Fact-checked by Grok 2 weeks ago

Index set

In , an index set is a set I whose elements, known as indices, serve to label or parameterize the members of a family of objects, such as sets, functions, or elements, enabling the systematic description of potentially infinite or arbitrarily structured collections. This concept is fundamental in , where it facilitates the extension of finite operations to transfinite or general cases without relying on explicit enumeration. An of sets, denoted \{A_i \mid i \in I\} or \{A_i\}_{i \in I}, associates each index i \in I with a corresponding set A_i, forming a structured collection that generalizes sequences (when I = \mathbb{N}) to arbitrary index sets. The index set I can be any set, including finite sets, the natural numbers, or even uncountable sets like the real numbers, allowing flexibility in modeling diverse mathematical structures. Key operations on indexed families include the \bigcup_{i \in I} A_i = \{ x \mid \exists i \in I \text{ such that } x \in A_i \}, which collects all elements appearing in at least one A_i, and the \bigcap_{i \in I} A_i = \{ x \mid \forall i \in I, x \in A_i \}, which consists of elements common to every A_i. These operations satisfy inclusion properties, such as \bigcap_{i \in I} A_i \subseteq A_j \subseteq \bigcup_{i \in I} A_i for any j \in I, and extend to indexed forms, like \left( \bigcup_{i \in I} A_i \right)^c = \bigcap_{i \in I} A_i^c. Examples illustrate the utility of index sets; for instance, with I = \mathbb{N} and A_n = \{1, 2, \dots, n\}, the \bigcup_{n \in \mathbb{N}} A_n = \mathbb{N} covers all natural numbers, while the \bigcap_{n \in \mathbb{N}} A_n = \{1\} isolates the common initial element. Another case is I = \mathbb{Z} \setminus \{0\} with A_i as the multiples of i, yielding \bigcup_{i \in I} A_i = \mathbb{Z} and \bigcap_{i \in I} A_i = \{0\}. Index sets are essential in advanced for defining Cartesian products, direct sums, and limits over arbitrary index classes, appearing in , , and to handle infinite constructions like the real numbers as unions over rational intervals or vector spaces as direct sums of subspaces. They also support concepts like pairwise disjoint families, where A_i \cap A_j = \emptyset for i \neq j, which is crucial in measure theory and probability for partitioning sample spaces.

Definition and Notation

Formal Definition

In , an is defined as any set I whose elements, referred to as indices, serve to label or parameterize the members of a family of mathematical objects, such as sets, functions, or individual elements. This structure allows for the systematic organization of collections where the indices provide a means of identification without implying any inherent order or on I itself. A indexed by I is formally a f: I \to \mathcal{C}, where \mathcal{C} is the collection of the relevant mathematical objects; for instance, when dealing with sets, this corresponds to a f: I \to \mathcal{P}(X) for some X, with \mathcal{P}(X) denoting the power set of X. Each element f(i) for i \in I is then labeled by the i, enabling operations like unions or intersections over the entire family via the indices. This functional perspective ensures that the family is well-defined and avoids ambiguities in unindexed collections. Unlike sequences, which are specifically indexed by the natural numbers \mathbb{N} and thus limited to countable structures, an index set I can possess arbitrary cardinality and no required linear order, accommodating uncountable or partially ordered indexings as needed in advanced set-theoretic constructions.

