Fact-checked by Grok 2 weeks ago

Uncountable set

In set theory, an uncountable set is an infinite set for which there exists no bijection with the natural numbers, meaning its elements cannot be enumerated in a sequence that includes every member exactly once. This property distinguishes uncountable sets from countable ones, which are either finite or equinumerous with the naturals via a one-to-one correspondence. The cardinality of an uncountable set is strictly greater than the countable infinity denoted by ℵ₀ (aleph-null). The notion of uncountability emerged from the foundational work of Georg Cantor (1845–1918), a German mathematician who revolutionized the understanding of infinity in the late 19th century. Cantor first demonstrated the uncountability of the real numbers in 1874 using a proof based on nested intervals, and later in 1891 provided the more famous diagonal argument to show that the interval (0,1) cannot be listed exhaustively. His diagonalization method assumes an enumeration of real numbers as infinite decimals and constructs a new number differing in the nth decimal from the nth listed number, proving no such complete list exists. Cantor's insights, initially met with controversy from figures like Leopold Kronecker, established set theory as a distinct branch of mathematics and introduced transfinite cardinalities. Uncountable sets play a central role in measuring the "sizes" of infinities, with the real numbers ℝ serving as the prototypical example, possessing cardinality 2^ℵ₀ known as the continuum c. Cantor's theorem generalizes this by proving that the power set of any set A, denoted 𝒫(A), has strictly greater cardinality than A itself, implying that the power set of the naturals 𝒫(ℕ) is also uncountable and equicardinal to the continuum. Other notable uncountable sets include the Cantor set, a fractal subset of [0,1] with measure zero yet cardinality c, and the set of all functions from ℕ to {0,1}. These concepts underpin broader implications in analysis, topology, and logic, such as the independence of the continuum hypothesis—which posits that no set has cardinality between ℵ₀ and c—from standard axioms like ZFC.

Fundamentals

Definition

In set theory, a set S is defined as uncountable if it is infinite and there exists no bijection between S and the set of natural numbers \mathbb{N}. This means that the elements of S cannot be listed in a sequence that exhausts the set, distinguishing it from sets that can be paired one-to-one with \mathbb{N}. Equivalently, an uncountable set S has cardinality |S| > \aleph_0, where \aleph_0 denotes the cardinality of \mathbb{N}, representing the smallest infinite cardinal. To contextualize uncountability, countability serves as the foundational prerequisite. A set is countable if it is finite or if it possesses cardinality \aleph_0, meaning it is countably infinite (also termed denumerable, indicating a bijection with \mathbb{N}). Thus, uncountable sets are precisely those infinite sets that exceed this countable threshold, marking the onset of larger infinities in the hierarchy of cardinals. The notion of uncountable sets was coined by Georg Cantor in the late 19th century, emerging from his pioneering investigations into the nature of infinite sets and their sizes. In set theory, cardinality |S| quantifies the "size" of a set S, with infinite cardinals indexed by alephs starting from \aleph_0 for countable infinities and proceeding to \aleph_1 and beyond for uncountable ones.

Countability Comparison

A fundamental distinction between countable and uncountable sets lies in their structural behaviors under unions and their cardinal sizes. Countable sets, which include both finite sets and those bijectable with the natural numbers \mathbb{N}, can be combined in limited ways while remaining countable, whereas uncountable sets exhibit strictly larger cardinality that prevents such combinations from encompassing them. A key theorem illustrating this contrast is that the union of countably many countable sets is itself countable. To see this, suppose \{A_n \mid n \in \mathbb{N}\} is a countable collection of countable sets. For each n, enumerate A_n = \{a_{n,k} \mid k \in \mathbb{N}\}. The union \bigcup_{n \in \mathbb{N}} A_n can then be enumerated using a pairing function that bijects \mathbb{N} \times \mathbb{N} onto \mathbb{N}, such as the Cantor pairing function \pi(m,n) = \frac{(m+n)(m+n+1)}{2} + n, which maps each pair (m,n) to a unique natural number. By composing enumerations of each A_n with this pairing, the entire union admits a bijection with \mathbb{N}, hence countable. This theorem has a direct implication for uncountable sets: no uncountable set can be expressed as a countable union of countable sets, as such a representation would render it countable by the above result. Thus, uncountable sets must transcend any such countable layering, underscoring their infinite and "larger" nature. Notably, while countable sets encompass finite sets (with cardinality less than or equal to \aleph_0), uncountable sets are invariably infinite, possessing cardinality strictly greater than \aleph_0. The Schröder-Bernstein theorem reinforces this inequality by establishing that if injections exist in both directions between two sets, then a bijection exists; however, since no injection maps an uncountable set onto a countable one (or vice versa, as countable sets inject into uncountable ones), their cardinalities cannot be equal, confirming the strict separation.

