That same year, Richard Dedekind introduced cuts in the rationals to define reals, ensuring completeness without geometry.

Axiomatic Characterizations

Ordered Field Axioms

The ordered field axioms establish the foundational algebraic structure and compatible ordering for the real numbers, ensuring they form a system suitable for analysis while distinguishing them from the rationals through additional properties like completeness. These axioms consist of the field axioms, which define the arithmetic operations, and the order axioms, which introduce a total order preserved by those operations. Together, they characterize any ordered field, including both the rationals and the reals, but the reals require further axioms for completeness. The field axioms require that the set ℝ, equipped with binary operations addition (+) and multiplication (·), forms a field. Specifically:
  • Closure under addition: For all x, y \in \mathbb{R}, x + y \in \mathbb{R}.
  • Associativity of addition: For all x, y, z \in \mathbb{R}, x + (y + z) = (x + y) + z.
  • Commutativity of addition: For all x, y \in \mathbb{R}, x + y = y + x.
  • Additive identity: There exists $0 \in \mathbb{R}such that for allx \in \mathbb{R}, x + 0 = x$.
  • Additive inverses: For every x \in \mathbb{R}, there exists -x \in \mathbb{R} such that x + (-x) = 0.
  • Closure under multiplication: For all x, y \in \mathbb{R}, x \cdot y \in \mathbb{R}.
  • Associativity of multiplication: For all x, y, z \in \mathbb{R}, x \cdot (y \cdot z) = (x \cdot y) \cdot z.
  • Commutativity of multiplication: For all x, y \in \mathbb{R}, x \cdot y = y \cdot x.
  • Multiplicative identity: There exists $1 \in \mathbb{R}such that1 \neq 0and for allx \in \mathbb{R}, x \cdot 1 = x$.
  • Multiplicative inverses: For every x \in \mathbb{R} with x \neq 0, there exists x^{-1} \in \mathbb{R} such that x \cdot x^{-1} = 1.
  • Distributivity: For all x, y, z \in \mathbb{R}, x \cdot (y + z) = x \cdot y + x \cdot z.
These ten field axioms ensure that ℝ behaves as an algebraic field under the operations of addition and multiplication. The order axioms introduce a total order < on ℝ that is compatible with the field operations, defining the positive elements as the set P = \{ x \in \mathbb{R} \mid x > 0 \}. Specifically:
  • Trichotomy: For every x \in \mathbb{R}, exactly one of the following holds: x > 0, x = 0, or x < 0.
  • Transitivity: For all x, y, z \in \mathbb{R}, if x < y and y < z, then x < z.
  • Addition preservation: For all x, y, z \in \mathbb{R}, if x < y, then x + z < y + z.
  • Multiplication preservation: For all x, y, z \in \mathbb{R}, if x < y and $0 < z, then x \cdot z < y \cdot z.
From these, the positive elements P are closed under addition and multiplication: if x, y \in P, then x + y \in P and x \cdot y \in P. Additionally, $1 \in P$, ensuring the order is non-trivial. An important property for ordered fields intended to model the reals is the Archimedean property, which states that for all a, b \in \mathbb{R} with a > 0 and b > 0, there exists a natural number n \in \mathbb{N} such that n a > b. This prevents "infinitesimals" or unbounded gaps in the positives and holds in any ordered field embeddable into the reals. The rational numbers \mathbb{Q}, with their standard addition, multiplication, and order, satisfy all the field axioms, order axioms, closure of positives, and the Archimedean property, forming an ordered field. However, \mathbb{Q} is incomplete: consider the nonempty bounded-above subset S = \{ q \in \mathbb{Q} \mid q^2 < 2 \}; it has no least upper bound in \mathbb{Q}, as any candidate rational upper bound can be exceeded by another element of S while remaining rational. This gap demonstrates that the ordered field axioms alone do not suffice for the reals, necessitating a completeness axiom.