Common Notations

In mathematics, index sets are commonly denoted by uppercase letters such as I, J, or \Lambda, with their elements typically represented by corresponding lowercase letters like i \in I, j \in J, or \lambda \in \Lambda. Capital Greek letters, such as \Lambda or \Gamma, are often employed for arbitrary indexing sets, while lowercase Greek letters denote the indices, particularly in contexts involving uncountable or general ordinals. An indexed family of sets is standardly notated as \{A_i \mid i \in I\} or, more compactly, \{A_i\}_{i \in I}, where the subscript i emphasizes the indexing by elements of I. The parenthetical form (A_i)_{i \in I} is also frequently used, especially to distinguish it from unordered collections. Alternative notations include uppercase letters for the entire family, such as F = \{f_\alpha \mid \alpha \in A\}, where A serves as an arbitrary index set and \alpha as a generic index. Conventions vary by field: in , the index set is often I = \mathbb{N} (the natural numbers) for sequences, with notation like (a_n)_{n=1}^\infty or \{a_n \mid n \in \mathbb{N}\}, reflecting the ordered, countable nature of sequences. In general , however, I can be any set, allowing for uncountable or arbitrary indexing without inherent order. Formally, an is a \phi: I \to \mathrm{Obj}, where \mathrm{Obj} is the class of relevant objects (such as sets), and \phi(i) = A_i for each i \in I. \phi: I \to \mathrm{Obj}, \quad \phi(i) = A_i This functional perspective underscores that the family is the image or of \phi, often written as \{\phi(i) \mid i \in I\}.

Indexed Families

Indexed Families of Sets

An indexed family of sets, also known as an indexed collection of sets, consists of a collection \{A_i \mid i \in I\} where I is an and each A_i is a of some set X. This parameterization allows for the systematic organization of sets using elements of I as indices, providing a structured way to handle potentially infinite collections without relying on . Key operations on an indexed family \{A_i \mid i \in I\} include the indexed and indexed . The indexed is defined as \bigcup_{i \in I} A_i = \{ x \in X \mid \exists i \in I \text{ such that } x \in A_i \}, which collects all elements belonging to at least one set in the family. The indexed is \bigcap_{i \in I} A_i = \{ x \in X \mid \forall i \in I, \, x \in A_i \}, comprising elements common to every set in the family. These operations generalize finite unions and intersections to arbitrary index sets, enabling the analysis of complex set structures. A disjoint indexed family \{A_i \mid i \in I\} satisfies A_i \cap A_j = \emptyset for all distinct i, j \in I. In such cases, the union simplifies significantly, and under the , the cardinality of the union equals the cardinal of the individual cardinalities: \left| \bigcup_{i \in I} A_i \right| = \sum_{i \in I} |A_i|. This relation holds because the disjointness ensures no overlapping elements, allowing a direct correspondence via choice functions for infinite families. Indexed families often relate to partitions of a set X, where \{A_i \mid i \in I\} is disjoint and \bigcup_{i \in I} A_i = X, with pairwise empty intersections for distinct indices. This decomposes X into non-overlapping subsets indexed by I, facilitating proofs and constructions in . The of the indexed can be expressed using the characteristic functions \chi_{A_i} of the individual sets: \chi_{\bigcup_{i \in I} A_i}(x) = 1 - \prod_{i \in I} (1 - \chi_{A_i}(x)), which evaluates to 1 if x belongs to at least one A_i and 0 otherwise, leveraging the product form to capture the existential condition.