Characterizations

Cardinality Criteria

A set S is uncountable if its cardinality |S| satisfies |S| > \aleph_0, where \aleph_0 denotes the cardinality of the natural numbers \mathbb{N}. Under the axiom of choice, the smallest such uncountable cardinal is \aleph_1, which exceeds all countable cardinals and marks the onset of transfinite infinities beyond the countable realm. Transfinite cardinals are indexed by ordinals via the aleph numbers \aleph_\alpha, where \alpha ranges over all ordinals, providing a hierarchy of infinite sizes. Specifically, \aleph_0 is the smallest infinite cardinal, corresponding to countably infinite sets, while uncountable cardinals commence at \aleph_1 and proceed as \aleph_2, \aleph_3, \dots, \aleph_\omega, \dots, each strictly larger than the previous under the standard ordering of cardinals. This sequence ensures that no uncountable cardinal lies between \aleph_0 and \aleph_1, establishing \aleph_1 as the minimal uncountable size in the cardinal hierarchy. The continuum hypothesis (CH) posits that the cardinality of the real numbers |\mathbb{R}| = 2^{\aleph_0} equals \aleph_1, implying no cardinal exists strictly between \aleph_0 and the continuum. However, Kurt Gödel demonstrated in 1940 that CH is consistent with Zermelo-Fraenkel set theory with the axiom of choice (ZFC), and Paul Cohen proved in 1963 its independence from ZFC, leaving its truth undecided within the standard axioms. Uncountable cardinals arise from well-orderings of sets, where each cardinal \kappa corresponds to an initial ordinal, the smallest ordinal of that cardinality. The countable ordinals culminate at \omega, the order type of \mathbb{N}, beyond which uncountable ordinals begin; the least uncountable ordinal \omega_1 has cardinality \aleph_1 and serves as the prototype for uncountable well-orderings, as every proper initial segment of \omega_1 is countable. Thus, uncountable cardinals are precisely the cardinalities of uncountable initial ordinals, which are limit ordinals exceeding \omega.

Bijection and Injection Tests

One practical criterion for establishing that a set S is uncountable is the absence of a bijection between S and the set of natural numbers \mathbb{N}. Equivalently, S is uncountable if there exists no injection from S into \mathbb{N}, or no surjection from \mathbb{N} onto S. This bijection test directly contrasts countability, where such a one-to-one correspondence exists, and serves as a foundational method to verify uncountability without invoking higher cardinalities. Cantor's theorem provides a generalized diagonalization argument to demonstrate the absence of a surjection from \mathbb{N} onto certain uncountable sets, such as the power set \mathcal{P}(\mathbb{N}). Specifically, assume for contradiction a surjection f: \mathbb{N} \to \mathcal{P}(\mathbb{N}), and define the set D = \{ n \in \mathbb{N} \mid n \notin f(n) \}; then D cannot equal f(k) for any k, as k \in D if and only if k \notin f(k), yielding a contradiction. This diagonalization technique extends to prove no surjection from \mathbb{N} onto \mathcal{P}(S) for any set S, confirming \mathcal{P}(S) is uncountable whenever S is infinite. An injection-based test for uncountability applies when a set S admits an injection into the real numbers \mathbb{R} but no injection into \mathbb{N}. Since \mathbb{R} is uncountable—established via diagonalization or nested intervals—and any subset of a countable set is countable, the lack of an injection into \mathbb{N} implies S cannot be countable. For instance, if an injection g: S \to \mathbb{R} exists and S is infinite, verifying no bijection with \mathbb{N} (e.g., by assuming an enumeration and deriving a contradiction) confirms uncountability. Under the axiom of choice, uncountable sets are Dedekind-infinite, meaning they admit a bijection with a proper subset of themselves, but the key bijection test emphasizes the absence of any bijection with \mathbb{N}. A set is Dedekind-infinite if there exists an injection from \mathbb{N} into it (equivalently, if it contains a countably infinite subset), allowing a proper subset to match its cardinality via shifting. However, the diagnostic focus remains on failing the bijection or injection test with \mathbb{N}, distinguishing uncountable sets from countable infinite ones. For ordered sets, the nested intervals method offers a constructive test for uncountability, particularly in complete ordered fields like \mathbb{R}. Assume an enumeration \{x_n\}_{n=1}^\infty of a bounded interval such as (0,1); construct nested closed intervals I_n = [a_n, b_n] with b_n - a_n = 1/2^n, ensuring x_n \notin I_n at each step, and I_{n+1} \subseteq I_n. By the nested interval property, the intersection \bigcap_{n=1}^\infty I_n contains a point x not in the enumeration, contradicting countability. This bisection-inspired approach generalizes to linearly ordered sets with completeness axioms, providing a mapping-free verification of uncountability through interval avoidance.

