Fact-checked by Grok 2 weeks ago

Least-upper-bound property

The least-upper-bound property, also known as the supremum property or completeness axiom, states that in an ordered set, every non-empty subset that is bounded above possesses a least upper bound (supremum) within the set itself. This property is a defining feature of the real numbers \mathbb{R}, distinguishing them from the rational numbers \mathbb{Q}, which lack it—for instance, the set \{q \in \mathbb{Q} \mid q^2 < 2\} is bounded above in \mathbb{Q} but has no least upper bound there, as \sqrt{2} is irrational. Formally, for a non-empty subset S \subseteq \mathbb{R} bounded above, there exists a unique \sup S \in \mathbb{R} such that \sup S is an upper bound for S and no smaller real number serves as an upper bound. Introduced by mathematician Richard Dedekind in his 1872 essay "Stetigkeit und irrationale Zahlen" (Continuity and Irrational Numbers), the property arises naturally from the construction of the real numbers via , where each cut corresponds to a real number and ensures the absence of "gaps" in the number line. Dedekind's approach axiomatized completeness independently of geometric intuitions, providing a rigorous algebraic foundation for the reals as the unique complete ordered field up to isomorphism. This axiomatic formulation resolved foundational issues in 19th-century analysis, such as the treatment of irrational numbers and limits, by embedding the property as a postulate rather than a derived theorem. The least-upper-bound property is equivalent to other completeness formulations, including the monotone convergence theorem for sequences and the Cauchy completeness of \mathbb{R}, and it underpins key results in real analysis, such as the intermediate value theorem, the extreme value theorem, and the existence of Riemann integrals. Without it, many theorems of calculus would fail, as seen in the rationals where bounded monotone sequences may not converge. In broader mathematics, the property extends to concepts like complete lattices and metric spaces, influencing fields from topology to functional analysis.

Basic Concepts

Upper bounds and least upper bounds

In a partially ordered set (poset) (P, \leq), an upper bound for a subset S \subseteq P is an element u \in P such that s \leq u for all s \in S. This means u is greater than or equal to every element in S under the partial order. The set of all upper bounds for S, often denoted U(S), forms an upset in the poset. The least upper bound, also known as the supremum and denoted \sup S or \lub(S), is the minimal element of U(S) when it exists. Specifically, \sup S is an upper bound for S, and for every other upper bound v \in U(S), it holds that \sup S \leq v. In posets, the supremum may not exist for every nonempty subset, but when it does, it is unique due to the antisymmetry of the partial order: if two such least upper bounds existed, each would be less than or equal to the other, implying they are equal. If \sup S \in S, then \sup S is the maximum element of S, as it bounds S from above and belongs to the set itself. Conversely, the maximum, if it exists, serves as the supremum. However, the supremum need not belong to S; for instance, in the poset of integers (\mathbb{Z}, \leq) under the usual order, \sup \{1, 2, 3\} = 3, which is the maximum element in the set. In the poset of real numbers (\mathbb{R}, \leq), consider the subset S = \{-1/n \mid n = 1, 2, \dots \}; here, \sup S = 0, which is an upper bound not attained by any element in S. This illustrates how the supremum provides the "tightest" upper bound, even when no maximum exists.

Bounded subsets of ordered sets

In a partially ordered set (P, \leq), a nonempty subset S \subseteq P is said to be bounded above if there exists an element k \in P such that k \geq s for all s \in S; any such k is called an upper bound for S. Similarly, S is bounded below if there exists an element m \in P such that m \leq s for all s \in S; any such m is a lower bound for S. A subset S is bounded (or fully bounded) if it is both bounded above and bounded below. These notions of boundedness apply to various ordered structures, but the existence of a least upper bound (supremum) for every nonempty subset that is bounded above is not guaranteed without additional completeness properties. For instance, consider the set of rational numbers \mathbb{Q} under the standard order. The subset S = \{ q \in \mathbb{Q} \mid q^2 < 2 \} is bounded above—for example, by $2 \in \mathbb{Q}—yet it has no least upper bound in \mathbb{Q}, as any candidate rational upper bound can be undercut by a smaller rational still greater than all elements of S.[13] Another counterexample is the set T = { x \in \mathbb{Q} \mid x < \sqrt{2} }, which is bounded above in \mathbb{Q}(e.g., by3/2) but lacks a supremum within \mathbb{Q}, since \sqrt{2}$ is irrational and no rational serves as the least such bound. In contrast, some subsets are unbounded. For example, the set of natural numbers \mathbb{N} \subseteq \mathbb{R} (under the standard order on the reals) has no upper bound in \mathbb{R}, as for any real r, there exists n \in \mathbb{N} with n > r. These examples from illustrate how ordered sets without the can contain bounded subsets that fail to attain suprema internally, motivating the need for axioms in constructing structures like the real numbers.