Indexed Families of Functions

An indexed family of functions, also known as a family of functions parametrized by an index set, consists of a collection of functions \{ f_i : X \to Y \mid i \in I \}, where I is the index set, X is the common , and Y is the common . Formally, this family is defined as a function f : I \to Y^X, where Y^X denotes the set of all functions from X to Y, and f_i = f(i) for each i \in I. This structure allows for the systematic organization of functions sharing the same domain and codomain, facilitating operations and analyses that depend on the indexing. Pointwise operations on such families are defined componentwise on the Y, assuming Y supports the relevant , such as being a or . For , if I is finite or the family has finite support at each point in X, the is the \left( \sum_{i \in I} f_i \right)(x) = \sum_{i \in I} f_i(x) for all x \in X. Similarly, multiplication or can be defined as (c \cdot f_i)(x) = c \cdot f_i(x) or (f_i \cdot f_j)(x) = f_i(x) \cdot f_j(x), extending to the entire family where holds . These operations preserve the indexed , yielding another family \{ g_i : X \to Y \mid i \in I \}. In analytic contexts, properties like uniform bounds and are examined relative to the index set. The supremum norm of the family is given by \sup_{i \in I} \| f_i \|, where \| f_i \| = \sup_{x \in X} |f_i(x)| assuming Y = \mathbb{R} or \mathbb{C}, providing a measure of the family's overall magnitude. For directed index sets I with a directed , pointwise limits such as \lim_{i \to \alpha} f_i(x) for x \in X and \alpha a limit point in I define a limiting , with requiring \sup_{x \in X} |\lim_{i \to \alpha} f_i(x) - f_\alpha(x)| \to 0 as i \to \alpha. These concepts ensure the family behaves cohesively under limiting processes. Indexed families of functions also arise in as components of morphisms between functors. Specifically, a \eta : F \dashv G between functors F, G : \mathcal{C} \to \mathcal{D} is an of morphisms \{ \eta_C : F(C) \to G(C) \mid C \in \mathrm{Ob}(\mathcal{C}) \}, one for each object C in the category \mathcal{C}, satisfying the naturality axiom that for every morphism h : C \to C' in \mathcal{C}, the diagram F(h) \circ \eta_C = \eta_{C'} \circ G(h) commutes. This interprets the index set as the class of objects in \mathcal{C}. For composition, given a family \{ f_i : X \to Y \mid i \in I \} and a function g : Y \to Z, the composed family is \{ g \circ f_i : X \to Z \mid i \in I \}, preserving the indexing.

Properties and Cardinality

Well-Ordering and Choice

A well-ordered index set I is equipped with a total order such that every nonempty subset of I has a least element. This property ensures that no infinite descending chains exist in I, allowing for the principle of transfinite induction: if a property holds for the least element of I and, assuming it holds for all elements less than some \alpha \in I, it also holds for \alpha, then the property holds for all elements of I. Such index sets facilitate the construction of transfinite sequences, where each term is defined inductively over the order of I, extending finite induction to infinite cases without gaps or cycles. Index sets often take the form of ordinal numbers in , which are themselves well-ordered sets under the membership relation. For instance, the ordinal \omega, the set of natural numbers ordered by \in, serves as an index set for countable sequences, while the class of all ordinals, denoted On, can index proper classes of sets in transfinite constructions. This identification allows arbitrary well-ordered index sets to be isomorphic to unique ordinals, providing a structure for indexing families beyond finite or countable ranges. The axiom of choice (AC) states that for any indexed family of nonempty sets (A_i)_{i \in I}, there exists a choice function selecting one element from each A_i. AC implies the well-ordering theorem, which asserts that every set admits a well-ordering, thereby enabling any index set I to be well-ordered and thus behave like an ordinal under AC. This equivalence, first proven by Zermelo in 1904, underpins the ability to index families over arbitrary sets by imposing a well-order. Zorn's lemma, equivalent to AC, applies to partially ordered families indexed by a set I: if every chain in the poset has an upper bound, then the poset contains a maximal element. In the context of indexed families, this manifests in the existence of maximal chains or selections, constructed via transfinite sequences over well-ordered indices derived from AC. The proof typically builds an ordinal-indexed ascending sequence of elements until a maximal one is reached, leveraging the well-ordering of the index set. In set theories without AC, such as certain models of ZF, some index sets may not admit well-orderings, precluding transfinite induction over them and affecting the existence of choice functions or bases for indexed families. For example, the real numbers cannot be well-ordered in Cohen's forcing model, implying that index sets of continuum cardinality may lack the least-element property for subsets, thus restricting inductive constructions.

