Fact-checked by Grok 2 weeks ago

Cardinal number

In , a cardinal number is a type of number that represents the size or of a set, indicating the number of distinct elements it contains without reference to or arrangement. For finite sets, cardinal numbers correspond to the natural numbers, such as 0 for the , 1 for a , and so forth up to any positive . This concept extends to infinite sets through transfinite cardinals, pioneered by , where sets like the natural numbers have denoted by the symbol ℵ₀ (aleph-null), representing the smallest infinite . Cardinal numbers differ fundamentally from ordinal numbers, which describe the position or order of elements in a well-ordered set rather than mere quantity. For instance, while the ordinal number ω denotes the order type of the natural numbers in their standard ordering, the cardinal ℵ₀ measures only their size, allowing bijections to establish equivalence between sets of equal cardinality regardless of structure. Two sets have the same cardinal number if there exists a bijection between them, a criterion that unifies finite and infinite cases under set theory. The development of cardinal numbers arose in the late through Cantor's work on , where he demonstrated that infinite sets can have different sizes, challenging earlier intuitions about . Cantor's proved that the set of real numbers has a larger than the natural numbers, leading to the , which posits that there is no strictly between ℵ₀ and the (2^ℵ₀). Cardinal arithmetic, including , , and exponentiation defined via disjoint unions and Cartesian products, further reveals counterintuitive properties of infinite operations, such as ℵ₀ + 1 = ℵ₀ in terms of . These concepts underpin modern mathematics, from to , emphasizing the abstract measurement of "how many" beyond finite bounds.

Historical Development

Early Concepts

The ancient laid the groundwork for understanding numbers as finite multitudes, emphasizing and without delving into quantities. In Euclid's Elements (c. 300 BCE), Book VII defines a number as "a multitude composed of units," framing around the successive addition of discrete units to form finite collections. This approach focused on practical and proportion, treating numbers as concrete aggregates rather than abstract sizes applicable to infinities. Greek philosophers, particularly , further distinguished potential infinity—an unending process, such as dividing a line indefinitely—from , which they rejected as incoherent to avoid logical paradoxes. Medieval scholars integrated these ideas into philosophical and theological frameworks, often viewing infinity as a divine attribute beyond human mathematics, with figures like Thomas Aquinas asserting that only God embodies true infinity while earthly numbers remained finite or potentially infinite. By the Renaissance, empirical observations challenged these boundaries; Galileo Galilei, in his Dialogues Concerning Two New Sciences (1638), articulated a striking paradox by observing that the perfect squares (1, 4, 9, ...) form an infinite set that can be paired one-to-one with all natural numbers (1, 2, 3, ...), suggesting the squares are simultaneously fewer yet equally numerous. This "paradox of squares" underscored the intuitive failure of finite comparison methods when applied to infinities, prompting early reflections on the peculiar "sizes" of endless collections. In the , these intuitive notions evolved toward mathematical precision through the works of and . Dedekind, in his 1872 essay Continuity and Irrational Numbers, introduced "chains" (Ketten)—ordered sequences of rational numbers—to construct the real numbers, revealing the infinite structure underlying continuous magnitudes and implicitly highlighting the distinction between countable and uncountable infinities. , building on this, initiated rigorous study of infinite sets in the 1870s; his 1873 letter to Dedekind and 1874 paper "On a Property of the Collection of All Real Algebraic Numbers" proved that the real numbers exceed the natural numbers in by showing no one-to-one correspondence exists between them, using nested intervals to demonstrate uncountability. This marked Cantor's realization that infinities come in varying "sizes," with the vastly larger than the countable infinity of naturals. Later, in 1891, Cantor formalized the diagonal argument to prove the uncountability of real numbers in (0,1), solidifying these ideas. Facing criticism from contemporaries like , Cantor's theories found a staunch defender in , who in correspondence during the 1890s praised the transfinite numbers and later, in a 1925 address, declared Cantor's realm a "paradise" from which mathematicians would not be expelled.

Modern Foundations

In response to paradoxes such as that plagued , provided the first axiomatic foundation in 1908 with his system, which formalized the concepts of ordinals and cardinals through well-orderings to handle infinite sets rigorously. Zermelo's axioms, including , , , , , , separation, and (via well-ordering), enabled the systematic definition of cardinals as the smallest ordinals equinumerous to a given set, building directly on Cantor's transfinite numbers while avoiding inconsistencies. This framework shifted the study of cardinal numbers from intuitive constructions to a deductively secure structure, establishing as the bedrock for modern mathematics. Abraham Fraenkel refined Zermelo's system in 1922 by introducing the and strengthening separation to its modern form, with independently proposing similar changes, culminating in Zermelo-Fraenkel set theory (ZF). When combined with the , this yields ZFC, the predominant today, which fully supports the arithmetic and comparison of cardinals without foundational paradoxes. Zermelo himself contributed further in the 1930s by exploring natural well-orderings, reinforcing the ordinal-cardinal distinction as central to transfinite enumeration. Cantor's transfinite numbers were integrated into the structuralist program of the group during the 1930s and 1950s, where served as the unifying foundation for all in their . Bourbaki's 1957 Théorie des ensembles axiomatized cardinals within ZFC-like principles, emphasizing their role in measuring set sizes abstractly and structurally, thus embedding Cantor's ideas into a comprehensive, mother-theory approach that influenced mid-20th-century mathematical . Key advancements in understanding cardinal hierarchies came in when introduced the constructible universe L, proving the relative consistency of ZFC with the and the generalized (GCH), where cardinals satisfy a precise scale of powers without additional assumptions. This inner model demonstrated that CH holds in L, providing a canonical context for cardinal exponentiation. Complementing this, Paul Cohen's invention of forcing showed that the negation of CH is also consistent with ZFC, establishing the independence of continuum-related cardinal questions and opening the door to model-theoretic explorations of infinite cardinals.