Formal Statement

For the real numbers

The least-upper-bound property for the real numbers, also known as the or , states that every non-empty subset of \mathbb{R} that is bounded above has a least upper bound (supremum) in \mathbb{R}. This axiom asserts the of \sup S \in \mathbb{R} for any such subset S \subseteq \mathbb{R}, where \sup S is the smallest element greater than or equal to every element in S. This property implies that the real numbers \mathbb{R} are Dedekind-complete, meaning every non-empty subset bounded above possesses a supremum within \mathbb{R} itself. In contrast, the rational numbers \mathbb{Q} lack this property; for instance, the set S = \{ x \in \mathbb{Q} \mid x^2 < 2 \} is bounded above in \mathbb{Q} (e.g., by 2), but its supremum is \sqrt{2}, which is irrational and thus not in \mathbb{Q}, so no least upper bound exists in \mathbb{Q}. This gap in \mathbb{Q} highlights how the least-upper-bound property fills the "holes" in the rationals to form the complete ordered field \mathbb{R}. A basic corollary follows symmetrically: every non-empty subset of \mathbb{R} that is bounded below has a greatest lower bound (infimum) in \mathbb{R}. This duality arises because the infimum of a set S is the supremum of its negatives, \inf S = -\sup (-S), preserving the property under the order-reversing map x \mapsto -x. For a simple illustration, consider the half-open interval [0,1) \subseteq \mathbb{R}, which is bounded above by 1 (among other values). Its least upper bound is \sup [0,1) = 1, which belongs to \mathbb{R}, even though 1 is not an element of the set itself.

Generalization to partially ordered sets

In partially ordered sets, or posets, the least-upper-bound property generalizes the concept from totally ordered sets by requiring that every non-empty subset that admits an upper bound possesses a least upper bound, known as the , within the poset. Formally, for a poset (P, \leq), a subset S \subseteq P is bounded above if there exists some u \in P such that s \leq u for all s \in S; the poset has the least-upper-bound property, or is , if every such non-empty S has a \sup S \in P, which is the minimal element among all upper bounds of S. This condition ensures that the order structure captures "completeness" for bounded subsets without assuming totality of the order. Variations of this property adapt the requirement to specific classes of subsets. For instance, a poset is finitely sup-complete if every non-empty finite subset bounded above has a supremum, which aligns with the definition of a lattice but extends only to finite cases. Another variation applies to directed subsets—subsets where every finite subcollection has an upper bound within the subset—yielding the notion of an up-complete poset, where every directed subset has a supremum; this is particularly relevant in contexts like domain theory for computing. Examples illustrate these generalizations effectively. Complete lattices represent the strongest form, where every subset (not just bounded ones) has both a supremum and an infimum; the power set of any set, ordered by inclusion, forms a complete lattice, with unions as suprema and intersections as infima. Conditionally complete posets, such as the non-negative reals [0, \infty) under the standard order, satisfy the least-upper-bound property for all non-empty bounded-above subsets—e.g., the set \{x \in [0, \infty) \mid x^2 < 2\} has supremum \sqrt{2}—but lack suprema for unbounded subsets like [0, \infty) itself. The property relates closely to chain completeness, where the focus narrows to totally ordered subsets, or chains: a poset is chain-complete if every non-empty chain has a supremum and infimum. This is a weaker condition than full sup-completeness, as it ignores incomparable elements; for example, the rationals \mathbb{Q} are not chain-complete, failing sup-completeness overall due to Dedekind-incomplete chains like \{q \in \mathbb{Q} \mid q^2 < 2\}, which is a bounded chain without a supremum in \mathbb{Q}. Counterexamples highlight limitations in non-complete posets. Finite posets without infinite chains, such as an antichain of two incomparable elements \{a, b\} with no additional structure, trivially satisfy the property since the subset \{a, b\} has no upper bound. However, more complex finite posets lacking lattice structure, like a poset with elements a, b, c where a and b are incomparable and both below c, may have bounded subsets without least upper bounds if extended appropriately; broader counterexamples include the rationals as a poset, where bounded-above subsets like those approximating \sqrt{2} lack suprema in \mathbb{Q}.

Equivalent Formulations

Cauchy completeness

A in a (X, d) is a sequence (x_n)_{n=1}^\infty such that for every \epsilon > 0, there exists N \in \mathbb{N} with the property that d(x_m, x_n) < \epsilon whenever m, n > N. This definition captures the intuitive notion that the terms of the sequence become arbitrarily close to each other as n increases, without necessarily converging to a point in the space. A is called Cauchy complete (or simply complete) if every in the space converges to some limit in the space. For the real numbers \mathbb{R}, equipped with the standard metric d(x, y) = |x - y|, this property ensures that sequences "intended" to approach irrational limits, such as approximations to \sqrt{2}, succeed in doing so within \mathbb{R}. The least-upper-bound property of \mathbb{R} is equivalent to its Cauchy completeness: an Archimedean ordered field satisfies the least-upper-bound property if and only if every converges to a in the field. To see one direction, assume the least-upper-bound property holds; for a (x_n) in \mathbb{R}, consider the nested intervals formed by partial sums or by bounding the sequence between increasing lower bounds and decreasing upper bounds derived from its terms—the intersection of these closed intervals is nonempty by the nested interval theorem (itself a consequence of the least-upper-bound property), yielding a to which (x_n) converges. The converse direction, that Cauchy completeness implies the least-upper-bound property, follows from constructing that approximate the supremum of a , but requires more involved arguments establishing the existence of suprema via . A classic example illustrating the necessity of this equivalence is a Cauchy sequence of rational numbers approximating \sqrt{2}, such as the decimal expansions $1, 1.4, 1.41, 1.414, \dots; this sequence is Cauchy in \mathbb{Q} but does not converge to any rational limit, whereas in \mathbb{R} it converges to \sqrt{2}, highlighting how the least-upper-bound property completes \mathbb{Q} into \mathbb{R}.