Cardinality Considerations

The cardinality of an indexed family of sets depends significantly on whether the index set I is finite or infinite. When I is finite, say with n elements, an indexed family \{A_i \mid i \in I\} reduces to an n-tuple of sets, and the cardinality of the union \bigcup_{i \in I} A_i is at most n \cdot \sup_{i \in I} |A_i|, while the cardinality of the product \prod_{i \in I} X_i is exactly the finite product \prod_{i \in I} |X_i|, computable without invoking additional axioms beyond basic set theory. In contrast, for infinite I, enumerating or selecting elements from the family often requires the axiom of choice (AC); for instance, AC ensures the existence of choice functions for infinite families of nonempty sets, enabling well-defined operations on cardinalities that might otherwise be ambiguous. For pairwise disjoint families \{A_i \mid i \in I\} of nonempty sets with infinite I, the cardinality of the union satisfies \left| \bigcup_{i \in I} A_i \right| = |I| \cdot \sup_{i \in I} |A_i|, where the supremum is taken in the sense of cardinal arithmetic under AC; this follows from the disjoint union construction, which embeds each A_i injectively into the total union via tagging with elements of I. Similarly, the set of all indexed families of functions \{f_i : X \to Y \mid i \in I\} has cardinality at most |Y|^{|X| \cdot |I|}, as it corresponds to the function space Y^{X \times I}, reflecting the exponential growth induced by the index set's size. An indexed family of distinct subsets \{A_i \subseteq X \mid i \in I\} induces an injection I \to \mathcal{P}(X) via i \mapsto A_i, implying |I| \leq |\mathcal{P}(X)| = 2^{|X|}, a bound central to comparing sizes in set-theoretic constructions. The product of cardinalities over an index set also relies on AC for infinite cases: under AC, \left| \prod_{i \in I} X_i \right| = \prod_{i \in I} |X_i|, where the right-hand side denotes the cardinal product, ensuring the Cartesian product is nonempty and its size matches the indexed multiplication of individual cardinalities. If |I| = 2^{\aleph_0} (the continuum), such indexed families can model uncountable structures without intermediate cardinalities, as per the continuum hypothesis (CH), which posits no sets exist with cardinality strictly between \aleph_0 and $2^{\aleph_0}; this allows families indexed by the reals to parameterize continuous phenomena while respecting CH's constraints on size hierarchies.

Applications

In Set Theory

In set theory, index sets facilitate the construction of direct sums and products of families of sets. The direct sum \oplus_{i \in I} A_i, also known as the disjoint union, is defined as the union \bigcup_{i \in I} (\{i\} \times A_i), where each A_i is embedded into a distinct copy via the index i, ensuring the components remain disjoint. This construction relies on the index set I to tag elements and prevent overlap, allowing the formation of a single set from the family \{A_i \mid i \in I\}. The direct product \prod_{i \in I} A_i, on the other hand, consists of all functions f: I \to \bigcup_{i \in I} A_i such that f(i) \in A_i for every i \in I, effectively selecting one element from each A_i coordinated by the index set. This equates to the set of all choice functions over the family, underscoring the role of I in structuring the Cartesian product for arbitrary index sets. Transfinite recursion employs well-ordered index sets to define sets iteratively across ordinals, building complex structures from simpler ones. For a well-ordered index set I (typically an ordinal), a F: V \to V (where V is the universe of sets) defines sets by : the value at stage \alpha \in I depends on prior stages \beta < \alpha. A seminal application is the von Neumann hierarchy, which constructs the cumulative hierarchy of sets indexed by ordinals: V_0 = \emptyset, and for successor ordinals \alpha = \beta + 1, V_\alpha = \mathcal{P}(V_\beta) (the power set of V_\beta); for limit ordinals \lambda, V_\lambda = \bigcup_{\beta < \lambda} V_\beta. This hierarchy, introduced by John von Neumann, models the iterative conception of sets, with each V_\alpha containing all sets of rank less than \alpha. The axiom schema of replacement leverages index sets to guarantee the existence of new sets from existing ones via definable substitutions. Specifically, for any set A and formula \phi(x, y) such that for every x \in A there is a unique y with \phi(x, y), the set \{y \mid \exists x \in A \, \phi(x, y)\} exists; here, A serves as the index set for the family \{y_x \mid x \in A\} defined by \phi. This schema ensures that images of sets under class functions are sets, enabling constructions like the transitive closure or ordinal exponentiation without exceeding set-sized bounds. Replacement is crucial for transfinite inductions, as it collects the outputs of recursions over well-ordered index sets into a single set. When the index set is a proper class, such as the class of all ordinals \mathrm{On}, constructions extend to class-sized families, modeling the entire universe. The von Neumann universe V is the proper class \bigcup_{\alpha \in \mathrm{On}} V_\alpha, where each V_\alpha is set-sized but the union over the class index \mathrm{On} is proper. This class-indexed union captures all sets under the iterative axiom, with proper classes like \mathrm{On} or V itself serving as index sets for global structures, such as the class of all ordinals or the cumulative hierarchy itself. Such extensions highlight how index sets, even proper classes, underpin foundational decompositions in set theory.