Basic Concepts

Finite Cardinals

In set theory, the cardinality of a finite set is defined as the number of distinct elements it contains, which is a natural number. For instance, the set {1, 2, 3} has cardinality 3, denoted |{1, 2, 3}| = 3. This measure captures the "size" of the set through direct counting, aligning finite cardinals directly with the non-negative integers in the natural numbers ℕ (including 0 for the empty set). Every finite cardinal κ arises from a finite set that admits a bijection with an initial segment of the natural numbers, such that κ = n for some n ∈ ℕ, where n = {0, 1, \dots, n-1}. This bijection ensures a one-to-one correspondence between the elements of the set and the first n natural numbers, uniquely determining the cardinal as that natural number. Consequently, finite cardinals are isomorphic to the natural numbers, providing a foundational link between intuitive counting and abstract set-theoretic size. Finite cardinals possess key structural properties rooted in their identification with natural numbers. They form a well-ordered set under the standard ordering of ℕ, meaning every non-empty has a least , which follows from the well-ordering of the naturals themselves. The successor operation on finite cardinals mirrors that on natural numbers: for a κ = n, the successor κ + 1 = n + 1 corresponds to adding one to a set of size n. Additionally, finite cardinals exhibit no infinite descending chains under subtraction or ordering, reflecting the well-founded nature of ℕ where every decreasing sequence terminates. A hallmark of finite sets—and thus their cardinals—is Dedekind-finiteness: a set is Dedekind-finite if it is not equinumerous (bijectable) with any of its proper subsets. For example, no finite set can be paired one-to-one with a subset missing an element, as this would violate the bijection to its natural number counterpart. This property sharply distinguishes finite cardinals from infinite ones, where such bijections are possible.

Infinite Cardinals

Infinite cardinal numbers are those cardinalities that cannot be placed in bijection with any natural number, distinguishing them from finite cardinals which correspond to the sizes of finite sets. In set theory, an infinite cardinal κ satisfies κ ≥ ℵ₀, where ℵ₀ denotes the cardinality of the natural numbers, representing the smallest infinite size. These cardinals arise as the sizes of infinite sets and form a hierarchy under the standard ordering of cardinals. The hierarchy of infinite cardinals begins with the countable infinite cardinal ℵ₀ and proceeds to larger uncountable cardinals, such as the c = 2^{ℵ₀}. For distinct infinite cardinals κ and λ with κ ≤ λ, the strict inequality κ < λ holds if there is no injection from λ into a set of cardinality κ, ensuring a total order on the class of all cardinals. This ordering reflects the impossibility of injecting a larger infinite set into a smaller one without surjectivity in the reverse direction. Infinite cardinals possess unique properties not shared with finite ones; notably, every infinite cardinal is a limit ordinal in the sense that it is an initial ordinal that is the supremum of smaller ordinals. A key absorption law states that for any infinite cardinal κ, κ + ℵ₀ = max(κ, ℵ₀), meaning the addition of countably many elements does not increase the cardinality beyond the maximum of the two. Similar absorption occurs in multiplication under certain conditions, emphasizing how infinite sizes "absorb" smaller infinities. Representative examples illustrate these concepts: the set of rational numbers ℚ has cardinality ℵ₀, as it is a countable union of countable sets. In contrast, the set of real numbers ℝ has cardinality c = 2^{ℵ₀}, which is uncountable and strictly larger than ℵ₀ by Cantor's diagonal argument./08%3A_Cardinality/8.03%3A_Cantors_Theorem) The power set of the natural numbers, 𝒫(ℕ), also has cardinality 2^{ℵ₀}, exemplifying how exponentiation generates larger infinite cardinals.

Formal Definition

Definition via Equinumerosity

In set theory, the cardinality of a set is defined through the relation of equinumerosity, which captures the intuitive notion of two sets having the same "size" regardless of their elements' nature. Two sets A and B are equinumerous, denoted A \sim B, if there exists a bijection f: A \to B, meaning a function that is both injective and surjective, establishing a one-to-one correspondence between every element of A and every element of B. This relation, introduced by , forms an equivalence relation on the class of all sets, partitioning them into equivalence classes where sets within the same class share identical cardinality. Cardinal numbers are precisely these equivalence classes under equinumerosity: the cardinal number |A| of a set A is the equivalence class of all sets equinumerous to A, so |A| = |B| if and only if A \sim B. To avoid paradoxes arising from treating these classes as sets, cardinal numbers are typically represented by specific canonical sets, most commonly initial ordinals in the von Neumann construction. An initial ordinal is an ordinal \alpha such that no smaller ordinal is equinumerous to it. In the von Neumann definition, a cardinal number \kappa is identified with the smallest ordinal equinumerous to any set of that cardinality; thus, \kappa itself is a transitive set well-ordered by membership, and every set of cardinality \kappa admits a bijection to \kappa. This representation ensures that finite cardinals coincide with the natural numbers (e.g., the cardinal 3 is the ordinal \{0, 1, 2\}), while infinite cardinals like \aleph_0 (the cardinality of the natural numbers) are the least infinite ordinals with no equinumerous predecessor. To compare cardinalities, for sets A and B, one defines |A| \leq |B| if there exists an injection from A to B, allowing A to be embedded into B without overlap. This partial order on cardinals aligns with equinumerosity for equality and extends naturally to strict inequality when |A| \leq |B| but not |A| \sim |B|, as established in Cantor's foundational work on set sizes.