Properties

Cardinal Operations

Cardinal arithmetic extends the operations of addition, multiplication, and exponentiation to infinite cardinals, revealing behaviors that differ markedly from finite arithmetic, particularly for uncountable cardinals. For addition, the sum of two infinite cardinals \kappa \geq \aleph_0 and \lambda is given by \kappa + \lambda = \max(\kappa, \lambda). This holds because the disjoint union of sets of these cardinalities can be bijected with the larger set. For uncountable cardinals, such as the smallest uncountable cardinal \aleph_1, adding a countable infinity yields \aleph_1 + \aleph_0 = \aleph_1. Multiplication of infinite cardinals \kappa \geq \aleph_0 and \lambda > 0 satisfies \kappa \cdot \lambda = \max(\kappa, \lambda), as the cardinality of the Cartesian product aligns with the dominant size. Thus, multiplying an uncountable cardinal by any countable or smaller infinite cardinal preserves the uncountable size, for instance, \aleph_1 \cdot \aleph_0 = \aleph_1. Exponentiation is defined as \kappa^\lambda being the cardinality of the set of functions from a set of size \lambda to one of size \kappa. Cantor's theorem states that for any cardinal \kappa, $2^\kappa > \kappa, implying that the power set cardinality strictly exceeds the original. In particular, $2^{\aleph_0}, the cardinality of the continuum, is uncountable and greater than \aleph_0. A key absorption property arises from these operations: the countable union of sets each of uncountable cardinality \kappa has cardinality \kappa, since it equals \aleph_0 \cdot \kappa = \kappa. This underscores how uncountable cardinals dominate countable additions and multiplications without growth.

Subset and Union Behaviors

One key structural property of uncountable sets in Polish spaces is the perfect set property, which asserts that every uncountable Polish space contains a perfect subset—a closed set with no isolated points, hence uncountable itself. This follows from the fact that Polish spaces, being separable complete metric spaces, allow the construction of such nonempty perfect subsets via a Cantor-Bendixson-like decomposition, where the scattered part is countable, leaving an uncountable perfect kernel if the space is uncountable. Consequently, uncountable Polish spaces cannot be scattered and must embed homeomorphically into the Cantor space, highlighting their topological richness despite potential pathologies in subsets. Bernstein sets provide a striking example of uncountable sets with atypical subset behaviors, constructed using the axiom of choice as subsets of the real line that intersect every uncountable closed set but contain no uncountable closed subset themselves. Such sets are uncountable, as they meet every perfect set, yet both the set and its complement avoid containing any perfect subset, rendering them "totally imperfect" in a strong sense. This construction relies on well-ordering the continuum and enumerating perfect sets to ensure transversal intersections, demonstrating how the axiom of choice enables sets that evade standard closure properties. Vitali sets exemplify non-measurable uncountable subsets of the reals, constructed by selecting one representative from each equivalence class under rational translations within [0,1), yielding a set that is uncountable but lacks Lebesgue measurability. The non-measurability arises because the countable disjoint union of its rational translates covers [0,1) up to measure zero, yet assigning a measure to the Vitali set leads to a contradiction with countable additivity and translation invariance of Lebesgue measure. Thus, not every uncountable subset of \mathbb{R} is Lebesgue measurable, underscoring limitations in measure theory for arbitrary subsets. Regarding decompositions, uncountable sets in separable metric spaces can possess countable dense subsets while remaining uncountable overall, as seen in the reals where the rationals form a countable dense subset. This separability ensures that the topology is generated by countably many basic open sets, allowing a countable dense subset to approximate the entire space, yet the uncountable cardinality persists due to the presence of continuum-many limit points. Such decompositions illustrate how uncountable sets can be "sparsely" populated topologically without sacrificing their size.