Dedekind completeness

A Dedekind cut in an ordered field, such as the rational numbers \mathbb{Q}, is defined as a partition of the field into two non-empty subsets L and U such that L \cup U equals the entire field, L \cap U = \emptyset, every element l \in L is less than every element u \in U, and L has no greatest element. This structure captures the idea of a "gap" in the ordering, where no element in the field separates L and U precisely without violating the no-maximum condition in L. For cuts corresponding to irrational numbers, U has no least element, while for rational numbers, U has a least element. Dedekind completeness for an ordered field states that every Dedekind cut has a least upper bound within the field itself, meaning the supremum of L (or equivalently, the infimum of U) belongs to the field and fills the gap defined by the cut. In this formulation, the real numbers \mathbb{R} are Dedekind complete, as every such cut corresponds uniquely to an element of \mathbb{R}. The least-upper-bound property is equivalent to Dedekind completeness in s: an ordered field satisfies the least-upper-bound property if and only if it is Dedekind complete, ensuring that every non-empty bounded-above has a supremum, which in turn guarantees that every cut is "completed" by an element of . This equivalence holds because, for any bounded , one can construct a cut whose lower set is that (adjusted to be downward closed), and the supremum of the serves as the least upper bound of the cut. A classic example is the Dedekind cut corresponding to \sqrt{2}, where L = \{ q \in \mathbb{Q} \mid q < 0 \lor q^2 < 2 \} and U is the complement in \mathbb{Q}. This cut has no greatest element in L (since rationals approximating \sqrt{2} from below can be refined indefinitely) and no least element in U, and its least upper bound in \mathbb{R} is \sqrt{2}. This concept played a pivotal role in Richard Dedekind's 1872 construction of the real numbers from the rationals, where \mathbb{R} is defined as the set of all Dedekind cuts on \mathbb{Q}, endowed with the appropriate order, thereby ensuring completeness.

Axiomatic Role and Proofs

Status as a completeness axiom

The least-upper-bound (LUB) property serves as a foundational completeness axiom in the axiomatic construction of the real numbers \mathbb{R}, complementing the field axioms (which establish addition, multiplication, and their properties) and the order axioms (which define the total ordering \leq). Together, these axioms characterize \mathbb{R} as a complete ordered field, distinguishing it from the rational numbers \mathbb{Q}, which satisfy the field and order axioms but lack completeness. This independence from the for the natural numbers \mathbb{N} underscores that completeness is an additional postulate required to extend the arithmetic of \mathbb{N} and \mathbb{Q} to the continuum of \mathbb{R}, as constructions of \mathbb{R} (such as or ) explicitly incorporate the LUB to ensure all bounded subsets have suprema./01%3A_Tools_for_Analysis/1.05%3A_The_Completeness_Axiom_for_the_Real_Numbers) In axiomatic frameworks like for Euclidean geometry or (ZFC), the LUB property manifests as a second-order axiom, quantifying over all subsets of \mathbb{R} to assert the existence of least upper bounds for bounded nonempty sets. Unlike first-order axioms, which can be expressed solely in terms of individual elements, the LUB cannot be fully captured in first-order logic due to its universal quantification over arbitrary subsets, making it non-elementary and reliant on higher-order comprehension principles in set-theoretic foundations. Within ZFC, the reals are constructed via the power set axiom applied to \mathbb{N}, and the LUB holds by virtue of this construction, but the axiom's second-order nature highlights its role in bridging arithmetic and analysis without being derivable from purely first-order set-theoretic primitives. Any ordered field satisfying the Archimedean property (no infinitesimals or infinite elements) and the LUB axiom is necessarily complete and isomorphic to \mathbb{R}, establishing the uniqueness of the reals up to order-preserving field isomorphism. This characterization implies that \mathbb{R} is the maximal Archimedean ordered field, as any proper extension would violate either Archimedeanness or completeness. The LUB's existential assertion—that a supremum exists without specifying a method to construct it—renders it non-constructive, sparking debates in intuitionistic logic where such existence claims are rejected unless accompanied by an effective procedure, contrasting with classical mathematics where the axiom is accepted unconditionally. The LUB property is logically equivalent to the monotone convergence theorem, which states that every bounded increasing sequence in \mathbb{R} converges to its supremum. This equivalence demonstrates how the axiom underpins sequential completeness, providing a bridge to alternative formulations like Cauchy or Dedekind completeness while emphasizing its role in ensuring the continuity of the real line.