In Topology and Analysis

In topology, index sets facilitate the description of bases and covers through indexed families of open sets. A basis for the topology on a space X is an indexed family \{U_i \mid i \in I\} of open subsets such that every open set in the topology can be expressed as a union of elements from some subfamily of \{U_i \mid i \in I\}. This structure ensures that the basis generates the entire topology while allowing for efficient local characterizations of open sets. Similarly, an open cover of X is an indexed family \{U_i \mid i \in I\} of open sets whose union equals X, often used to study compactness via subcovers. Nets and filters in topology rely on index sets equipped with a directed partial order to generalize sequences beyond metric spaces. A directed set I serves as the index set for a net (x_i)_{i \in I} in a topological space X, where convergence to a point x \in X is defined such that for every neighborhood U of x, the set \{i \in I \mid x_i \in U\} is cofinal in I. x_i \to x \iff \forall U \ni x, \{i \in I \mid x_i \in U\} \text{ is cofinal in } I. This cofinality condition captures "eventual" membership in U, enabling nets to detect continuity and limits in arbitrary topological spaces; for instance, the index set I = [0,1] with the usual order can index nets approximating continuous functions on compact intervals. Filters, dual to nets, use index sets to define adherent points via bases of neighborhoods. In analysis, index sets underpin series expansions and integrals by providing ordered structures for convergence. Orthogonal expansions in Hilbert spaces, such as Fourier series, represent elements as sums \sum_{i \in I} \langle f, e_i \rangle e_i over countable index sets I (e.g., \mathbb{Z} or \mathbb{N}), where \{e_i\} forms an orthonormal basis ensuring Parseval's identity for the norm. In Banach spaces, Schauder bases generalize this to indexed families \{e_i \mid i \in I\} (typically countable) such that every x \in X admits a unique representation x = \sum_{i \in I} c_i e_i with convergence in the norm, providing dense spanning with biorthogonal functionals for coefficient extraction. Integrals in real analysis emerge as limits of nets over directed sets of partitions; for the Riemann integral on [a,b], the index set consists of tagged partitions ordered by refinement, with Riemann sums forming a net that converges to the integral when the function is Riemann-integrable.