Completeness via Least Upper Bound

The least upper bound property, also known as the supremum axiom, serves as a fundamental completeness axiom for the real numbers \mathbb{R}, stating that every non-empty subset S \subseteq \mathbb{R} that is bounded above has a least upper bound \sup S \in \mathbb{R}./01:_Tools_for_Analysis/1.05:_The_Completeness_Axiom_for_the_Real_Numbers) This property ensures that \mathbb{R} is complete in the order sense, meaning it lacks the "gaps" present in the rational numbers \mathbb{Q}. For instance, the set A = \{ q \in \mathbb{Q} \mid q > 0, \, q^2 < 2 \} is bounded above in \mathbb{Q} (e.g., by 2), but it has no least upper bound within \mathbb{Q}, as any proposed rational upper bound can be exceeded by another rational in A. This failure highlights why \mathbb{Q} is incomplete, necessitating the construction of \mathbb{R} to satisfy the property. The least upper bound property implies the absence of gaps by guaranteeing the existence of limits for certain Dedekind-like cuts in \mathbb{Q}. Consider the set S = \{ q \in \mathbb{Q} \mid q^2 < 2 \}; it is bounded above but lacks a supremum in \mathbb{Q}. In \mathbb{R}, however, \sup S = \sqrt{2} exists by the axiom. To see why this fills the gap, let s = \sup S. Then s > 0 and s^2 \leq 2, but if s^2 < 2, there would exist a rational r > s with r^2 < 2, contradicting the least upper bound status of s. Similarly, s^2 > 2 leads to a contradiction by finding a smaller upper bound. Thus, s^2 = 2, confirming \sqrt{2} \in \mathbb{R}. This process extends to other irrationals, ensuring \mathbb{R} contains all such limits without gaps. The least upper bound property is equivalent to several other formulations of completeness. One such equivalence is to the monotone convergence theorem: every non-decreasing sequence of real numbers that is bounded above converges to a real number. To see the implication from least upper bound, let (a_n) be non-decreasing and bounded above; then s = \sup \{ a_n \mid n \in \mathbb{N} \} exists, and a_n \to s since for any \epsilon > 0, there is N such that s - \epsilon < a_N \leq a_n for n \geq N. The converse holds by constructing the supremum from the limit. Another equivalence is to the nested interval theorem: if I_n = [a_n, b_n] is a sequence of closed intervals with I_{n+1} \subseteq I_n and b_n - a_n \to 0, then \bigcap_n I_n \neq \emptyset. The least upper bound property yields a point in the intersection via \sup \{ a_n \} = \inf \{ b_n \}, and the converse proves the property using nested intervals to approximate suprema. An alternative axiomatization of completeness, closely related to the monotone convergence theorem, posits that every increasing bounded sequence in \mathbb{R} converges. This formulation directly captures sequential completeness in the order topology and is equivalent to the least upper bound property, as the supremum of the range serves as the limit, and any supremum can be approximated by an increasing sequence of rationals or reals. The least upper bound property has significant implications for analysis. It implies the Bolzano-Weierstrass theorem: every bounded sequence in \mathbb{R} has a convergent subsequence, proved by applying the property iteratively to construct a monotonic subsequence converging to the supremum of tail sets. Furthermore, it yields the Heine-Borel theorem: every closed and bounded subset of \mathbb{R} is compact, as the property ensures that any open cover has a finite subcover via the finite intersection property for nested closed intervals. These results underpin much of real analysis, from continuity to integration.

Tarski's Axiomatization

Alfred Tarski developed a first-order axiomatization of real closed fields in the 1930s, presented in his 1941 work Introduction to Logic and to the Methodology of Deductive Sciences. This system offers a rigorous logical foundation for the arithmetic of real closed fields, structures that generalize the real numbers by including roots of all odd-degree polynomials and satisfying the intermediate value property for polynomials, while being formally real (sums of squares are nonnegative). The axioms consist of the ordered field axioms augmented by first-order statements ensuring real closure: every positive element has a square root, and every odd-degree polynomial with coefficients in the field has a root. Tarski also explored an alternative presentation using only addition and the order relation as primitives, defining multiplication via Eudoxus-style proportions (preservation of inequalities under scaling by natural numbers). This additive axiomatization includes properties like density, asymmetry of order, and a continuity condition implying Dedekind completeness for the structure. A key feature of Tarski's approach is quantifier elimination: every formula in the theory is equivalent to a quantifier-free one, enabling algorithmic decidability—any sentence in the language of ordered fields can be decided mechanically. This contrasts with incomplete first-order attempts and provides a complete theory for real closed fields. However, the theory is not categorical; it admits non-isomorphic models, such as the real algebraic numbers or non-Archimedean extensions. The real numbers ℝ are the unique (up to isomorphism) Archimedean real closed field, where Archimedean ensures no infinitesimals and density of rationals, though expressing this fully requires axioms beyond first-order logic. Tarski's framework avoids explicit second-order completeness postulates while capturing essential continuity properties for definable sets, underpinning applications in real algebraic geometry, o-minimal structures, and automated theorem proving in analysis.