Derivation from Cauchy sequences

The least-upper-bound property for the real numbers \mathbb{R}, equipped with the standard metric induced by the absolute value, is equivalent to the Cauchy completeness of \mathbb{R}, meaning every Cauchy sequence in \mathbb{R} converges to a limit in \mathbb{R}. To show that the least-upper-bound property implies Cauchy completeness, consider a Cauchy sequence (x_n) in \mathbb{R}. First, note that every Cauchy sequence is bounded: for \epsilon = 1, there exists N \in \mathbb{N} such that |x_m - x_n| < 1 for all m, n \geq N, so the terms from n \geq N lie within [x_N - 1, x_N + 1], and the finite initial segment is bounded, hence the entire sequence is bounded. Define, for each k \in \mathbb{N}, the set S_k = \{x_n \mid n \geq k\}, which is nonempty and bounded above (by the bound of the sequence). Let z_k = \sup S_k, so z_k exists by the least-upper-bound property. The sequence (z_k) is decreasing because S_{k+1} \subseteq S_k, so z_{k+1} \leq z_k. Moreover, (z_k) is bounded below (e.g., by \inf S_1), so by the least-upper-bound property applied to the set of its terms, L = \inf_k z_k exists in \mathbb{R}. Similarly, define y_k = \inf S_k, yielding an increasing sequence (y_k) bounded above, with M = \sup_k y_k existing. Since y_k \leq z_k for all k, it follows that M \leq L. For each fixed n, y_n \leq x_n \leq z_n, and as k \to \infty, y_k \to M and z_k \to L by the definitions of infimum and supremum. To show M = L, observe that the Cauchy condition implies z_k - y_k \to 0 as k \to \infty: for \epsilon > 0, there exists N such that |x_m - x_n| < \epsilon for m, n \geq N, so \sup_{m \geq k} x_m - \inf_{m \geq k} x_m \leq \epsilon for k \geq N, hence z_k - y_k \leq \epsilon. Thus, L - M = \lim (z_k - y_k) = 0, so L = M. Convergence of (x_n) to L follows: for \epsilon > 0, choose N_1 such that z_k - y_k < \epsilon for k \geq N_1, and N_2 such that |x_m - x_n| < \epsilon for m, n \geq N_2. For n \geq \max(N_1, N_2), |x_n - L| \leq z_n - y_n < \epsilon, since y_n \leq x_n \leq z_n and y_n \to L, z_n \to L. Alternatively, |x_n - L| \leq \sup_{m \geq n} |x_m - x_n| < \epsilon for sufficiently large n, by the Cauchy condition. For the converse, that Cauchy completeness implies the least-upper-bound property, consider a nonempty subset S \subseteq \mathbb{R} that is bounded above by some b \in \mathbb{R}. Choose a \in S with a < b (possible since S is nonempty and bounded above). Construct nested closed intervals [a_n, b_n] as follows: set a_1 = a, b_1 = b; for n \geq 1, let m_n = (a_n + b_n)/2. If m_n is an upper bound for S, set a_{n+1} = a_n, b_{n+1} = m_n; otherwise, set a_{n+1} = m_n, b_{n+1} = b_n. Each interval contains elements of S or is refined accordingly, and the length b_n - a_n = (b - a)/2^{n-1} \to 0. The sequences (a_n) and (b_n) are both Cauchy because their difference halves each step, so |a_n - a_m| \leq |b_n - a_n| < (b - a)/2^{n-1} for m > n. By Cauchy completeness, a_n \to L and b_n \to L for some L \in \mathbb{R}. Thus, L is an upper bound for S: if there exists s \in S with s > L, then for large n, s > b_n, contradicting the construction where b_n remains an upper bound unless refined below s. Moreover, L is the least upper bound: if c < L is an upper bound, then for large n, c < a_n, but a_n \in S or the interval contains points of S exceeding c, a contradiction.

Applications in Analysis

Intermediate value theorem

The intermediate value theorem states that if f: [a, b] \to \mathbb{R} is a continuous function with f(a) < c < f(b), then there exists some x \in (a, b) such that f(x) = c. This theorem highlights a fundamental property of continuous functions on closed intervals, guaranteeing that the function attains every value between its endpoints. To prove the theorem using the least-upper-bound property of the real numbers, consider the set S = \{x \in [a, b] \mid f(x) \leq c\}. This set is nonempty since a \in S, and it is bounded above by b. By the least-upper-bound property, S has a supremum \xi = \sup S. Continuity of f at \xi implies that f(\xi) = c, as follows from analyzing the possible cases for f(\xi). Suppose f(\xi) < c. Then, by , there exists \delta > 0 such that f(x) < c for all x \in (\xi, \xi + \delta) \cap [a, b], implying that points greater than \xi belong to S, which contradicts \xi being the least upper bound of S. Similarly, if f(\xi) > c, continuity yields \delta > 0 such that f(x) > c for all x \in (\xi - \delta, \xi) \cap [a, b], meaning no points in that left neighborhood are in S, again contradicting \xi = \sup S. Thus, the only possibility is f(\xi) = c, and since f(a) < c < f(b), it follows that \xi \in (a, b). For example, consider f(x) = x^2 on [0, 1], where f(0) = 0 < 0.5 < 1 = f(1). The theorem ensures there exists x \in (0, 1) such that x^2 = 0.5, namely x = \sqrt{0.5}.