Examples

Real Numbers

The set of real numbers \mathbb{R} serves as the canonical example of an uncountable set, illustrating a cardinality strictly larger than that of the natural numbers \mathbb{N}. To demonstrate this, Georg Cantor employed his diagonal argument, which shows that no surjective function exists from \mathbb{N} to the open interval (0,1), and hence to \mathbb{R}. Assume, for contradiction, that there is an enumeration \{r_n\}_{n=1}^\infty of all real numbers in (0,1), where each r_n has a decimal expansion r_n = 0.d_{n1}d_{n2}d_{n3}\dots with digits d_{ni} \in \{0,1,\dots,9\}. Construct a new real number x = 0.x_1 x_2 x_3 \dots \in (0,1) by defining x_i = 4 if d_{ii} \neq 4 and x_i = 5 if d_{ii} = 4 for each i. This x differs from every r_n in the nth decimal place, so x is not in the enumeration, contradicting the assumption of surjectivity. The cardinality of \mathbb{R}, denoted |\mathbb{R}| = \mathfrak{c} (the continuum) or $2^{\aleph_0}, is uncountable and equals the cardinality of the power set of \mathbb{N}. This follows from the existence of an injection from (0,1) to the set of all subsets of \mathbb{N} via binary expansions (with care for non-unique representations like $0.0111\dots = 0.1000\dots), and Cantor's theorem that |\mathcal{P}(S)| > |S| for any set S. Under the continuum hypothesis (CH), formulated by Cantor in 1878, \mathfrak{c} = \aleph_1, the smallest uncountable cardinal, though CH is independent of standard set theory axioms. The sets \mathbb{R} and (0,1) have the same cardinality, as shown by the bijection f: (0,1) \to \mathbb{R} given by f(x) = \tan(\pi(x - 1/2)), which is continuous, strictly increasing, and maps the endpoints asymptotically to \pm \infty. Despite this equivalence, \mathbb{Q} is countable and dense in \mathbb{R}, meaning that between any two distinct reals a < b, there exists q \in \mathbb{Q} with a < q < b; this follows from the Archimedean property, as one can find integers m,n such that a < m/n < b. Cantor's enumeration proves |\mathbb{Q}| = \aleph_0 by listing positives as p/q in lowest terms via a diagonal traversal of the grid, extending to all rationals. Thus, the uncountability of \mathbb{R} arises from the irrationals, which are dense as well.

Power Sets and Generalizations

One fundamental example of an uncountable set arises from the power set operation applied to the natural numbers. The power set 𝒫(ℕ) is the set of all subsets of ℕ, and by Cantor's theorem, its cardinality satisfies |𝒫(ℕ)| > |ℕ| = ℵ₀, establishing that 𝒫(ℕ) is uncountable. More precisely, the cardinality of 𝒫(ℕ) is 2^{ℵ₀}, which equals the cardinality of the continuum 𝔠. Cantor's theorem generalizes beyond ℕ: for any set S, the cardinality of its power set satisfies |𝒫(S)| > |S|, implying that the power set of any infinite set is uncountable. In particular, since ℕ is infinite and countable, 𝒫(ℕ) provides a concrete uncountable set of cardinality 2^{ℵ₀}. This construction highlights how power sets generate larger infinities iteratively, producing an infinite hierarchy of uncountable cardinals. Applying the power set to larger sets yields even greater uncountable cardinalities. For instance, the power set 𝒫(ℝ) of the real numbers has cardinality 2^𝔠, which exceeds 𝔠 and thus represents a strictly larger uncountable infinity. Uncountable sets also emerge in function spaces. The set of all functions from ℕ to ℕ, denoted ^ℕℕ, is uncountable with cardinality 2^{ℵ₀}. This follows from an injection into 𝒫(ℕ × ℕ) via graphs of functions (subsets where each first coordinate appears exactly once), since |ℕ × ℕ| = ℵ₀ and |𝒫(ℕ × ℕ)| = 2^{ℵ₀}, implying |^ℕℕ| ≤ 2^{ℵ₀}. For the reverse inequality, the set of functions from ℕ to {0,1} (subsets of ℕ, cardinality 2^{ℵ₀}) injects into ^ℕℕ by identifying {0,1} with {0,1} ⊆ ℕ, so |^ℕℕ| = 2^{ℵ₀}. In cardinal arithmetic, such exponentiations generalize the power set construction to mappings between sets of arbitrary sizes.