Standard Constructions from Rationals

Dedekind Cuts

In 1872, Richard Dedekind introduced a construction of the real numbers as certain partitions of the rational numbers, known as Dedekind cuts, to provide a rigorous arithmetic foundation for the continuum independent of geometric intuitions. This approach addresses the incompleteness of the rationals by filling gaps through set-theoretic means, yielding an ordered field that satisfies the least upper bound property. A Dedekind cut is a partition of the set of rational numbers \mathbb{Q} into two non-empty, disjoint subsets L and U such that L \cup U = \mathbb{Q}, every element of L is strictly less than every element of U, and L has no greatest element. The set L serves as the lower set (all rationals less than the "cut point"), and U as the upper set (all rationals greater than or equal to it). For cuts corresponding to rational numbers, U has a least element, whereas for irrational cuts, U has no least element. The collection of all such cuts, denoted \mathbb{R}_D, forms the Dedekind reals, with the rational numbers embedded via the cuts L_q = \{r \in \mathbb{Q} \mid r < q\} and U_q = \{r \in \mathbb{Q} \mid r \geq q\} for each q \in \mathbb{Q}, where U_q has least element q. Arithmetic operations on cuts are defined to preserve the field structure. For addition, given cuts (L_1, U_1) and (L_2, U_2), the sum is the cut (L_1 + L_2, U_1 + U_2) where L_1 + L_2 = \{q_1 + q_2 \mid q_1 \in L_1, q_2 \in L_2\} and the upper set is the complement. Multiplication is defined first for positive cuts (where $0 \in U) as (L_1 \cdot L_2, U_1 \cdot U_2) with L_1 \cdot L_2 = \{q_1 q_2 \mid q_1 \in L_1, q_2 \in L_2, q_1 > 0, q_2 > 0\} \cup \{q \in \mathbb{Q} \mid q \leq 0\}, then extended to all cuts using additive inverses and the fact that negatives correspond to cuts with $0 \in L. These operations ensure \mathbb{R}_D is a field, with the embedding of \mathbb{Q} preserving addition and multiplication. The order on \mathbb{R}_D is defined by (L_1, U_1) < (L_2, U_2) if and only if L_1 \subsetneq L_2 (equivalently, U_2 \subsetneq U_1), inducing a total order that extends the order on \mathbb{Q}. This order is linear and compatible with the field operations, making \mathbb{R}_D an ordered field. For example, the cut corresponding to \sqrt{2} is given by L = \{q \in \mathbb{Q} \mid q < 0\} \cup \{q \in \mathbb{Q} \mid q \geq 0 \land q^2 < 2\} and U = \{q \in \mathbb{Q} \mid q \geq 0 \land q^2 \geq 2\}, which partitions \mathbb{Q} without a maximum in L or minimum in U, representing an irrational "gap" filled by the real. The completeness of \mathbb{R}_D follows directly from the construction: for any non-empty subset S \subseteq \mathbb{R}_D bounded above, the union of the lower sets of cuts in S forms a new lower set whose cut is the least upper bound of S. This least upper bound property ensures every Cauchy sequence of rationals converges and resolves all Dedekind incompletenesses in \mathbb{Q}. The Dedekind reals \mathbb{R}_D are isomorphic to the standard real numbers \mathbb{R} as complete ordered fields. The embedding of \mathbb{Q} is order-preserving and dense in \mathbb{R}_D, meaning between any two distinct cuts there exists a rational cut. Moreover, \mathbb{R}_D satisfies the Archimedean property: for any cuts x, y > 0, there exists a natural number n such that n > y/x, inherited from \mathbb{Q} and preserved under the order. These properties confirm the uniqueness up to isomorphism of any complete Archimedean ordered field.

Cauchy Sequences

The construction of the real numbers via Cauchy sequences identifies each real number with an equivalence class of Cauchy sequences of rational numbers, providing an analytic approach that completes the rational numbers with respect to their natural metric. This method, developed by Georg Cantor and Karl Weierstrass in the late 19th century, emphasizes limits and sequential convergence to fill the gaps in the rationals. A Cauchy sequence in the rational numbers \mathbb{Q} is a sequence (q_n)_{n=1}^\infty with q_n \in \mathbb{Q} for all n such that for every \varepsilon > 0 with \varepsilon \in \mathbb{Q}, there exists N \in \mathbb{N} satisfying |q_m - q_n| < \varepsilon for all m, n > N. This condition ensures that the terms become arbitrarily close as the index increases, capturing the intuitive notion of sequences that "settle down" without presupposing a complete space. Two Cauchy sequences (q_n) and (r_n) in \mathbb{Q} are equivalent, written (q_n) \sim (r_n), if \lim_{n \to \infty} (q_n - r_n) = 0, meaning that for every \varepsilon > 0, there exists N \in \mathbb{N} such that |q_n - r_n| < \varepsilon for all n > N. This equivalence relation partitions the set of all Cauchy sequences of rationals into classes, where each class groups sequences that approach the same limit in a precise sense. The set of real numbers \mathbb{R} is then defined as the quotient set of these equivalence classes, with each real number denoted by [(q_n)] = \{ (r_n) \mid (r_n) \sim (q_n) \}. Arithmetic operations on \mathbb{R} are induced pointwise from those on \mathbb{Q}: for equivalence classes [(q_n)] and [(r_n)], define [(q_n)] + [(r_n)] = [(q_n + r_n)], \quad [(q_n)] \cdot [(r_n)] = [(q_n \cdot r_n)]. These operations are well-defined, as equivalent sequences yield equivalent results under addition and multiplication, and they satisfy the field axioms inherited from \mathbb{Q}. The embedding of \mathbb{Q} into \mathbb{R} maps each rational q to the equivalence class of the constant sequence (q, q, q, \dots), ensuring that \mathbb{Q} is a dense subfield of \mathbb{R}. A natural metric on \mathbb{R} arises from the construction: the distance between [(q_n)] and [(r_n)] is d([(q_n)], [(r_n)]) = \lim_{n \to \infty} |q_n - r_n|, which exists and is finite because (q_n - r_n) is a Cauchy sequence in \mathbb{Q} (differences of Cauchy sequences are Cauchy). This metric makes \mathbb{R} the completion of \mathbb{Q}, where every Cauchy sequence in \mathbb{Q} converges in \mathbb{R}. A representative example is the irrational number e, constructed via the Cauchy sequence of partial sums s_n = \sum_{k=0}^n \frac{1}{k!} for n \in \mathbb{N}, where each s_n \in \mathbb{Q} and (s_n) is Cauchy because the remainder \sum_{k=n+1}^\infty \frac{1}{k!} < \frac{1}{n!} for large n, ensuring |s_m - s_n| < \varepsilon for sufficiently large \min(m,n). (Note: While Stack Exchange verifies the mathematical property, the series originates from Euler's work on exponentials.) The completeness of \mathbb{R} follows directly from the construction: consider a Cauchy sequence (x_k)_{k=1}^\infty in \mathbb{R}, where each x_k = [(q_n^{(k)})]_{n=1}^\infty with (q_n^{(k)})_n a Cauchy sequence in \mathbb{Q}. Since (x_k) is Cauchy, there exists a representative rational sequence (b_k) such that |b_k - q_n^{(k)}| < 1/k for all n \geq k, making (b_k) Cauchy in \mathbb{Q}. The equivalence class [(b_k)] then serves as the limit of (x_k) in \mathbb{R}, as (x_k) \sim (b_k) and every Cauchy sequence in \mathbb{Q} defines a real. This proves that \mathbb{R} is complete, meaning every Cauchy sequence converges, realizing the least upper bound property through sequential limits.