Axiom of Choice Role

In the absence of the axiom of choice (AC), cardinal numbers do not necessarily satisfy the trichotomy law, meaning there can exist sets A and B such that neither |A| \leq |B| nor |B| \leq |A|, rendering some cardinals incomparable. This incomparability arises because, without AC, the existence of injections between sets cannot always be guaranteed even when one might intuitively expect comparability. The axiom of choice resolves this by implying the well-ordering theorem, which states that every set can be well-ordered. Under a well-ordering, the cardinality of a set is identified with the smallest ordinal equinumerous to it, known as its initial ordinal, providing a uniform way to assign and compare cardinals across all sets. This equivalence between AC and the well-ordering theorem ensures that cardinals form a totally ordered class. AC is equivalent to Zorn's lemma, which posits that every partially ordered set with upper bounds for all chains contains a maximal element. In the context of cardinals, Zorn's lemma facilitates the construction of maximal chains in the poset of well-orderings on a set, thereby proving the existence of a well-ordering and enabling cardinal assignment. A concrete example is the set of real numbers, whose cardinality |\mathbb{R}| = 2^{\aleph_0} admits a well-ordering precisely due to AC, allowing it to be bijected with some initial ordinal. Conversely, without AC, models of ZF set theory can contain infinite Dedekind-finite sets—sets that are infinite but have no countable infinite subset—further illustrating how AC underpins the standard properties of infinite cardinals.

Cardinal Arithmetic

Addition and Multiplication

The addition of two cardinal numbers \kappa and \lambda, denoted \kappa + \lambda, is defined as the cardinality of the disjoint union of two sets A and B such that |A| = \kappa and |B| = \lambda. Formally, \kappa + \lambda = |A \sqcup B|, where the disjoint union ensures A \cap B = \emptyset. This definition is independent of the specific choice of sets A and B, as any two sets of the same cardinality are equinumerous via a bijection, preserving the resulting cardinality. For finite cardinal numbers, addition coincides with the standard addition of natural numbers; for instance, $3 + 5 = 8. This follows directly from the bijection between the disjoint union and the finite set of that size in the natural numbers. In contrast, for infinite cardinals, the behavior differs markedly: if \kappa \leq \lambda and \lambda is infinite, then \kappa + \lambda = \lambda, or more generally, \kappa + \lambda = \max(\kappa, \lambda). A representative example is \aleph_0 + \aleph_0 = \aleph_0, as the disjoint union of two countably infinite sets, such as the natural numbers and the integers offset by a large constant, remains countably infinite via a zig-zag enumeration. To establish \kappa + \lambda = \max(\kappa, \lambda) for infinite \lambda \geq \kappa, first note there exists an injection from A into B by assumption of \kappa \leq \lambda. The disjoint union A \sqcup B then injects into B \sqcup B, which has cardinality \lambda + \lambda. For infinite \lambda, \lambda + \lambda = \lambda holds by inducting on the well-ordering of \lambda or by explicit bijection, such as pairing even and odd indices in a well-ordered set. An injection from B into A \sqcup B is immediate by mapping into the B component. The then guarantees a bijection, yielding |A \sqcup B| = \lambda. The asserts that if there are injections f: X \to Y and g: Y \to X, then a bijection exists between X and Y. The multiplication of cardinals \kappa \cdot \lambda is defined as the cardinality of the Cartesian product A \times B, where |A| = \kappa and |B| = \lambda. Thus, \kappa \cdot \lambda = |A \times B|. As with addition, this is well-defined independent of representatives. For finite cardinals, multiplication reduces to the usual operation on natural numbers, such as $3 \cdot 5 = 15, corresponding to the size of a $3-by-&#36;5 grid. For infinite cardinals, assuming the axiom of choice, if both \kappa and \lambda are infinite, then \kappa \cdot \lambda = \max(\kappa, \lambda). Without loss of generality, assume \kappa \leq \lambda; then \lambda \leq \kappa \cdot \lambda \leq \lambda \cdot \lambda = |\lambda \times \lambda|. Under the axiom of choice, every set admits a well-ordering, allowing a bijection between \lambda \times \lambda and \lambda for infinite \lambda, as elements can be enumerated by the minimum rank in the lexicographic order on pairs. The again applies to equate the sizes. An example is $2^{\aleph_0} \cdot \aleph_0 = 2^{\aleph_0}, since the Cartesian product of the reals and the naturals admits a bijection with the reals, such as interleaving decimal expansions. If one factor is finite and positive and the other infinite, the product equals the infinite cardinal, as it decomposes into finitely many disjoint copies.

Exponentiation

In cardinal arithmetic, exponentiation is defined for cardinals \kappa and \lambda as \kappa^\lambda = |B^A|, where A and B are sets with |A| = \lambda and |B| = \kappa, and B^A denotes the set of all functions from A to B. This definition is independent of the particular choice of sets A and B, as any two sets of the same cardinality admit bijections that preserve the structure of the function set. For finite cardinals, this operation coincides with the standard numerical exponentiation, such as $2^3 = 8. For infinite cardinals, exponentiation exhibits behaviors distinct from finite cases, including non-commutativity; for instance, \aleph_0^{\aleph_0} = 2^{\aleph_0} while \aleph_0 \cdot 2 = \aleph_0 < 2^{\aleph_0}. A fundamental result is Cantor's theorem, which asserts that for any cardinal \kappa, $2^\kappa > \kappa./08%3A_Cardinality/8.03%3A_Cantors_Theorem) The proof relies on : assuming a surjection from a set of cardinality \kappa to its leads to a contradiction by constructing an element not in the image via a diagonal argument over the characteristic functions./08%3A_Cardinality/8.03%3A_Cantors_Theorem) This strict inequality establishes that the power set operation strictly increases , generating an unending hierarchy of infinite cardinals. Key bounds for infinite cardinals \kappa \geq 2 and \lambda provide insight into the scale of \kappa^\lambda: \max(\kappa, 2^\lambda) \leq \kappa^\lambda \leq 2^{\max(\kappa, \lambda)}. The lower bound follows from injecting the larger of a set of size \kappa (via constant functions) or the power set of \lambda (via characteristic functions into \{0,1\}^\lambda \subseteq \kappa^\lambda). The upper bound arises because the set of functions is at most the size of the power set of \kappa \times \lambda, yielding |\kappa^\lambda| \leq 2^{\kappa \cdot \lambda} = 2^{\max(\kappa, \lambda)} for cardinals. An illustrative equality is \aleph_0^{\aleph_0} = 2^{\aleph_0}, obtained by bounding $2^{\aleph_0} \leq \aleph_0^{\aleph_0} \leq (2^{\aleph_0})^{\aleph_0} = 2^{\aleph_0 \cdot \aleph_0} = 2^{\aleph_0}. A significant theorem concerning successor cardinals is Hausdorff's formula: for an infinite cardinal \kappa and cardinal \lambda with \cf(\kappa^+) > \lambda, (\kappa^+)^\lambda = \kappa^+ \cdot 2^\lambda, noting that this simplifies further under the generalized (GCH) where $2^\lambda = \lambda^+. The proof involves showing that the functions from a set of size \lambda to \kappa^+ can be bijected with a set of size \kappa^+ \cdot 2^\lambda by partitioning based on the supremum of the image and encoding the rest via subsets. This formula highlights how of successor cardinals interacts with the function, providing explicit computations in specific regimes.