Examples

Basic Examples

A finite index set provides a straightforward illustration of an indexed family. Consider the index set I = \{1, 2, 3\} and the corresponding family of sets \{A_i \mid i \in I\} where A_1 = \{a\}, A_2 = \{b\}, and A_3 = \{c\}. The union of this family is \bigcup_{i \in I} A_i = \{a, b, c\}, demonstrating how indexing allows for systematic combination of distinct elements across the sets. For a countable index set, the natural numbers \mathbb{N} often serve as indices for sequences, which are indexed families of real numbers or sets. An example is the sequence (a_n)_{n=1}^\infty defined by a_n = \frac{1}{n}, where each a_n can be viewed as a singleton set \{a_n\}. This family converges to 0 as n \to \infty, since for any \epsilon > 0, there exists N \in \mathbb{N} such that for all n > N, |a_n - 0| < \epsilon, illustrating the utility of countable indexing in analyzing limits. The empty index set I = \emptyset yields the empty family of sets, with no elements to index. In this case, the union \bigcup_{i \in \emptyset} A_i = \emptyset, as there exists no i \in \emptyset such that an element belongs to some A_i. Conversely, the intersection \bigcap_{i \in \emptyset} A_i is the universal set (or the ambient space containing all possible elements), due to the vacuous truth that every element satisfies the condition of belonging to all A_i when no such sets exist. In the context of products, an index set I = \{x, y\} indexes a family \{A_i \mid i \in I\}, and the Cartesian product \prod_{i \in I} A_i consists of all functions f: I \to \bigcup A_i such that f(i) \in A_i for each i. For instance, if A_x = \{\alpha, \beta\} and A_y = \{\gamma, \delta\}, then \prod_{i \in I} A_i = A_x \times A_y = \{(\alpha, \gamma), (\alpha, \delta), (\beta, \gamma), (\beta, \delta)\}, reducing to the standard ordered pairs. Finally, the trivial case of a singleton index set I = \{*\} simplifies an indexed family to a single object, such as \{A_* \mid * \in I\} where the family is just \{A_*\}. Operations like union or product then revert to the set itself, as \bigcup_{i \in I} A_i = A_* and \prod_{i \in I} A_i = A_*, highlighting how indexing encompasses ordinary sets as special cases.

Advanced Examples

In set theory, an advanced example of an indexed family arises when the index set is the ordinal I = \omega + 1, which consists of all finite ordinals together with the least infinite ordinal \omega. This structure enables the definition of transfinite sequences, where the term at index \omega serves as the limit of the preceding finite-indexed terms, often constructed as the supremum or union of those terms. For instance, consider a family (A_\alpha)_{\alpha \in \omega + 1} of sets where A_n = \{ n \} for finite n \in \omega, and A_\omega = \bigcup_{n < \omega} A_n; this illustrates how ordinal indexing captures transfinite progression beyond countable sequences. Another sophisticated case involves an uncountable index set, such as I = \mathbb{R}, with the family of singletons (\{r\})_{r \in \mathbb{R}}. Here, each set in the family contains exactly one real number, and their union yields \bigcup_{r \in \mathbb{R}} \{r\} = \mathbb{R}, demonstrating how an uncountable indexing reconstructs the continuum from disjoint components. The intersection, however, is empty, as no element belongs to every singleton. This example highlights the utility of uncountable indices in partitioning and reassembling sets of continuum cardinality. For applications in topology, consider the directed set I = \mathbb{Q} equipped with the usual order, which forms a directed partial order since any two rationals have an upper bound. This index set is used to define Cauchy nets in metric spaces, generalizing sequences to handle non-sequential convergence. A net (x_t)_{t \in \mathbb{Q}} in a metric space (X, d) is Cauchy if for every \epsilon > 0, there exists t_0 \in \mathbb{Q} such that d(x_s, x_t) < \epsilon for all s, t \geq t_0; in complete metric spaces, such nets converge, providing a tool for limits in spaces without a countable basis. To illustrate cardinality in indexed families, take the index set I = \mathcal{P}(\mathbb{N}), the power set of the natural numbers, which has $2^{\aleph_0}, the . This uncountable index set can index a family (A_S)_{S \in \mathcal{P}(\mathbb{N})} where each A_S = S, directly associating subsets with themselves; the size of I underscores that the power set operation exponentially increases beyond \aleph_0, as no exists between \mathbb{N} and its subsets. Finally, in linear algebra over fields, the Hamel basis for \mathbb{R} as a over \mathbb{Q} provides an example reliant on the . This basis I is a linearly independent set such that every is a unique finite rational of elements from I, and under the , |I| = 2^{\aleph_0}, matching the of the . Thus, \mathbb{R} decomposes as a \bigoplus_{i \in I} \mathbb{Q} \cdot e_i, where (e_i)_{i \in I} are the basis vectors, emphasizing the uncountable nature required for spanning \mathbb{R}.