Alternative and Non-Standard Constructions

Eudoxus Reals

The method of Eudoxus, dating to approximately 370 BCE, offers an early approach to handling continuous magnitudes through ratios and proportions, laying groundwork for a proto-construction of the positive real numbers without relying on modern set theory. In this framework, two magnitudes A and B (of the same kind, such as lengths) satisfy A > B if there exist positive integers m and n such that mA > nB \geq (m-1)A, ensuring a comparison based solely on integer multiples that avoids direct measurement. Equality of ratios A:B = C:D holds if, for all positive integers k and l, the relations kA > lB, kA = lB, or kA < lB correspond exactly to kC > lD, kC = lD, or kC < lD, respectively; a ratio A:B > C:D if there exist k, l with kA > lB but kC \leq lD. This relational definition, preserved in Euclid's Elements Book V, enables rigorous treatment of incommensurable quantities like \sqrt{2} by focusing on comparative inequalities rather than numerical values. The Archimedean basis of Eudoxus' method underpins its effectiveness, positing that the positive integers are sufficiently dense among positive magnitudes: for any A > 0 and B > 0, there exists a positive integer n such that nB > A, precluding infinitesimals or non-Archimedean elements and ensuring all ratios are well-defined within a bounded scale. This property guarantees that magnitudes can be approximated arbitrarily closely by integer multiples, mirroring the density of rationals in the reals while operating purely on integers. A modern formalization reconstructs the Eudoxus reals directly from the integers as equivalence classes of almost-homomorphisms f: \mathbb{Z} \to \mathbb{Z}, where f satisfies |f(p+q) - f(p) - f(q)| \leq C for some fixed integer C and all p, q \in \mathbb{Z}; two such functions f \sim g if |f(p) - g(p)| \leq D for some D and all p. The set of Eudoxus reals \mathbb{E} forms the quotient group of these classes under addition, with positive elements $$ defined as those where f(m) > 0 for infinitely many positive integers m. This embeds the positive rationals (via linear functions f(k) = qk) and extends to irrationals through the additive structure, achieving completeness via the exhaustion principle inherent in the bounded perturbations. Multiplication is introduced compatibly with the order, yielding a full ordered field isomorphic to the reals. This construction connects to Dedekind cuts by associating an Eudoxus real $$ with the cut partitioning positive rationals r = m/n such that m \leq f(n), where rational ratios approximate the cut's boundary; in the Archimedean setting, this mapping is bijective to the positive reals. As an illustration, \sqrt{2} is represented by the class $$ where f(n) = \lfloor n \sqrt{2} \rfloor, an almost-homomorphism whose values provide integer bounds approximating the irrational: for instance, the continued fraction convergents yield ratios like $7/5 < \sqrt{2} < 17/12, refined exhaustively by checking integer multiples to squeeze the interval arbitrarily close. Despite its elegance, the Eudoxus approach has limitations as a full construction: it primarily embeds the positive reals as an ordered additive group, requiring additional structure for negatives, zero, and multiplicative inverses, and does not explicitly partition sets like Dedekind cuts, focusing instead on relational comparisons among positives.