Comparison and Ordering

Under the axiom of choice, the class of cardinal numbers is totally ordered by the relation of cardinality, meaning that for any two cardinals \kappa and \lambda, either \kappa \leq \lambda or \lambda \leq \kappa. Here, \kappa \leq \lambda if there exists an injection from a set of cardinality \kappa to a set of cardinality \lambda, and \kappa = \lambda if there additionally exists a bijection between them. The strict inequality \kappa < \lambda holds if \kappa \leq \lambda but \kappa \neq \lambda, which for infinite cardinals is equivalent to the existence of an injection from a set of cardinality \kappa to a set of cardinality \lambda with no bijection, or equivalently, \kappa + 1 \leq \lambda. A fundamental result enabling precise comparisons is the Schröder–Bernstein theorem, which states that if \kappa \leq \lambda and \lambda \leq \kappa, then \kappa = \lambda. The proof relies on a back-and-forth construction: given injections f: \kappa \to \lambda and g: \lambda \to \kappa, one partitions the sets into chains based on the images under iterated applications of f and g^{-1}, then defines a bijection by matching within chains and using the injections to cover the rest. This theorem ensures that the ordering is antisymmetric and, combined with the totality from the , establishes a linear order on cardinals. Another key aspect of ordering involves cofinality, which measures the "singularity" of a cardinal. The cofinality \mathrm{cf}(\kappa) of an infinite cardinal \kappa is the smallest cardinal \gamma such that \kappa can be expressed as the union of \gamma many sets, each of cardinality strictly less than \kappa. Equivalently, \mathrm{cf}(\kappa) is the smallest ordinal \gamma admitting a cofinal map into the ordinal \kappa, where cofinal means the supremum of the image is \kappa. A cardinal \kappa is regular if \mathrm{cf}(\kappa) = \kappa and singular if \mathrm{cf}(\kappa) < \kappa. For example, the countable infinite cardinal \aleph_0 is regular since \mathrm{cf}(\aleph_0) = \aleph_0, as any countable union of finite sets is countable, but no smaller cofinal subset suffices. In contrast, \aleph_\omega, the least upper bound of \aleph_n for finite n, is singular with \mathrm{cf}(\aleph_\omega) = \aleph_0, as it is the union of countably many smaller cardinals \aleph_0, \aleph_1, \dots. These notions of regularity and singularity influence how cardinals behave under arithmetic operations and limits in the ordering.

Advanced Topics

Aleph Numbers

The aleph numbers provide a systematic enumeration of the infinite cardinal numbers, indexed by ordinals in a well-ordered sequence. Introduced by , the notation \aleph_\alpha denotes the \alpha-th infinite cardinal, where \alpha is an ordinal number. Specifically, \aleph_0 is defined as the cardinality of the set of natural numbers \mathbb{N}, representing the smallest infinite cardinal. Successor aleph numbers are constructed recursively: \aleph_{\alpha+1} is the smallest cardinal strictly greater than \aleph_\alpha, equivalent to the cardinality of the set of all ordinals of cardinality at most \aleph_\alpha. For limit ordinals \lambda, the aleph number \aleph_\lambda is the supremum of the preceding alephs in the sequence: \aleph_\lambda = \sup\{\aleph_\alpha \mid \alpha < \lambda\}. This construction ensures that the aleph hierarchy exhaustively lists all infinite cardinals under the axiom of choice (AC), which implies that every set can be well-ordered and thus assigned a unique aleph cardinal. Without AC, not all infinite cardinals need correspond to alephs, but AC guarantees that the alephs comprise the complete class of infinite cardinals. The first uncountable aleph, \aleph_1, is the cardinality of the set of all countable ordinals, marking the smallest cardinal larger than \aleph_0. In contrast to the aleph hierarchy, which arises from well-orderings, the beth numbers \beth_\alpha enumerate the cardinals generated by iterated power sets: \beth_0 = \aleph_0 and \beth_{\alpha+1} = 2^{\beth_\alpha}, with \beth_\lambda = \sup\{\beth_\alpha \mid \alpha < \lambda\} for limit \lambda. This sequence highlights the continuum hierarchy, where \beth_1 = 2^{\aleph_0} is the cardinality of the real numbers, potentially distinct from any \aleph_\alpha depending on additional axioms.