Bolzano–Weierstrass theorem

The Bolzano–Weierstrass theorem asserts that every bounded sequence in the real numbers \mathbb{R} has a convergent subsequence. This result highlights the completeness of \mathbb{R} provided by the least-upper-bound property, ensuring that boundedness implies the existence of accumulation points within \mathbb{R}. One standard proof constructs a convergent subsequence using nested closed intervals. Consider a bounded sequence \{x_n\}_{n=1}^\infty in \mathbb{R}, so there exists a closed interval I_1 = [a_1, b_1] with a_1 = \inf_n x_n and b_1 = \sup_n x_n, both finite by the least-upper-bound property applied to the sets of terms less than or greater than any bound. Divide I_1 at its midpoint into two closed subintervals of equal length, and select I_2 as the one containing infinitely many terms of the sequence; such an I_2 exists because the sequence is infinite and bounded. Inductively, for each k \geq 1, divide I_k = [a_k, b_k] at its midpoint and choose I_{k+1} as a closed subinterval of half the length containing infinitely many terms. Select indices n_k such that x_{n_k} \in I_k and n_{k+1} > n_k. The intervals \{I_k\} are nested and closed, with diameters tending to 0. By the nested interval theorem—derived from the least-upper-bound property, as the intersection is nonempty and singleton—the subsequence \{x_{n_k}\} converges to the unique point in \bigcap_k I_k. An alternative proof first establishes the existence of a monotonic , then uses boundedness to ensure . Every in \mathbb{R} has a monotonic : classify terms as "dominant" if no later term exceeds them in value; if infinitely many dominant terms exist, they form a decreasing , while if finitely many, select an increasing by greedily choosing terms larger than all previous ones. For a bounded , this monotonic is bounded above (or below), so by the least-upper-bound property, it converges to its supremum (or infimum). For example, consider the unbounded defined by alternating terms x_{2k-1} = k and x_{2k} = 1/k for k \in \mathbb{N}, which has a converging to 0 (the even terms) and another diverging to \infty (the odd terms). Restricting to a bounded version, such as x_n = \sin(n) (bounded in [-1, 1]), yields convergent to points in [-1, 1] by the , despite the full not converging.

Extreme value theorem

The extreme value theorem states that if f: [a, b] \to \mathbb{R} is continuous on the closed interval [a, b], then f attains both a maximum and a minimum value on [a, b]. That is, there exist points x_1, x_2 \in [a, b] such that f(x_1) \leq f(x) \leq f(x_2) for all x \in [a, b]. To prove this using the least-upper-bound property and the (established in the previous subsection), first show that the image f([a, b]) is bounded. Suppose it is unbounded above; then for each positive n, there exists x_n \in [a, b] such that f(x_n) > n. The sequence \{x_n\} is bounded, so by the , it has a convergent subsequence x_{n_k} \to x^* \in [a, b]. By , f(x_{n_k}) \to f(x^*), but f(x_{n_k}) > n_k \to \infty, implying f(x^*) = \infty, which is impossible in \mathbb{R}. Thus, f([a, b]) is bounded above (and similarly below). By the least-upper-bound property, f([a, b]) has a supremum M = \sup f([a, b]) \in \mathbb{R}, so f(x) \leq M for all x \in [a, b]. To show attainment of the maximum, for each positive integer n, there exists x_n \in [a, b] such that f(x_n) > M - 1/n (since M is the least upper bound). The sequence \{x_n\} is bounded in [a, b], so by the , it has a convergent subsequence x_{n_k} \to x^* \in [a, b]. By , f(x_{n_k}) \to f(x^*). Since f(x_{n_k}) > M - 1/n_k \to M and M is an upper bound, it follows that f(x^*) = M. The proof for the minimum is symmetric: the image is bounded below, has an infimum m = \inf f([a, b]), and choosing x_n with f(x_n) < m + 1/n yields a convergent to x_1 \in [a, b] where f(x_1) = m by a similar argument. For example, the f(x) = \sin x on [0, \pi] is continuous and attains its maximum value of 1 at x = \pi/2 and minimum value of 0 at x = 0 and x = \pi.

Heine–Borel theorem

The Heine–Borel theorem provides a characterization of compact subsets of \mathbb{R}^n: a subset K \subseteq \mathbb{R}^n is compact if and only if it is closed and bounded. This equivalence relies fundamentally on the least-upper-bound property of the real numbers, which ensures the existence of suprema and infima for bounded sets, enabling key steps in the proof. To sketch the proof that closed and bounded subsets are compact, consider first the case n=1 for a closed bounded interval [a, b]. Given an open cover \{U_\alpha\}_{\alpha \in A} of [a, b], define the set S = \{x \in [a, b] : [a, x] admits a finite subcover from \{U_\alpha\}\}. The set S is nonempty (since a \in S) and bounded above by b, so by the least-upper-bound property, t = \sup S exists in \mathbb{R}. A finite subcollection covers [a, t], and since some U_{\alpha_0} contains t and is open, it covers an interval around t, allowing extension of the subcover to points beyond t. If t < b, the closedness of [a, b] and the cover imply a contradiction unless t = b, yielding a finite subcover for the entire interval. For higher dimensions, the result follows by projecting onto coordinate axes (using the one-dimensional case) or via the finite product of compact intervals, where the least-upper-bound property underpins the nested interval constructions ensuring compactness in each factor. The least-upper-bound property plays a pivotal role by guaranteeing that bounded closed sets in \mathbb{R} possess suprema and infima, which extends to higher dimensions through coordinate-wise applications and facilitates the –Weierstrass theorem's generalization: every bounded infinite of \mathbb{R}^n has a point. This, in turn, supports the compactness criterion by ensuring closed bounded sets contain all points of their sequences. As a consequence, the theorem implies the for continuous functions on compact sets. A representative example illustrates the theorem: the closed interval [0, 1] \subseteq \mathbb{R} is compact, as any open cover admits a finite subcover by the above argument. In contrast, the open interval (0, 1) is bounded but not closed, hence not compact; consider the open cover \{U_n = (1/n, 1) : n = 2, 3, \dots \}, whose union is (0, 1), but no finite subcollection covers points near 0, as the left endpoints have a positive infimum.