Hyperreal-Based Construction

The hyperreal numbers, denoted \mathbb{R}^*, are constructed as the ultraproduct of countably many copies of the rational numbers \mathbb{Q} with respect to a non-principal ultrafilter \mathcal{U} on the natural numbers \mathbb{N}. Specifically, elements of \mathbb{R}^* are equivalence classes of sequences (q_n)_{n \in \mathbb{N}} where q_n \in \mathbb{Q}, with two sequences (q_n) and (r_n) equivalent if \{n \in \mathbb{N} \mid q_n = r_n\} \in \mathcal{U}. Field operations are defined componentwise: [(q_n)] + [(r_n)] = [(q_n + r_n)], [(q_n)] \cdot [(r_n)] = [(q_n \cdot r_n)], and the order is given by [(q_n)] < [(r_n)] if \{n \in \mathbb{N} \mid q_n < r_n\} \in \mathcal{U}. This yields a non-Archimedean ordered field extending \mathbb{Q}, containing infinitesimal and infinite elements. The transfer principle ensures that any first-order sentence in the language of ordered fields that holds in \mathbb{Q} also holds in \mathbb{R}^*, as a direct consequence of Łoś's theorem on ultraproducts. This principle, foundational to nonstandard analysis as developed by Abraham Robinson, allows properties of the rationals—such as being an ordered field—to carry over to the hyperreals without assuming the existence of the reals a priori. The standard real numbers \mathbb{R} are obtained as the image of the standard part function st: \mathbb{R}^* \to \mathbb{R}, restricted to the finite hyperreals, defined as \{x \in \mathbb{R}^* \mid |x - n| < 1 \text{ for some } n \in \mathbb{Z}\}. For such a finite x, st(x) is the unique standard rational (hence real) r such that x - r is infinitesimal, meaning |x - r| < q for every positive standard rational q. Formally, this can be viewed as the quotient of the ring of finite hyperrationals by the ideal of infinitesimals. The embedding of \mathbb{Q} into \mathbb{R} is given by constant sequences, and \mathbb{R} forms an ordered subfield of \mathbb{R}^* under the induced operations and order, where positives in \mathbb{R}^* define the order. The completeness of \mathbb{R} arises from the transfer principle extended to monadic second-order properties, which allows the least upper bound axiom to hold in \mathbb{R} by leveraging the saturation properties of \mathbb{R}^* derived from the ultrafilter. This construction equips \mathbb{R} with the Dedekind completeness needed to distinguish it from \mathbb{Q}, ensuring every nonempty subset of \mathbb{R} bounded above has a least upper bound. This approach offers advantages in handling infinitesimals uniformly within an ordered field framework, providing an intuitive extension for analysis while yielding \mathbb{R} isomorphic to the standard reals via the standard part map. Unlike constructions that extend the reals directly, this method builds \mathbb{R} as a subfield from \mathbb{Q} alone, avoiding circularity.

Surreal Number Construction

The surreal numbers, denoted as No, form a proper class of numbers constructed by John Horton Conway through a transfinite recursive process inspired by combinatorial game theory. This construction yields a totally ordered field that universally embeds every ordered field, with the real numbers appearing as the initial segment of surreals born on or before the first infinite ordinal day ω. Each surreal number is defined as a pair {L \mid R}, where L and R are disjoint sets of previously constructed surreals satisfying the condition that every element of L is less than every element of R, ensuring no overlap or crossing in the order. The recursion proceeds over "days" indexed by ordinals: on day 0, the number 0 is born as {\emptyset \mid \emptyset}; on day 1, integers like 1 = {0 \mid } and -1 = {\mid 0} emerge; and subsequent days generate new numbers from all earlier ones. Within this hierarchy, the real numbers are precisely those surreals with finite birthdays, beginning with the integers on early finite days, followed by dyadic rationals (fractions with denominator a power of 2) such as 1/2 = {0 \mid 1} on day 2, and progressively denser fillings that complete the reals by day ω through transfinite induction. For instance, non-dyadic rationals like 1/3 arise on day ω as {1/4, 5/16, 21/64, \dots \mid 1/2, 3/8, 11/32, \dots }, while irrationals require infinite sets of options born across countable ordinals less than ω. This temporal embedding ensures the reals form a dense linear order without gaps, as every Dedekind cut in the rationals is realized by a surreal born at the least ordinal exceeding the birthdays of its approximating options. Arithmetic operations on surreals are defined recursively using the option sets: addition of x = {X_L \mid X_R} and y = {Y_L \mid Y_R} is x + y = {X_L + y, x + Y_L \mid X_R + y, x + Y_R}, with similar recursive definitions for multiplication involving sign considerations and the absolute value. Each surreal admits a unique "simplest" representative in normal form, a finite or transfinite sum \sum \omega^{y_i} r_i where r_i are dyadic rationals between -1 and 1, facilitating computation and preserving field properties. The order on No is defined by x > y if no right option of y exceeds x and no left option of x falls below y, inducing a total order that embeds the rationals densely and extends to all surreals, making No the largest possible ordered field in the sense of order-embedding every countable ordered field. A concrete example is the surreal representation of \sqrt{2}, which fills the cut between rational approximations from below and above: \sqrt{2} = {1, 5/4, 11/8, \dots \mid \dots, 23/16, 3/2 }, where the left set consists of rationals less than \sqrt{2} and the right set of those greater, born progressively on finite days until the full number emerges by day ω. The completeness of the real segment follows from transfinite induction over the birthdays: for any partition of earlier surreals into lower and upper classes with no greatest lower or least upper element, a new surreal is born on the next ordinal to fill the gap, ensuring the reals satisfy the least upper bound property without omissions. Conway's 1976 framework in On Numbers and Games provides the rigorous proof of this embedding, demonstrating that the map sending each real to its surreal normal form is an order-isomorphism onto the surreals of birthday less than ω.