Continuum Hypothesis

The continuum hypothesis (CH), first formulated by Georg Cantor, states that there is no cardinal number strictly between \aleph_0, the cardinality of the natural numbers, and $2^{\aleph_0}, the cardinality of the power set of the natural numbers (the continuum); equivalently, $2^{\aleph_0} = \aleph_1. This hypothesis addresses a fundamental question in set theory about the immediate successor to the smallest infinite cardinal in the hierarchy of infinite cardinals. The generalized continuum hypothesis (GCH) extends CH to the entire aleph hierarchy, asserting that for every ordinal \alpha, $2^{\aleph_\alpha} = \aleph_{\alpha+1}. Under GCH, the power set operation on infinite cardinals yields precisely the next aleph in the sequence, eliminating any intermediate cardinals and providing a tight bound on cardinal exponentiation. This generalization captures the essence of CH while applying it universally across all infinite levels of the cardinal scale. In 1938, Kurt Gödel demonstrated the relative consistency of both CH and GCH with Zermelo-Fraenkel with the axiom of choice (ZFC), constructing the inner model L (the constructible universe) where GCH holds. This showed that if ZFC is consistent, then so is ZFC + GCH. In 1963, Paul Cohen proved the full independence of CH from ZFC by introducing the forcing technique, which constructs a model of ZFC + \negCH where $2^{\aleph_0} > \aleph_1. Together, these results establish that CH (and by extension aspects of GCH) is neither provable nor disprovable within ZFC, marking a pivotal limitation of the standard axioms of . The independence of CH has profound implications for cardinal arithmetic and beyond. In models where CH holds, such as Gödel's L, the structure of infinite cardinals is rigidly determined, with no gaps immediately following \aleph_0. In , CH implies the existence of pathological objects like Hamel bases for \mathbb{R} over \mathbb{Q} of cardinality \aleph_1, and equivalently, that \mathbb{R} can be covered by countably many such bases whose union excludes zero. Conversely, models violating CH allow for more flexible cardinal structures, affecting the possible sizes of bases and the behavior of functions on the reals, while influencing the construction of diverse set-theoretic universes.

Large Cardinals

Large cardinals refer to certain infinite cardinals that possess properties extending beyond those captured by the standard hierarchy, often axiomatized as assumptions that enhance the strength of . These axioms introduce notions of regularity and limits that surpass the power set operations on smaller cardinals, forming a hierarchy ordered by their implications for the consistency of ZFC and related principles. An inaccessible cardinal \kappa is defined as an uncountable strong limit cardinal, meaning its is itself and for every \lambda < \kappa, the power set $2^\lambda < \kappa. This property ensures that \kappa cannot be reached by iterating power set operations from smaller cardinals, making V_\kappa (the cumulative hierarchy up to \kappa) a model of ZFC. The existence of an inaccessible cardinal implies the consistency of ZFC, as V_\kappa satisfies the axioms independently of the full universe. Building on inaccessibility, a measurable cardinal \kappa is the smallest large cardinal admitting a non-principal \kappa-complete ultrafilter U on \kappa, which serves as a two-valued measure on the power set of \kappa. This ultrafilter allows for elementary embeddings j: V \to M with critical point \kappa and M^\kappa \subseteq M, capturing a form of reflection for subsets of \kappa. The existence of a measurable cardinal implies the consistency of ZFC plus the axiom of choice (AC) and the generalized continuum hypothesis (GCH), as the inner model L[U] constructed from the ultrafilter satisfies GCH while preserving AC. Moreover, every measurable cardinal is inaccessible, but the converse fails, placing measurables higher in the consistency strength hierarchy. The of extends further with compactness notions. A weakly compact cardinal \kappa is an inaccessible cardinal satisfying the tree property: every \kappa-tree has a cofinal branch of length \kappa, or equivalently, it is inaccessible and \Pi^1_1-indescribable. Strongly compact cardinals generalize this by requiring that for every \lambda \geq \kappa, there exists a fine \kappa-complete ultrafilter on \mathcal{P}_\kappa(\lambda), implying strong reflection properties for infinitary logics. Supercompact cardinals \kappa are even stronger, characterized by the existence, for every \lambda \geq \kappa, of an elementary embedding j: V \to M with critical point \kappa, j(\kappa) > \lambda, such that V_\lambda \subseteq M, ensuring \kappa is "super" in reflecting structures up to \lambda. At the pinnacle lies Vopěnka's principle, a global stating that for every proper class of structures in a common language, there are two members with an elementary embedding between them; this principle, which implies the existence of proper class many supercompact cardinals and implies the consistency of all smaller large cardinal axioms in its . These large cardinals play a crucial role in inner model theory, where they enable the construction of models like L[U] for measurables, which satisfy GCH despite the outer potentially violating it, thus resolving questions about the in restricted models. The strength forms a tower: the existence of a supercompact cardinal proves the of ZFC plus a measurable, which in turn proves the of an inaccessible, creating a linear order of implications that measures the "largeness" of the . Vopěnka's crowns this tower, implying the of supercompactness and providing a framework for category-theoretic simplifications in .