References

  1. [1]
    2.3Indexed families of sets - SIUE
    We call a set an index set if it plays a role similar to that of N N in the previous example. Moreover, we say that it indexes the family A, ...
  2. [2]
    1.6 Families of Sets
    Suppose I is a set, called the index set, and with each i∈I we associate a set Ai. We call {Ai:i∈I} an indexed family of sets. Sometimes this is denoted by ...
  3. [3]
  4. [4]
    [PDF] Notes on Sets, Mappings, and Cardinality - UC Berkeley math
    Oct 18, 2010 · Members of I are referred to as indices. Indexed families are often denoted {Ei}i∈I , or even {Ei}, if the indexing set is clear from the ...
  5. [5]
    [PDF] Chapter 1 - Proof, Sets, and Functions
    In this case, the set N is called the index set and the set S is called an indexed set or indexed family. Since this concept is used so often in mathematics, we ...
  6. [6]
    Index Set -- from Wolfram MathWorld
    A set whose members index (label) members of another set. For example, in the set A= union _(k in K)A_k, the set K is an index set of the set A.
  7. [7]
    [PDF] 2.6. Indexed Sets
    Jan 1, 2022 · Note/Definition. We can describe a set by associating its elements with members of an index set. We can write B = {bi | i ∈ I} where set I ...
  8. [8]
    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.<|control11|><|separator|>
  9. [9]
    family in nLab
    Jan 27, 2025 · The easiest case is when the elements are from a set S S ; then an I I -indexed family of elements of S S , also called an indexed subset of S S ...
  10. [10]
    5.5: Indexed Families of Sets
    ### Summary of Indexed Families of Sets Content
  11. [11]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · There it is shown that, given any indexed family \(\sA\) of nonempty sets, there is a natural bijection between choice functions on \(\sA\) and ...
  12. [12]
    [PDF] Math 230.01, Fall 2012: HW 5 Solutions
    c) For any collection of events A1, ··· An, the indicator of their union is. I∪Ai = 1 − (1 − IA1 )(1 − IA2 ) ··· (1 − IAn ). SOLUTION. Note that. [∪n i=1Ai] c.
  13. [13]
    [PDF] Foundations of Algebraic Specification and Formal ... - mimuw
    Sep 29, 2010 · sorted sets. An S-sorted function f:X → Y is an S-indexed family of functions f = hfs:Xs → Ysis∈S; X is called the domain (or source) of ...
  14. [14]
    [PDF] arXiv:1906.04871v1 [math.CO] 12 Jun 2019
    Jun 12, 2019 · Recall that an indexed family of functions FE from A to k is ... of pointwise convergence. In this case, our construction becomes a ...<|separator|>
  15. [15]
    [PDF] arXiv:2307.10889v3 [math.PR] 10 May 2024
    May 10, 2024 · theorem for the ǫ-indexed family of functions ǫ−1/2h(ǫ2t, ǫx), see ... such that fn converges to f pointwise. Since for a sequence of ...
  16. [16]
    [PDF] arXiv:2307.16076v2 [math.CT] 16 Aug 2024
    Aug 16, 2024 · It establishes an equivalence between indexed categories and Grothendieck fibrations. The construction reorganizes the data of an indexed family ...<|control11|><|separator|>
  17. [17]
    [PDF] Ordinal numbers - Cornell Mathematics
    Moreover, every ordinal is equal to the set of its predecessor ordinals, and every ordinal has a successor ordinal α + 1 defined by (2). Finally, nonempty.
  18. [18]
    [PDF] Transfinite Induction - Penn Math
    Meanwhile, {1} is totally ordered by ∈, hence well ordered because it is finite, but is not a concrete ordinal because 1 := {∅} is an element but not a subset.
  19. [19]
    [PDF] Notes on the Axiom of Choice - ETH Zürich
    Jul 24, 2025 · The Well Ordering Principle. Every set admits a well ordering. Theorem 1. The Lemma of Zorn implies the Axiom of Choice. Proof. Assume the ...