Historical Development

19th-century origins

The , developed by the ancient Greek mathematician in the 4th century BCE, provided an early approach to approximating suprema by inscribing and circumscribing polygons around curved figures, allowing proofs of area and volume results through successive refinements that approached the least upper bound without directly invoking limits. This technique foreshadowed the need for a complete but remained geometric and did not address numerical completeness. In the , the foundations of , initially developed using rational numbers, encountered fundamental issues with irrational numbers, as these created gaps in the rational line that prevented rigorous handling of limits and . A classic illustration of such gaps is the irrationality of \sqrt{2}, known since but highlighted in early modern proofs like those building on , where no rational p/q satisfies p^2 = 2q^2, demonstrating that the rationals lack density in certain intervals and fail to fill the continuum required for analysis. These gaps became acute in calculus applications, such as proving the , motivating mathematicians to seek a complete number system. Joseph Fourier's work on trigonometric series in the , particularly in his 1822 treatise on conduction, exposed the incompleteness of by attempting to represent arbitrary functions—including discontinuous ones—via infinite series, leading to problems that demanded a denser, complete domain for real variables. This challenged the sufficiency of rational approximations in and spurred efforts toward rigorous foundations. In 1817, Bernard Bolzano recognized the need for completeness in his paper "Rein analytischer Beweis des Lehrsatzes," where he proved the intermediate value theorem by establishing a greatest lower bound property for certain sets, implicitly relying on a form of completeness without geometric appeals. Augustin-Louis Cauchy advanced this in his 1821 "Cours d'analyse," introducing the concept of Cauchy sequences—sequences where terms become arbitrarily close—and implicitly assuming the completeness of the reals to ensure their convergence, though without explicit axiomatization, as part of rigorizing limits in calculus.

Key contributions and formalizations

In the mid-19th century, played a pivotal role in rigorizing through his lectures at the University of during the 1850s and 1860s. He introduced the epsilon-delta definition of limits and , which provided a precise arithmetic foundation for without relying on geometric intuition or infinitesimals. Although Weierstrass did not explicitly state the least-upper-bound property, his framework implicitly depended on it to ensure the existence of limits for bounded sequences, forming a cornerstone for subsequent developments in . Richard Dedekind made an explicit formalization in 1872 with his pamphlet Stetigkeit und irrationale Zahlen, where he constructed the real numbers using —partitions of the rational numbers into two non-empty classes with no greatest element in the . This directly incorporates the least-upper-bound property as an equivalence, stating that every non-empty subset of reals bounded above has a least upper bound, thereby defining the of the reals independently of geometric notions. Dedekind's approach emphasized the arithmetic of the number system, influencing the axiomatic treatment of . Georg Cantor, in the 1870s, advanced the formalization through his work on and the construction of real numbers. In his 1872 paper "Über die Ausdehnung eines Satzes aus der Theorie der trigonometrischen Reihen," Cantor employed nested intervals—sequences of closed intervals with lengths approaching zero—to demonstrate the of the reals, ensuring that such intervals contain a unique limit point. This method, building on Cauchy sequences, implicitly relies on the least-upper-bound property to guarantee and the density of rationals in the reals, laying groundwork for transfinite . Later contributions included David Hilbert's axiomatization around 1900, where in his lectures and writings on the foundations of and , he incorporated a for the real numbers as an , ensuring every bounded non-empty set has a least upper bound to characterize the reals up to . Henri , in his 1904 Leçons sur l'intégration et la recherche des fonctions primitives, built upon the complete of reals, utilizing the least-upper-bound property to define outer measures and integrable functions rigorously. In the , Alfred extended this in the 1940s by proving the of the of real closed fields—ordered fields satisfying the and every positive element having a square root—where the least-upper-bound property holds categorically.