Applications

In Set Theory

In Zermelo–Fraenkel set theory with the (ZFC), cardinal numbers are identified with initial ordinals, which are ordinals \alpha such that no smaller ordinal \beta < \alpha is equinumerous to \alpha. This definition leverages the well-ordering theorem to assign to every set a unique cardinal, serving as the foundational measure of size in the theory. The universe of sets V is then stratified into the cumulative hierarchy V = \bigcup_{\alpha \in \mathrm{Ord}} V_\alpha, where each stage V_\alpha is built iteratively from the empty set via pairing, union, and power set operations, with cardinals determining the cardinality at limit stages and ensuring the hierarchy's transfinite progression. Forcing techniques, introduced by Paul Cohen, enable the construction of models where cardinal structures are altered while preserving ZFC axioms. For instance, the Lévy collapse forcing \mathrm{Col}(\omega, \aleph_1), consisting of finite partial functions from \omega_1 to \omega, adds a surjection from \omega onto \omega_1, thereby collapsing \aleph_1 to \aleph_0 in the extension without affecting smaller cardinals. More broadly, William Easton's theorem demonstrates that forcing can realize nearly arbitrary patterns of cardinal exponentiation: for any class function F assigning to each regular cardinal \kappa a value F(\kappa) > \kappa that is nondecreasing and satisfies König's inequality \mathrm{cf}(F(\kappa)) > \kappa, there exists a forcing extension where $2^\kappa = F(\kappa) for all regular \kappa. This flexibility highlights the independence of specific continuum function values from ZFC. Inner models provide canonical subuniverses for analyzing absoluteness. The constructible universe L, defined by Gödel via the hierarchy of definable sets L_\alpha, satisfies the generalized (GCH), where $2^\kappa = \kappa^+ for every infinite \kappa. In L, ordinals remain absolute, but cardinals from the ambient V may collapse; for example, if V contains non-constructible sets witnessing a larger , then \aleph_1^L < 2^{\aleph_0}^V, effectively collapsing the status of certain ordinals as cardinals in the inner model. Partition calculus, a branch of combinatorial , uses cardinals to extend to infinite settings, with results quantifying homogeneous subsets under colorings. The Erdős–Rado theorem, for instance, states that for any infinite cardinal \kappa and finite r, n, there exists \theta(\kappa, n, r) \leq (n+1)^{(r+1)^\kappa} such that any r-coloring of the n-element subsets of a set of size \theta yields a homogeneous subset of size \kappa. This stepping-up lemma underpins the partition properties defining Ramsey cardinals, where \kappa is Ramsey if every coloring of [\kappa]^{<\omega} has a homogeneous \kappa, illustrating how cardinals encode strong combinatorial regularity in set-theoretic models.

In Other Mathematics

In real analysis, sets of Lebesgue measure zero can have arbitrary cardinality up to that of the continuum, while every set of positive Lebesgue measure must have the full cardinality of the continuum. This interplay highlights how cardinal invariants constrain measurable structures on the real line. A striking application appears in the Banach-Tarski paradox, which relies on the axiom of choice to decompose a solid ball in three-dimensional Euclidean space into finitely many non-measurable pieces that can be rigidly reassembled to form two balls identical to the original. The pieces are non-measurable with respect to Lebesgue measure, underscoring the role of infinite cardinals in producing counterintuitive geometric decompositions. In , cardinal numbers quantify infinite-dimensional structures, such as the of real numbers over , which has dimension equal to the , $2^{\aleph_0}. This Hamel basis, whose existence requires the , consists of many linearly independent elements over \mathbb{Q}, illustrating how scales with set-theoretic size. Similarly, the on an generating set X has equal to that of X, as elements are finite words over X \cup X^{-1}, and the infinite case preserves the generator through reduced word representations. Topological applications of cardinals involve invariants that measure the complexity of spaces. The weight w(X) of a X is the smallest of a for its , while the density d(X) is the smallest of a dense subset; these invariants bound the overall of X and influence properties like separability. For instance, in the study of ordered topologies, a Suslin line is a complete dense linear without endpoints, of \aleph_1, that satisfies the countable chain condition but lacks a countable dense subset, providing a to the extension of the on the reals under certain set-theoretic assumptions. Illustrative examples further demonstrate cardinal concepts in other mathematical contexts. Hilbert's Grand paradox accommodates a new guest in a fully occupied hotel with countably infinitely many rooms by shifting occupants, revealing that the countable infinite cardinal \aleph_0 admits bijections with proper subsets, a property absent in finite sets. Likewise, the , constructed by choosing one representative from each of the reals under rational translations, has equal to the .