Without Axiom of Choice

Dependent Properties

The axiom of choice (AC) implies the well-ordering theorem, which states that every set can be well-ordered, including uncountable sets. A well-ordering on a set is a total order such that every nonempty subset has a least element, allowing the assignment of an ordinal number to represent the order type of the set. For uncountable sets like the real numbers, this yields an ordinal of cardinality \mathfrak{c}, the continuum, though the explicit well-ordering remains non-constructive without additional assumptions. This property enables the extension of transfinite induction and recursion to uncountable domains, facilitating proofs in set theory and analysis. AC also ensures the comparability of cardinals, meaning that for any two cardinal numbers \kappa and \lambda, either \kappa \leq \lambda or \lambda \leq \kappa. Without AC, there exist models where infinite cardinals are incomparable, but under AC, every uncountable cardinal can be compared to others, such as \aleph_1 \leq 2^{\aleph_0} or the negation thereof via the continuum hypothesis. This comparability underpins the structure of the cardinal hierarchy and is equivalent to AC via Zorn's lemma applied to chains of cardinals. The Vitali construction provides an example of an uncountable non-Lebesgue measurable set in \mathbb{R}, relying on AC to select representatives. Consider the equivalence relation on \mathbb{R} where x \sim y if x - y \in \mathbb{Q}; the equivalence classes form a partition of \mathbb{R} into uncountably many disjoint sets, each dense in \mathbb{R}. AC allows choosing one representative from each class to form the Vitali set V \subseteq [0,1], which is uncountable and cannot be assigned a Lebesgue measure, as the countable disjoint translates V + q for q \in \mathbb{Q} \cap [-1,1] cover [0,2] while overlapping minimally. This demonstrates the existence of pathological uncountable subsets incompatible with intuitive measure properties. The Banach-Tarski paradox illustrates a counterintuitive decomposition of uncountable sets using AC. Specifically, the unit ball in \mathbb{R}^3 can be partitioned into finitely many (as few as five) non-measurable pieces that can be rigidly reassembled via isometries into two copies of the unit ball. The construction exploits the free group structure within the rotation group SO(3) and applies AC to select representatives from orbits under group actions, yielding sets without well-defined volume. This uncountable phenomenon highlights how AC enables paradoxical behaviors in geometric measure theory, where volume preservation fails for such decompositions.