  20. [20]
    [PDF] Zorn's lemma - Cornell Mathematics
    Zorn's lemma states: If a poset has every chain with an upper bound, then it has at least one maximal element.
  21. [21]
    [PDF] SET THEORY - Analysis
    This book provides an introduction to relative consistency proofs in axiomatic set theory, and is intended to be used as a text in beginning.
  22. [22]
    [PDF] A Formal Proof of the Independence of the Continuum Hypothesis
    Any uncountable family of finite sets has an uncountable subfamily forming a Δ-system. We say that a topological space has the CCC if every family of ...
  23. [23]
    [PDF] Notes on Introductory Point-Set Topology - Cornell Mathematics
    A collection B of open sets in a topological space X is called a basis for the topology if every open set in X is a union of sets in B. A topological space can ...
  24. [24]
    [PDF] An outline summary of basic point set topology - UChicago Math
    A topology on a set X is a set of subsets, called the open sets, which satisfies the following conditions. (i) The empty set ∅ and the set X are open. (ii) Any ...
  25. [25]
    [PDF] Nets and filters (are better than sequences)
    Nets are not bound by this restriction though, and in fact we can use the directed set structure of this collection of open sets itself to index an object that ...
  26. [26]
    [PDF] Schauder basis
    Dec 4, 2012 · Every Banach space with a Schauder basis has the approximation property. Basis problem. A theorem of Mazur asserts that every infinite ...
  27. [27]
    [PDF] how the concept of a net arrives from riemann integration in calculus
    A typical example for a directed set is the set P of all finite partitions of a closed interval [a, b]. Now, do you remember what a sequence of real numbers ...
  28. [28]
    When indexing set is empty, how come the union of an indexed ...
    Oct 3, 2016 · In the edge case when the list is empty the intersection is all of X . The whole set is the (multiplicative) identity for intersection just as ...The empty set as an Indexing set. [duplicate] - Math Stack ExchangeUnion and Intersection of Indexed Family - Math Stack ExchangeMore results from math.stackexchange.com
  29. [29]
    Cartesian Product of an Indexed Family by example
    Nov 13, 2019 · Mendelson's definition is good. An indexed family of sets is literally an indexing of a family of sets. In your example, your different ...
  30. [30]
    8.5 Indexed sets
    Indexed sets use indices to distinguish sets, where the indexing set is the set of indices and the sets are the indexed sets. Indices can be from any set.<|separator|>
  31. [31]
    [PDF] 1 Directed Sets 2 Nets
    Roughly speaking, nets are a generalization of sequences wherein the indexing set N is replaced by a directed set. As the name suggests, these sets have a ...
  32. [32]
    [PDF] SET THEORY CONCEPTS Cardinality (power) of a set = “number of ...
    Cardinality (power) of a set is the number of elements in a set. Sets A and B have the same cardinality if their elements can be paired.
  33. [33]
    [PDF] The Point-to-Set Principle and the Dimensions of Hamel Bases - arXiv
    Sep 21, 2023 · The rest of this proof establishes that B is a Hamel basis of R over Q and that dimH (B) = s. We first show that span(B) = R. For this, since we ...
  34. [34]
    [PDF] The cardinality of Hamel bases of Banach spaces
    By the axiom of choice, every vector space E over a field K possesses a Hamel base. A. Hamel base is hence a set H of vectors such that. (i) H spans E, i.e. E ...