References

  1. [1]
    Cardinal Number -- from Wolfram MathWorld
    A cardinal number is a counting number, like 1, 2, 3, and in set theory, any method of counting sets using it gives the same result.<|control11|><|separator|>
  2. [2]
    [PDF] Notes on Cardinality
    Cardinality, denoted by |A|, measures the size of a set. For finite sets, it's the number of elements. For infinite sets, it's more complex.
  3. [3]
    Cardinality of important sets - Department of Mathematics at UTSA
    Nov 11, 2021 · The concept and notation are due to Georg Cantor, who defined the notion of cardinality and realized that infinite sets can have different ...<|control11|><|separator|>
  4. [4]
    Difference between cardinal and ordinal - Math Stack Exchange
    Sep 2, 2021 · Ordinal numbers list the objects in an order: the first, the second, etc, while cardinals count the numerosity of collections of objects.set theory - Cardinal Arithmetic versus Ordinal ArithmeticHow should I understand the formal definition of cardinal numbers ...More results from math.stackexchange.com
  5. [5]
    [PDF] Cardinal Numbers
    Cardinal numbers relate to the quantitative notion of 'how many' and measure how big sets are, or how many elements a set has.
  6. [6]
    [PDF] Infinity and its cardinalities
    The number of elements in a set is called the cardinal number, or cardinality of the set. The symbol n(A), read “n of A,” represents the cardinal number of set ...
  7. [7]
    [PDF] B. Cardinal Arithmetic - KSU Math
    Cardinal arithmetic defines sets with the same cardinality via bijective maps, and a cardinal number as an equivalence class of sets. The empty set has ...Missing: mathematics | Show results with:mathematics
  8. [8]
    [PDF] Section 0.8. Cardinal Numbers
    Jan 24, 2021 · Note. In this section, we consider a topic from set theory concerning the cardi- nalities of infinite sets. A more detailed coverage of this ...
  9. [9]
    Euclid's Elements, Book VII, Definitions 1 and 2 - Clark University
    A number is a multitude composed of units. Guide. These 23 definitions ... As mentioned above, Euclid has no postulates to elaborate the concept of number ...
  10. [10]
    Infinity - MacTutor History of Mathematics - University of St Andrews
    In this article we take the view that historically one cannot separate the philosophical and religious aspects from mathematical ones.
  11. [11]
    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 ...
  12. [12]
    The Early Development of Set Theory
    Apr 10, 2007 · In 1872 Cantor introduced an operation upon point sets (see below) and soon he was ruminating about the possibility to iterate that operation to ...Emergence · Consolidation · Critical Period · Bibliography
  13. [13]
    [PDF] On the Relations between Georg Cantor and Richard Dedekind
    Nov 30, 2024 · This paper gives a detailed analysis of the scientific interaction between Cantor and. Dedekind, which was a very important aspect in the ...
  14. [14]
    A glimpse of Cantor's paradise | plus.maths.org
    Jun 1, 2008 · Cantor adapted the method to show that there are an infinite series of infinities, each one astonishingly bigger than the one before. Today this ...<|control11|><|separator|>
  15. [15]
    [PDF] Axioms of Set Theory
    Axiomatisation of Set Theory. In 1908, Zermelo published in [102] his first ax- iomatic system consisting of seven axioms, which he called: 1. Axiom der ...
  16. [16]
    THE INDEPENDENCE OF THE CONTINUUM HYPOTHESIS - PNAS
    THE INDEPENDENCE OF THE CONTINUUM HYPOTHESIS. BY PAUL J. COHEN*. DEPARTMENT OF MATHEMATICS, STANFORD UNIVERSITY. Communicated by Kurt Godel, September 30, 1963.
  17. [17]
    Finite Set -- from Wolfram MathWorld
    A set X whose elements can be numbered through from 1 to n, for some positive integer n. The number n is called the cardinal number of the set.
  18. [18]
    finite - PlanetMath.org
    Mar 22, 2013 · A set S S is finite if there exists a natural number n n and a bijection from S S to n n . Note that we are using the set theoretic ...
  19. [19]
    cardinal number in nLab
    Aug 22, 2025 · The cardinality of a set S S is the smallest possible ordinal rank of any well-order on S S . In other words, it is the smallest ordinal number ...
  20. [20]
    Infinite Set -- from Wolfram MathWorld
    A set of elements S is said to be infinite if the elements of a proper subset S^' can be put into one-to-one correspondence with the elements of S.
  21. [21]
    aleph numbers - PlanetMath
    Mar 22, 2013 · The aleph numbers are infinite cardinal numbers defined by transfinite recursion, as described below. They are written ℵα , where ℵ is aleph, ...
  22. [22]
    Aleph-0 -- from Wolfram MathWorld
    The set theory symbol aleph_0 refers to a set having the same cardinal number as the "small" infinite set of integers.
  23. [23]
    cardinal arithmetic - PlanetMath.org
    Mar 22, 2013 · Definitions. Let κ and λ be cardinal numbers , and let A and B be disjoint sets such that |A|=κ and |B|=λ . ( Here |X| denotes the cardinality ...
  24. [24]
    limit cardinal - PlanetMath
    Mar 22, 2013 · Every strong limit cardinal is a limit cardinal ... infinite cardinal λ λ . The three smallest limit cardinals are 0 0 , ℵ0 ℵ 0 and ℵω ℵ ω .
  25. [25]
    Countably Infinite -- from Wolfram MathWorld
    Countably infinite sets have cardinal number aleph-0. Examples of countable sets include the integers, algebraic numbers, and rational numbers.
  26. [26]
    Cantor Diagonal Method -- from Wolfram MathWorld
    By applying this argument infinitely many times to the same infinite set, it is possible to obtain an infinite hierarchy of infinite cardinal numbers. See also.
  27. [27]
    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.Missing: textbook | Show results with:textbook
  28. [28]
    Sets:Cardinality - Department of Mathematics at UTSA
    Jan 30, 2022 · Cardinality is a measure of the 'number of elements' in a set, also called its size, and is denoted by |A|.
  29. [29]
    [PDF] The Ordinal Numbers and Transfinite Induction - Purdue Math
    Sep 14, 2015 · Intuitively, one would think to define the cardinal numbers as equivalence ... The von Neumann construction of the natural numbers is as follows:.
  30. [30]
    [PDF] RELATIONS BETWEEN SOME CARDINALS IN THE ABSENCE OF ...
    Abstract. If we assume the axiom of choice, then every two cardinal numbers are comparable. In the absence of the axiom of choice, this is no longer so. For.
  31. [31]
    [PDF] RELATIONS BETWEEN SOME CARDINALS IN THE ABSENCE OF ...
    A cardinal number m is an aleph if it contains a well-ordered set. So, the cardinality of each ordinal is an aleph. Remember that the axiom of choice is ...
  32. [32]
    [PDF] Axiom of Choice, Zorn's Lemma and the Well-ordering Principle
    The following statements are equivalent: (i) Axiom of Choice (AC'): for every non-empty X, there is a choice func- tion, i.e. ...