References

  1. [1]
    Least Upper Bound Axiom
    The least upper bound property can then be proved as a theorem about the set of real numbers, defined as the set of Dedekind cuts. Theorem (Least Upper Bound ...
  2. [2]
    [PDF] Lecture 17 - Section 10.1 Least Upper Bound Axiom Section 10.2 ...
    Mar 11, 2008 · Definition. Let S be a nonempty subset S of R. S is bounded above if ∃M ∈ R such that x ≤ M for all x ∈ S;. M is called an upper bound for S ...
  3. [3]
    Least Upper Bounds
    A real number β is a least upper bound or supremum of S if (i) β is an upper bound of S, and (ii) we have β≤b for every upper bound b of S. At most one real ...
  4. [4]
    Dedekind's Contributions to the Foundations of Mathematics
    Apr 22, 2008 · Richard Dedekind (1831–1916) was one of the greatest mathematicians ... “Continuity and Irrational Numbers”, in (Dedekind 1901a), pp. 1 ...
  5. [5]
    [PDF] An Introduction to Real Analysis John K. Hunter - UC Davis Math
    Abstract. These are some notes on introductory real analysis. They cover the properties of the real numbers, sequences and series of real numbers, limits.
  6. [6]
    [PDF] Dedekind's forgotten axiom and why we should teach it (and why we ...
    Mar 14, 2010 · What people usually call Dedekind completeness: Every bounded non-empty set S of real numbers has a least upper bound. It's a good completeness ...
  7. [7]
    [PDF] Basic Analysis: Introduction to Real Analysis
    May 23, 2025 · ... real number system, most importantly its completeness property, which is the basis for all that follows. We then discuss the simplest form ...
  8. [8]
    [PDF] Notes on Lattice Theory J. B. Nation University of Hawaii
    We say that x is the least upper bound for S if x is an upper bound for S and x ≤ y for every upper bound y of S. If the least upper bound of S exists,.
  9. [9]
    [PDF] Notes on Ordered Sets - UC Berkeley math
    Feb 27, 2012 · Definition 1.5 When minU(E) exists it is called the least upper bound of. E, or the supremum of E, and is denoted sup E. Dually, when max L(E) ...
  10. [10]
    [PDF] Math 127: Posets
    We define upper bound and supremum symmetrically. Note that this definition of lower/upper bound and infimum/supremum is identical to the definition given ...
  11. [11]
    [PDF] The supremum and infimum - UC Davis Math
    A set is bounded if it is bounded both from above and below. The supremum of a set is its least upper bound and the infimum is its greatest upper bound.
  12. [12]
    Real Numbers:Bounded Subsets - UTSA
    Nov 14, 2021 · A subset S of a partially ordered set P is called bounded above if there is an element k in P such that k ≥ s for all s in S. The element k is ...
  13. [13]
    Real Analysis: Definition: Lower and Greatest Lower Bound
    An element b is called a lower bound for the set X if every element in X is greater than or equal to b. If such a lower bound exists, the set X is called ...
  14. [14]
    [PDF] CHAPTER 6 - Max, Min, Sup, Inf - Purdue Math
    In fact, we know that √2=1.414213562 + . Each of the numbers 1.4, 1.41, 1.414, 1.4142, etc. is rational and has square less than 2. Their limit is √2. Thus, ...
  15. [15]
    [PDF] Axioms for the Real Numbers
    (P13) (Existence of least upper bounds): Every nonempty set A of real numbers which is bounded above has a least upper bound.
  16. [16]
    Supremum Axiom in Real Numbers: Definitions & Examples - Studocu
    THE SUPREMUM AXIOM FOR THE REAL NUMBERS the supremum axiom for the real numbers definitions. nonempty subset is bounded above if: any such is an upper bound ...
  17. [17]
    [PDF] Lecture Notes The Least Upper Bound Property and Intermediate ...
    The least upper bound property is a very fundamental one: it is actually the single axiom that distinguishes the set of rational numbers from the set of real ...<|control11|><|separator|>
  18. [18]
    Dedekind Completion - The Unapologetic Mathematician
    Dec 5, 2007 · We define an ordered field to be “Dedekind complete” if every nonempty set S with an upper bound has a least upper bound \sup S.
  19. [19]
    [PDF] Basic Analysis: Introduction to Real Analysis - IRL @ UMSL
    May 16, 2022 · The least-upper-bound property is sometimes called the completeness property or the Dedekind completeness property . As we will note in the next ...
  20. [20]
    [PDF] Completeness and compact generation in partially ordered sets
    A poset P is called conditionally complete if it is both conditionally U-complete and conditionally L-complete. We observe that every U-complete poset has the ...
  21. [21]
    Complete Lattice -- from Wolfram MathWorld
    A partially ordered set (or ordered set or poset for short) (L,<=) is called a complete lattice if every subset M of L has a least upper bound.
  22. [22]
    [PDF] Cauchy's Construction of R - UCSD Math
    To complete our discussion, we must show that this crazy set R of equivalence classes of Cauchy sequences really satisfies the least upper bound property!
  23. [23]
    [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.
  24. [24]
    [PDF] The least upper bound principle is equivalent to the Dedekind cut ...
    By the Dedekind property, either H has a last element or K has a first element. If H had a last element, say w, then w would also be an upper bound of H, a.Missing: completeness textbook