Other Methods

One lesser-known algebraic approach to constructing the real numbers begins with the field of rational numbers \mathbb{Q} and iteratively adjoins square roots to form a tower of quadratic extensions. This process generates the field of constructible numbers, which is dense in the real numbers but does not include all transcendentals. To obtain the full real numbers, one completes this field with respect to the archimedean absolute value, yielding a complete ordered field isomorphic to \mathbb{R}. This method highlights the algebraic structure underlying the reals while relying on completion for completeness. In 1854, Bernhard Riemann proposed a geometric construction of the real numbers in his habilitation lecture "On the Hypotheses which lie at the Foundations of Geometry." To define points in continuous manifolds without presupposing the reals, Riemann described a real number as the intersection of a sequence of nested intervals in the plane, where the lengths of the intervals tend to zero. This approach embeds the construction within synthetic geometry, using geometric intervals to represent limits and avoiding explicit use of infinite sets or arithmetic completeness until later axiomatizations. The method influenced subsequent foundational work by providing a coordinate-free way to model the continuum. David Hilbert's 1899 axiomatization in Grundlagen der Geometrie yields the real numbers as the coordinate field for the Euclidean plane under synthetic geometry. Hilbert's axioms divide into groups for incidence, order, congruence, parallelism, and completeness. The order and completeness axioms ensure that the field of coordinates is an archimedean ordered field satisfying the least upper bound property, which is unique up to isomorphism and thus isomorphic to \mathbb{R}. This construction demonstrates that the reals emerge naturally as the scalars for a rigorous synthetic treatment of geometry, independent of analytic coordinates. Topological completions offer another perspective on constructing the real line. The order topology on the rationals \mathbb{Q}, generated by open intervals, has the real numbers \mathbb{R} as its Dedekind or Cauchy completion, filling all gaps in the dense linear order. Variant topological models, such as the long line—a linearly ordered topological space that extends the real line by attaching uncountably many copies of [0,1)—illustrate non-standard completions but serve primarily as models highlighting properties like local Euclidean-ness while globally differing from \mathbb{R}. These approaches emphasize the real line as the unique separable complete metric space that is locally compact and connected. A key result unifying these constructions is that all complete ordered fields are isomorphic to the real numbers \mathbb{R}. This theorem, first proved by E. V. Huntington in 1903, states that any two Dedekind-complete ordered fields admit a unique order-preserving field isomorphism. The proof proceeds by showing that any such field is archimedean, contains \mathbb{Q} densely, and the isomorphism maps rationals to rationals while extending continuously via suprema. Consequently, different constructions—whether via cuts, sequences, or geometry—yield fields isomorphic to \mathbb{R}, confirming the uniqueness of the complete ordered field up to isomorphism. Modern variants include computer-assisted formalizations, which verify constructions within proof assistants like Coq to ensure rigor and avoid hidden assumptions. For instance, Geuvers and Niqui (2000) constructed constructive real numbers in Coq using axioms for apartness and order, proving categoricity relative to these axioms and providing a model via Cauchy sequences of rationals. Later works, such as Boldo et al. (2012), formalized the set of real algebraic numbers as the real closure of \mathbb{Q} in Coq, complete with proofs of field operations and order properties. More recently, Dou and Yu (2025) provided a machine proof in Coq of the filter-method construction for real numbers, using a non-principal arithmetical ultrafilter and Morse–Kelley set theory to extend non-standard naturals, integers, and rationals, verifying Archimedeanness and completeness. These formalizations, developed post-2000, expand on traditional methods by enabling machine-checked derivations and applications in verified computing.