Models and Counterexamples

In set theory without the axiom of choice (AC), various models demonstrate the independence of statements involving uncountable sets, highlighting how the failure of AC can lead to counterintuitive cardinal behaviors. One prominent example is provided by Cohen forcing, a technique developed by Paul Cohen to construct models of ZF where the continuum hypothesis (CH) fails. Specifically, Cohen showed that it is consistent with ZF that the cardinality of the continuum, $2^{\aleph_0}, equals \aleph_2, meaning the power set of the natural numbers has cardinality strictly larger than \aleph_1 but not arbitrarily large. This model illustrates that uncountable cardinals like the continuum can take on values beyond the immediate successor in the aleph hierarchy without violating ZF axioms. Fraenkel-Mostowski (FM) permutation models offer another class of counterexamples, constructed using a set of atoms and group actions to interpret the universe in a way that violates AC while preserving ZF minus AC (ZFA). In the basic Fraenkel model, the set of atoms A is infinite, but every subset of A that is "symmetric" (invariant under finite-support permutations) lacks a countable infinite subset. This implies that A is not countable, as there is no injection from \omega into A, yet proofs of uncountability relying on AC—such as well-ordering principles—fail, rendering A a paradoxical infinite set whose uncountability cannot be established via choice-dependent methods. Without AC, Dedekind-finite infinite sets provide further counterexamples to standard notions of infinity tied to uncountability. A set is Dedekind-finite if it is not equinumerous to any proper subset of itself, meaning no injection from \omega exists into the set. In ZF alone, it is consistent that infinite Dedekind-finite sets exist, such as amorphous sets where every subset is finite or cofinite, but the set itself admits no countably infinite subset. These sets are uncountable in the sense of lacking a bijection with \omega, yet they defy the ZFC equivalence between Dedekind-finiteness and actual finiteness, complicating the characterization of uncountable structures. The Russell socks analogy underscores the failure of AC in handling families of pairs, particularly when extended to uncountable collections. Bertrand Russell illustrated that while one can select the left shoe from each pair in an infinite collection of distinguishable shoe pairs without AC (via an explicit rule), selecting one sock from each pair in an infinite collection of identical pairs requires arbitrary choice, as no distinguishing feature exists. In models without AC, this extends to uncountable families of two-element sets, where no choice function can be defined without additional assumptions, leading to counterexamples where uncountable unions or products cannot be well-defined in the expected way.