  25. [25]
    [PDF] The fabulous destiny of Richard Dedekind
    Sep 30, 2021 · As proved in Section 4 the D–completeness is equivalent to the least upper bound property2, as well as to the greatest lower bound property.
  26. [26]
    [PDF] An Axiomatic Construction of the Real Number System
    In this section, we construct the set of natural numbers, N, from a set of five axioms known as the Peano Axioms. The primary tool used is mathematical.
  27. [27]
    [PDF] PRIMITIVE TERMS GROUP I: FIELD AXIOMS
    (The least upper bound axiom) Every nonempty set of real numbers that has an upper bound has a least upper bound. An ordered field satisfying the least upper ...
  28. [28]
    Second-order and Higher-order Logic
    Dec 20, 2007 · The result is an infinite set of first-order axioms, assuring that any definable set that is non-empty and bounded has a least upper bound. The ...
  29. [29]
    [PDF] Supplement. The Real Numbers are the Unique Complete Ordered ...
    Oct 2, 2024 · A related work concerning a construction of the real numbers is due to Richard Dedekind (October 6, 1831–February 12, 1916). Dedekind was.
  30. [30]
    [DOC] Real Numbers - UCLA Mathematics
    An ordered field that satisfies the LUB axiom is called a complete ordered field. ... Archimedean ordered field is isomorphic to a subfield of the real numbers.
  31. [31]
    Constructive Mathematics | Internet Encyclopedia of Philosophy
    Every form of constructive mathematics has intuitionistic logic at its core ... For example, the classical least upper bound principle (LUB) is not constructively ...
  32. [32]
    [PDF] completeness of the real numbers - UTK Math
    Proof. We use the fact that the supremum property implies the Archimedean property and the monotone convergence property. Let (xn) be a Cauchy sequence ...
  33. [33]
    3.3 Cauchy and Completeness - GitLab
    3 Completeness implies least upper bound property. Theorem 3.3.22. The completeness of the reals implies that the reals satisfy the least upper bound property.
  34. [34]
    Bolzano's Intermediate Value Theorem
    Contemporary proofs are usually based on the Least Upper Bound Principle (LUBP), that a nonempty, bounded-above set of real numbers has a lub. In a previous ...
  35. [35]
    Intermediate Value Theorem
    Hence by the "completeness property" of the reals, the least upper bound, or supremum c = sup S exists. We claim that f(c)=k. Suppose our claim is false. First, ...
  36. [36]
    [PDF] The Bolzano-Weierstrass Theorem: Every sequence {xn}
    By the least-upper-bound property of the real numbers, s = supn(an) exists. Now, for every > 0, there exists a natural number N such that aN > s − , since ...
  37. [37]
    [PDF] 2.3. Bolzano-Weierstrass Theorem - East Tennessee State University
    Feb 5, 2024 · The Bolzano-Weierstrass Theorem gives a condition under which a set must have at least one limit point.
  38. [38]
    [PDF] Math 140A - Notes - UCI Mathematics
    Mar 7, 2025 · Theorem 2.41 (Bolzano–Weierstraß). Every bounded sequence has a convergent subsequence. Proof 1. Lemma 2.38 says there exists a monotone ...
  39. [39]
  40. [40]
  41. [41]
    proof of Heine-Borel theorem - PlanetMath
    Mar 22, 2013 · It goes by bisecting the rectangle along each of its sides. At the first stage, we divide up the rectangle A into 2n subrectangles.
  42. [42]
    [PDF] Exhaustion: From Eudoxus to Archimedes
    Apr 22, 2005 · Exhaustion is a method invented by Eudoxus to prove results about lengths, areas and volumes of geometrical objects.
  43. [43]
    Real numbers 2 - MacTutor History of Mathematics
    Cauchy did not have explicit formulations for the completeness of the real numbers. Among the forms of the completeness property he implicitly assumed are that ...
  44. [44]
    [PDF] On the history of analysis. The formation of definitions - arXiv
    In the middle of 19th century mathematics was in need of the theory of real numbers. Fourier series appeared, class of discontinuous and non-differentiable ...<|control11|><|separator|>
  45. [45]
    [PDF] Teaching Analysis with Bolzano's Original Text
    Jun 17, 2024 · Even though neither sets nor the completeness of the reals were formally defined until much later, we see in Bolzano's proof an exploration of ...
  46. [46]
    Continuity and Infinitesimals - Stanford Encyclopedia of Philosophy
    Jul 27, 2005 · Like Weierstrass and Dedekind, Cantor aimed to formulate an adequate definition of the real numbers which avoided the presupposition of their ...4. The Continuum And The... · 5. The Continuum And The... · 9. Smooth Infinitesimal...
  47. [47]
    [PDF] Cantor and Continuity
    May 1, 2018 · Cantor's construction of the real numbers was not sui generis. ... from his [1870] and relied on the Cantorian construction of the real numbers.
  48. [48]
    Hilbert's Program - Stanford Encyclopedia of Philosophy
    Jul 31, 2003 · It calls for a formalization of all of mathematics in axiomatic form, together with a proof that this axiomatization of mathematics is consistent.
  49. [49]
    Real closed field - Encyclopedia of Mathematics
    Jun 6, 2020 · Model theory. In 1940, A. Tarski proved that the elementary theory of all real closed fields is complete; this is known as the Tarski principle.