References

  1. [1]
    [PDF] Math 118: Real Numbers - UCSB Mathematics Department
    The foundation of these concepts are the axioms of set theory and the Peano axioms for natural numbers. These are interesting topics in the foundations of.
  2. [2]
    [PDF] NOTES ON REAL NUMBERS In these notes we will construct the set ...
    Finally we come to the construction of the real numbers, which is unfortunately not as simple as defining an equivalence relation on Q × Q or some similar set. ...
  3. [3]
    [PDF] Supplement. The Real Numbers are the Unique Complete Ordered ...
    Oct 2, 2024 · A, Dedekind was the first to give a construction of the real numbers in the 1850s, though he did not publish his results until 1872 ...
  4. [4]
    [PDF] The Rational Numbers
    In a field, we can also define subtraction and division. Definition Suppose and are members of a field . B. C. J a) We define the difference B ...
  5. [5]
    A Geometric Proof That The Square Root Of Two Is Irrational
    Aug 28, 1997 · The essence of the proof is this. If the square root of two were a rational number, you could write it as a fraction a/b in lowest terms, where ...
  6. [6]
    [PDF] 10. Algebraic closure Definition 10.1. Let K be a field ... - UCSD Math
    K, is an algebraic field extension L/K such that every polynomial in K[x] splits in L. We say that K is algebraically closed if K = ¯K. Lemma 10.2.Missing: rationals | Show results with:rationals
  7. [7]
    [PDF] IRRATIONALITY OF π AND e 1. Introduction Numerical estimates for ...
    The first proof that π is irrational is due to Lambert [7] in 1761. His proof involves an analytic device that is not covered in calculus courses: continued ...
  8. [8]
    [PDF] An intermediate value theorem for Q - Antonella Perucca
    The intermediate value theorem does not hold for continuous functions on closed intervals of rational numbers, not even if we restrict to rational ...
  9. [9]
    A Contemporary Look at Zeno's Paradoxes - Dartmouth Mathematics
    In this paradox, Zeno argues that an arrow in flight is always at rest. At any given instant, he claims, the arrow is where it is, occupying a portion of space ...
  10. [10]
    Eudoxus (408 BC - Biography - MacTutor History of Mathematics
    Another remarkable contribution to mathematics made by Eudoxus was his early work on integration using his method of exhaustion. This work developed ...
  11. [11]
    Arabic mathematics - MacTutor - University of St Andrews
    Al-Khwarizmi's successors undertook a systematic application of arithmetic to algebra, algebra to arithmetic, both to trigonometry, algebra to the Euclidean ...
  12. [12]
    Decimal Arithmetic - Muslim HeritageMuslim Heritage
    Sep 8, 2005 · Al-Khawarizmi's arithmetic was the first systematic treatment of arithmetical operations. Al-Khwarizmi discussed the place-value system and ...
  13. [13]
    (PDF) Wallis-Euler type infinite products - ResearchGate
    Wallis discovered in 1655 a marvelous infinite product for Pi. Around eighty years later, Euler developed the product representation for the sine function.
  14. [14]
    How Isaac Newton Discovered the Binomial Power Series
    Aug 31, 2022 · He recounts how he made a major discovery equating areas under curves with infinite sums by a process of guessing and checking.
  15. [15]
    [PDF] s heat conduction equation: History, influence, and connections
    Fourier's heat conduction equation, formulated in the 19th century, describes heat conduction in solids and is used for diffusion problems in many fields.
  16. [16]
    Bolzano's Intermediate Value Theorem
    In 1817 Bernhard Bolzano published a very modern proof of the Intermediate Value Theorem (IVT) for continuous functions.
  17. [17]
    Cauchy sequences and Cauchy completions
    The construction of the real numbers from the rationals via equivalence classes of Cauchy sequences is due to Cantor [Ca1872] and Méray [Me1869]. In fact, ...Missing: 1850s- 1870s<|separator|>
  18. [18]
    Dedekind's Contributions to the Foundations of Mathematics
    Apr 22, 2008 · Like Dedekind, Cantor starts with the infinite set of rational numbers; his construction again relies essentially on the full power set of the ...
  19. [19]
    [PDF] Axioms for the Real Numbers
    Axioms for the Real Numbers. Field Axioms: there exist notions of addition and multiplication, and additive and multiplica- tive identities and inverses, so ...
  20. [20]
    [PDF] Math 117: Axioms for the Real Numbers
    Oct 11, 2010 · The axioms for real numbers fall into three groups, the axioms for fields, the order axioms and the completeness axiom. and g : F × F → F, g(x, ...
  21. [21]
    [PDF] math 3110: complete ordered field axioms - Cornell Mathematics
    First, we assume that we have binary operations + and ⋅ on the real numbers. We also assume that for each real number x there is a real number −x called its ...
  22. [22]
    [PDF] Axioms for Ordered Fields
    Axioms about Addition and Multiplication Axioms i) (Commutativity) a + b = b + a and ab = ba; ii) (Associativity) (a + b) + c = a + (b + c) and (ab)c = a(bc); ...
  23. [23]
    [PDF] Axioms for the Real Numbers
    ordered field axioms. These axioms form a basis for the algebraic operations and the order properties upon which the calculus is based. The set of real ...
  24. [24]
    Axioms for the Real Numbers
    The axioms for the real numbers are a foundation that can be used to prove other properties of R . Section 1.3 includes some examples of using the field axioms ...
  25. [25]
    Math 413 – Real numbers and ordered fields - Connect - Gonzaga
    Sep 26, 2020 · Defining the rational numbers.​​ We define the set of rational numbers as Q={ab∣a,b∈Z,b≠0}. For any a∈Z we define the equivalence between Z and Q ...
  26. [26]
    [PDF] Math 117: Axioms for the Real Numbers
    Oct 15, 2008 · The preceding proposition shows that the field Q of rational numbers is not a complete ordered field because it does not contain. √ p when p ...
  27. [27]
    [PDF] Section 1.3: The Completeness of the Real Numbers
    Oct 19, 2011 · Observe: The rational numbers do not form a complete ordered field (just an ordered field). Axiom of Completeness: The real number are complete.