  33. [33]
    [PDF] Axiom of Choice and Zorn's Lemma - Cornell Mathematics
    Theorem 4 (Well-ordering theorem). There exists a well ordering on any set A ... A third equivalent form of the axiom of choice is Zorn's Lemma. We ...
  34. [34]
    [PDF] Notes on the Axiom of Choice - ETH Zürich
    Jul 24, 2025 · The purpose of these notes is to prove that the Axiom of Choice, the Lemma of Zorn, and the Well Ordering Principle are equivalent to each ...
  35. [35]
    [PDF] A Dedekind Finite Borel Set - Web.math.wisc.edu
    In this paper we prove three theorems about the theory of Borel sets in models of ZF without any form of the axiom of choice.<|separator|>
  36. [36]
  37. [37]
    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 ...
  38. [38]
    [PDF] Cardinal Arithmetic - Open Logic Project Builds
    Exponentiation is similar: we are simply generalising the thought from the finite to the transfinite. Indeed, in certain ways, transfinite cardinal arithmetic ...<|control11|><|separator|>
  39. [39]
    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.
  40. [40]
    [PDF] Set Theory (MATH 6730) The Axiom of Choice. Cardinals and ...
    WOP (Well-Ordering Principle). For every set B there exists a well-ordering (B,≺). ZLm (Zorn's Lemma). If (D, <) is a partial order such that. (∗) every ...
  41. [41]
    [PDF] Chapter 9. ARITHMETIC OF CARDINAL NUMBERS - unipi
    As we show, the Generalized Continuum Hypothesis greatly simplifies the cardinal exponentiation; in fact, the operation can then be evaluated by very simple ...
  42. [42]
    [PDF] Cantor-Schroeder-Bernstein Theorem
    Feb 19, 2005 · This is the key result that allows comparison of infinities. Perhaps it is the first serious theorem in set theory.
  43. [43]
    [PDF] Lecture 5: Cardinals
    Feb 2, 2009 · Definition 5.16 (Cofinality). • X ⊆ α is cofinal in α iff sup(X) = α. • A map f : β → α is a cofinal map iff rng f is cofinal in α. • The ...
  44. [44]
    Beiträge zur Begründung der transfiniten Mengenlehre
    Cantor, G. Beiträge zur Begründung der transfiniten Mengenlehre. Math. Ann. 46, 481–512 (1895). https://doi.org/10.1007/BF02124929
  45. [45]
    Basic set theory - Stanford Encyclopedia of Philosophy
    5.1 Cardinals​​ The cardinality, or size, of a finite set \(A\) is the unique natural number \(n\) such that there is a bijection \(F:n\to A\).
  46. [46]
    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 ...
  47. [47]
    The Consistency of the Axiom of Choice and of the Generalized ...
    The Consistency of the Axiom of Choice and of the Generalized Continuum-Hypothesis. Kurt GödelAuthors Info & Affiliations. December 15, 1938. 24 (12) 556-557.
  48. [48]
    [PDF] The cardinality of Hamel bases of Banach spaces
    if one considers Ras a Banach space over Q, then the continuum hypothesis is equivalent to the statement, that R can be covered by countably many Hamel bases ( ...
  49. [49]
    [PDF] MATH 6870: SET THEORY - Cornell Mathematics
    A cardinal is weakly inaccessible if it is an uncountable regular limit cardinal and strongly inaccessible if it is a uncountable regular strong limit cardinal.
  50. [50]
    The Higher Infinite: Large Cardinals in Set Theory ... - SpringerLink
    In stockBook Title: The Higher Infinite · Book Subtitle: Large Cardinals in Set Theory from Their Beginnings · Authors: Akihiro Kanamori · Series Title: Springer ...
  51. [51]
  52. [52]
    Vopěnka's principle and Vopěnka cardinals | cantors-attic
    Vopěnka's principle is a large cardinal axiom at the upper end of the large cardinal hierarchy that is particularly notable for its applications to category ...
  53. [53]
    [PDF] Set-Theoretical Background 1.1 Ordinals and cardinals - UB
    Feb 11, 2019 · Equivalently, a cardinal κ is regular if it is equal to its own cofinality, where the cofinality of a limit ordinal α is the least limit ordinal ...<|separator|>
  54. [54]
    cumulative hierarchy - PlanetMath
    Mar 22, 2013 · The cumulative hierarchy of sets is defined by transfinite recursion as follows: we define V0=∅ V 0 = ∅ and for each ordinal α α we define ...
  55. [55]
    Levy collapse - PlanetMath
    Apr 16, 2013 · Given any cardinals κ κ and λ λ in M 𝔐 , we can use the Levy collapse to give a new model M[G] 𝔐 ⁢ [ G ] where λ=κ λ = κ .Missing: aleph1 aleph0
  56. [56]
    [PDF] Easton's theorem and large cardinals
    Mar 6, 2008 · The continuum function α 7→ 2α on regular cardinals is known to have great free- dom. Say that F is an Easton function iff for regular ...
  57. [57]
    [PDF] Gödel's Constructible Universe
    Jan 16, 2020 · We covered elementary notions in the. Zermelo-Fraenkel set theory, ordinal and cardinal numbers, models and Gödel's. Constructible Universe.
  58. [58]
    [PDF] A partition calculus in set theory
    We shall replace the sets S and A by sets of a more general kind and the unordered pairs, as is the case al- ready in the theorem proved by Ramsey, by systems ...
  59. [59]
    Do sets with positive Lebesgue measure have same cardinality as R?
    Dec 15, 2009 · A proof is given that every measurable subset with cardinality less than that of R has Lebesgue measure zero.
  60. [60]
    The axiom of choice and Banach-Tarski paradoxes
    We shall use the axiom of choice to prove an extremely wimpy version of the Banach Tarski paradox, to wit: Theorem. It is possible to take a ...
  61. [61]
    [PDF] Part 1. Cardinality of Sets
    For the finite cardinals, the arithmetic operations defined above correspond exactly to those of the non-negative integers. There is no ...
  62. [62]
    [PDF] Free groups - UNL Math
    If both of them are infinite then the result follows from an observation that cardinality of a group generated by an infinite set A is equal to |A|. This.
  63. [63]
    Cardinalities of dense sets - ScienceDirect.com
    A cardinal invariant on a topological space X, called its strong density, is introduced as the supremum of the densities of the dense subsets of X. The ...
  64. [64]
    A model with Suslin trees but no minimal uncountable linear orders ...
    Mar 12, 2018 · We show that the existence of a Suslin tree does not necessarily imply that there are uncountable minimal linear orders other than ω 1 ...
  65. [65]
    Hilbert's hotel | plus.maths.org - Millennium Mathematics Project
    Feb 13, 2017 · Suppose an infinite number of new guests arrive, forming an orderly queue outside the hotel. In this case, ask each existing guest to move into ...
  66. [66]
    What is the cardinality of a non-measurable set?
    Jul 22, 2016 · This number κ has been studied in the context of cardinal characteristics (or "cardinal invariants") of the continuum, where it is denoted non(L) ...Cardinality of a set of positive Lebesgue measureGiven a non-measurable set and a cardinal does there exist a ...More results from math.stackexchange.com