References

  1. [1]
  2. [2]
    4.10 Cantor's Theorem
    Cantor's theorem implies that there are infinitely many infinite cardinal numbers, and that there is no largest cardinal number.
  3. [3]
    Cantor, Georg (1845-1918) - CMU School of Computer Science
    Georg Cantor astonished the mathematical world with his discovery that the integers and the real numbers cannot be placed into one-to-one onto correspondance.
  4. [4]
    [PDF] Cantor's Other Proofs that R Is Uncountable
    Cantor's first proof of the uncountability of R was published in 1874 and is based on the fact that bounded monotonic sequences of real numbers converge.
  5. [5]
    Set Theory - Stanford Encyclopedia of Philosophy
    Oct 8, 2014 · Set theory is the mathematical theory of well-determined collections, called sets, of objects that are called members, or elements, of the set.
  6. [6]
    9.2: Countable Sets - Mathematics LibreTexts
    Sep 29, 2021 · A set that is countably infinite is sometimes called a denumerable set. A set is countable provided that it is finite or countably infinite.Infinite Sets · Countably Infinite Sets · The Set of Positive Rational...
  7. [7]
    [PDF] Naive set theory. - Whitman People
    union of a countably infinite family of countable sets is countable. Proof: given the family {Xn\ (n e w) of countable sets, find a family {/„} of func.
  8. [8]
    Uncountably Infinite -- from Wolfram MathWorld
    An infinite set, such as the real numbers, which is not countably infinite.
  9. [9]
    Aleph-1 -- from Wolfram MathWorld
    Aleph-1 is the set theory symbol aleph_1 for the smallest infinite set larger than aleph_0 (Aleph-0), which in turn is equal to the cardinal number of the set
  10. [10]
    Aleph -- from Wolfram MathWorld
    The set theory symbol ( aleph ) for the cardinal number of a well-orderable infinite set. See also Aleph-0, Aleph-1, Countable Set, Countably Infinite, Finite, ...
  11. [11]
    Continuum Hypothesis -- from Wolfram MathWorld
    The proposal originally made by Georg Cantor that there is no infinite set with a cardinal number between that of the small infinite set of integers aleph_0.
  12. [12]
    The Continuum Hypothesis - Stanford Encyclopedia of Philosophy
    May 22, 2013 · The continuum hypothesis (under one formulation) is simply the statement that there is no such set of real numbers. It was through his attempt ...
  13. [13]
    [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.<|control11|><|separator|>
  14. [14]
    [PDF] math 112, spring 2019
    Mar 5, 2019 · A set A is uncountable if it is infinite and not countable. When sets are in bijection, we think of them as having the same number of elements.<|control11|><|separator|>
  15. [15]
    [PDF] Cantor's Legacy - andrew.cmu.ed
    If S is finite, the power set of S has cardinality 2|S|. Cantor's Theorem: The power set of a set has a greater cardinal number than the set itself,. |S| < |P(S)|.
  16. [16]
    [PDF] Lecture Notes 2023 - Analysis I: One Variable - ETH Zürich
    Dec 8, 2023 · • A set is called countable if there is a bijection to N. The ... that R is uncountable, it suffices to prove the existence of an injection.
  17. [17]
    Real Analysis: Theorem 2.1.4: Dedekind Theorem - MathCS.org
    Countable Infinity. Theorem 2.1.4: Dedekind Theorem. A set S is infinite if and only if there exists a proper subset A of S which has the same cardinality as S.
  18. [18]
    [PDF] A Dedekind Finite Borel Set - University of Wisconsin–Madison
    An infinite D ⊆ 2ω cannot be amorphous. We don't know if there could be an uncountable Borel set D ⊆ 2ω such that every subset is countable or co-countable ...
  19. [19]
    [PDF] An Introduction to Real Analysis - SUNY Geneseo
    May 4, 2022 · (i) The set S is countably infinite if there is a bijection from. N ... . Using the Nested Interval property of R, we give a proof that R is.
  20. [20]
    Cardinal Addition -- from Wolfram MathWorld
    Cardinal Addition. Cite this as: Weisstein, Eric W. "Cardinal Addition." From MathWorld--A Wolfram Resource. https://mathworld.wolfram.com/CardinalAddition.html ...Missing: arithmetic | Show results with:arithmetic
  21. [21]
    Cardinal Multiplication -- from Wolfram MathWorld
    Let A and B be any sets. Then the product of |A| and |B| is defined as the Cartesian product |A|*|B|=|A×B|.
  22. [22]
    Cardinal Exponentiation -- from Wolfram MathWorld
    Let A and B be any sets, and let |X| be the cardinal number of a set X. Then cardinal exponentiation is defined by |A|^(|B|)=|set of all functions from B into ...
  23. [23]
    [PDF] Descriptive Set Theory
    Theorem 2.25 (Perfect Set Theorem for Borel Sets) If X is a Polish space and B ⊆ X is an uncountable Borel set, then B contains a perfect set. Proof Let τ ...Missing: source | Show results with:source
  24. [24]
    [PDF] Some descriptive set theory
    Aug 13, 2008 · Let X be a Polish space and A ⊆ X be analytic and uncountable. A contains a perfect set and therefore |A| = 2ℵ0 . Proof. Because A is ...
  25. [25]
    SOME REMARKS ON DIFFERENCE SETS OF BERNSTEIN SETS
    this classification. Definition 1 A set C is called a Bernstein set if both C and C have nonempty intersection with every perfect set. (A set is perfect ...
  26. [26]
    [PDF] 4. Countability
    Definition 6.1. A set A that is not countable is called uncountable. By this point, the following should not be surprising. Proposition 6.2. If A is uncountable ...
  27. [27]
    [PDF] Density of the Rationals - UC Davis Math
    The density of rationals means there is a rational number strictly between any two distinct real numbers, however close together they may be.
  28. [28]
    [PDF] Infinity and its cardinalities
    In fact, Cantor would prove that, in general, this is not true. He showed that some infinite sets have a greater cardinality than others, thus implying the.
  29. [29]
    Peter Suber, "Infinite Sets"
    If S is a set of any infinite cardinality, then its power set has a greater infinite cardinality, or |*S| > |S|. This follows directly from Cantor's Theorem ( ...
  30. [30]
    [PDF] Theorem 1. The set F of all functions from N to N is uncountable.
    It is impossible for g to simultaneously be a member of F and not be, and so we conclude that the set F of all functions from N to N is in fact not countable.
  31. [31]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · There is a Lebesgue nonmeasurable set of real numbers (Vitali 1905). This was shown much later to be a consequence of BPI (see below) and ...
  32. [32]
    The Axiom of Choice > Notes (Stanford Encyclopedia of Philosophy)
    A millionaire possesses an infinite number of pairs of shoes, and an infinite number of pairs of socks. One day, in a fit of eccentricity, the millionaire ...