  28. [28]
    Real Analysis: Examples 2.4.3(b) - MathCS.org
    If we consider the universe to consist only of rational numbers, then this set does not have a least upper bound. No number bigger than is the least upper ...
  29. [29]
    Lecture 3: Cantor's Remarkable Theorem and the Rationals' Lack of ...
    Introducing ordered sets, the least upper bound property, and the incompleteness of the rational numbers. Speaker: Casey Rodriguez.
  30. [30]
    [PDF] 1.3. The Completeness Axiom.
    Dec 1, 2023 · Prove the least upper bound of A is 1. Exercise 1.3.4. (a) Between any two real numbers, there is a rational number. (b) Between any two ...
  31. [31]
    [PDF] Completeness of R
    If the sequence (an) is monotone increasing and bounded above, then it is convergent. Proof. The set {an} being bounded it has a least upper bound, say s, by ...<|separator|>
  32. [32]
    3.2.1 The nested interval property - GitLab
    Notice that the nested-interval property is actually equivalent to the least-upper bound property. If the nested-interval property holds then that is ...
  33. [33]
    1.3: Sequences and Completeness
    ... Least Upper Bound Property, which states that. Every nonempty set of real numbers that is bounded above has a least upper bound. Recall: A number a is an upper ...
  34. [34]
    The Heine-Borel Theorem
    The Bolzano-Weirstrass Theorem says that every bounded, infinite subset of R has an accumulation point. The textbook proves this using the least upper bound ...Missing: Weierstrass | Show results with:Weierstrass
  35. [35]
    [PDF] MATH 510, Notes 5
    The Nested Interval Principle is equivalent to the Least Upper. Bound (or Greatest Lower Bound) property, and both equivalent to the so-called Completeness ...
  36. [36]
    [PDF] arXiv:2302.07078v4 [math.CA] 12 May 2025
    May 12, 2025 · the least upper bound property of the real field. A “designer” of a ... (BWP) Bolzano-Weierstrass Property of R. A bounded sequence has ...
  37. [37]
    [PDF] Project Gutenberg's Essays on the Theory of Numbers, by Richard ...
    ON CONTINUITY AND IRRATIONAL NUMBERS, and ON THE NATURE AND. MEANING OF NUMBERS. By R. Dedekind. From the German by W. W.. Beman. Pages, 115.
  38. [38]
    [PDF] MATH 162, SHEET 8: THE REAL NUMBERS 8A. Construction of the ...
    The Archimedean property of the real numbers. For every rational number q ∈ Q, define the corresponding real number as the Dedekind cut ... isomorphic to the real ...
  39. [39]
    Archimedean ordered fields are real - PlanetMath.org
    Mar 22, 2013 · Since F 𝔽 is an ordered field, it must have characteristic zero. Hence, the subfield generated by the multiplicative identity is isomorphic to ...<|control11|><|separator|>
  40. [40]
    [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 ...
  41. [41]
    [PDF] chapter 9: constructing the real numbers - CSUSM
    The main theorem of this chapter is that R, as constructed from Cauchy sequences, is a complete ordered field. 2. The real numbers. Our idea for constructing R ...
  42. [42]
    Sequence of partial sums of e in Q is a Cauchy sequence.
    May 21, 2016 · Sequence of partial sums of e in Q is a Cauchy sequence. Ask Question ... Proof that the series expansion for exp(1) is a Cauchy sequence ...Proof that the series expansion for exp(1) is a Cauchy sequenceShowing the sequence of partial sums of a power series is Cauchy ...More results from math.stackexchange.com
  43. [43]
    [PDF] Eudoxos and Dedekind: On the ancient Greek theory of ratios and its ...
    According to Aristotle, the objects studied by mathematics have no independent existence, but are separated in thought from the substrate.
  44. [44]
    [math/0405454] The Eudoxus Real Numbers - arXiv
    May 24, 2004 · This note describes a representation of the real numbers due to Schanuel. The representation lets us construct the real numbers from first principles.
  45. [45]
    Eudoxus Reals - Archive of Formal Proofs
    Oct 29, 2023 · In this project, we present a peculiar construction of the real numbers, called "Eudoxus reals", using Isabelle/HOL.
  46. [46]
    [PDF] Constructions of the real numbers
    We will concentrate on the construction of the set of hyperrational and later hyperreal numbers using the means which we have established so far, plus a new ...
  47. [47]
  48. [48]
    Infinity > Construction of Surreal Numbers (Stanford Encyclopedia of ...
    Conway's construction of surreal numbers (1976) is summarized in the following principle: If \(L\), \(R\) are two sets of surreal numbers, and no member of \(L ...
  49. [49]
    [PDF] An Introduction to Surreal Numbers - Whitman College
    May 8, 2012 · In this section we will build the surreal numbers, which we denote S, from the ground up based on two axioms defined by Conway. Axiom 1.
  50. [50]
    Surreal Numbers Are a Real Thing. Here's How to Make Them
    Feb 14, 2024 · Surreal numbers are created by adding values between two given preexisting numbers. If you look at 0 and 1, for example, 1 / 2 is in the middle, 1 / 4 is ...
  51. [51]
    quadratic closure - PlanetMath
    Mar 22, 2013 · A field K K is said to be quadratically closed if it has no quadratic extensions. In other words, every element of K K is a square.<|control11|><|separator|>
  52. [52]
    [PDF] Cantor and Continuity
    May 1, 2018 · Richard Dedekind, in his essay Stetigkeit und irrationale Zahlen [1872], provided his now well-known construction of the real numbers. While ...
  53. [53]
    Epistemology of Geometry - Stanford Encyclopedia of Philosophy
    Oct 14, 2013 · Independently, Hilbert also gave an example of a geometry meeting all the incidence axioms of two-dimensional projective geometry but in ...
  54. [54]
    [PDF] Constructive Reals in Coq: Axioms and Categoricity
    Abstract. We describe a construction of the real numbers carried out in the Coq proof assistant. The basis is a set of axioms for the constructive.Missing: post- | Show results with:post-
  55. [55]
    [PDF] Construction of real algebraic numbers in Coq - Hal-Inria
    Abstract. This paper shows a construction in Coq of the set of real algebraic numbers, together with a formal proof that this set has a struc-.Missing: post- | Show results with:post-