Fact-checked by Grok 2 weeks ago

Partial function

In mathematics, a partial function from a set X to a set Y is a relation that assigns to each element in a subset D \subseteq X (known as the domain of definition) exactly one element in Y, while being undefined for elements outside D. This structure generalizes the notion of a total function, which requires definition on the entire set X, allowing partial functions to model situations where mappings are incomplete or conditional. Partial functions arise naturally in various mathematical contexts, such as set theory, where they are formalized as triples \langle A, G, B \rangle with G \subseteq A \times B being a functional graph (ensuring at most one output per input), and the effective domain \operatorname{dom}(f) = \{a \in A \mid \exists b \in B, (a, b) \in G\}. In category theory, they can be represented as spans A \leftarrow D \to B where D is a subset of A and the right leg is a total function, enabling composition via pullbacks in categories with such limits. Common examples include the square root function \sqrt{x} over the reals, undefined for x < 0, or division x / y undefined when y = 0. In computer science, partial functions are crucial for describing algorithms and computations that may halt or fail on certain inputs, such as Turing machines defining partial recursive functions that are computable but not necessarily total. They facilitate error handling and domain restrictions in programming, often modeled using types like the maybe monad to distinguish defined and undefined cases. The collection of all partial functions between sets forms a category \mathbf{Set}_*, which is equivalent to the category of sets with partial map classifiers, highlighting their role in foundational structures.

Definition and Fundamentals

Formal Definition

In set theory, a partial function from a set A to a set B is a binary relation f \subseteq A \times B such that for every a \in A, the set \{b \in B \mid (a, b) \in f\} has at most one element. This ensures that whenever the function is defined at a point, it assigns a unique output, but it allows for points in A where no output is specified. Such a partial function is commonly denoted by f: A \rightharpoonup B. The points in A where no corresponding element in B exists—meaning the set \{b \in B \mid (a, b) \in f\} is empty—are precisely where the partial function f is undefined. To denote this explicitly, various notations are used; for instance, f(a) = \uparrow indicates that f is undefined at a, where \uparrow symbolizes the absence of a value. A partial function can also be represented explicitly as an ordered triple (A, B, f), where A is the domain set, B is the codomain set, and f is the underlying binary relation satisfying the functional property. This formulation emphasizes the structural components and aligns with foundational approaches in set theory, such as those in Bourbaki's framework, adapted for partiality. Total functions arise as a special case of partial functions, where the relation is defined for every element of A.

Relation to Total Functions

A partial function f: A \rightharpoonup B between sets A and B is called a total function if its domain of definition equals the entire source set A. In this case, f assigns a unique element of B to every element of A, aligning with the standard notion of a function in set theory. Partial functions generalize total functions by permitting the domain of definition to be any subset of A, including the possibility of being undefined for some elements. Every total function qualifies as a partial function, but the converse does not hold, as partial functions naturally accommodate scenarios where the mapping is incomplete or undefined over parts of the source set. This generalization proves useful in contexts like computability and analysis, where operations may fail for certain inputs without invalidating the overall structure. To relate partial functions more closely to total ones, extension techniques allow enlarging the domain while preserving the original mapping. In set theory, partial functions can often be extended to total functions by assigning values from B to points outside the original domain, assuming B is nonempty. However, achieving maximal extensions—particularly in the poset of compatible extensions ordered by inclusion—relies on the axiom of choice or equivalent principles like Zorn's lemma to guarantee the existence of a maximal element, which would be total. For instance, in the case of partial choice functions, the axiom of choice ensures that any partial selection can be extended to a total choice function across a family of nonempty sets. In certain set theories, such as Zermelo-Fraenkel without the axiom of choice, there exist models featuring partial functions that cannot be extended to total functions without invoking additional axioms, particularly when extensions must satisfy specific structural constraints like well-definedness across infinite domains. These situations highlight the foundational role of choice principles in bridging partial and total mappings. This distinction underscores the flexibility of partial functions in modeling undefined behaviors while distinguishing them from fully defined total functions.

Properties and Operations

Domain, Codomain, and Graph

In mathematics, the domain of a partial function f: A \rightharpoonup B is defined as the set \dom(f) = \{ a \in A \mid \exists b \in B \text{ such that } (a, b) \in f \}, which represents the effective input set where the function is defined. This domain is a subset of the ambient set A, distinguishing partial functions from total functions, where \dom(f) = A. The codomain of a partial function f: A \rightharpoonup B is the fixed set B, which specifies the possible outputs, though not all elements of B need to be attained. In contrast, the image (or range) of f is the actual output set \im(f) = \{ b \in B \mid \exists a \in A \text{ such that } (a, b) \in f \}, which is a subset of the codomain. This distinction allows partial functions to model scenarios where outputs are restricted to certain values within B. The graph of a partial function f: A \rightharpoonup B is the relation G(f) = f \subseteq A \times B, consisting of all ordered pairs (a, b) where b = f(a) for a \in \dom(f). By definition, the graph ensures uniqueness: for each a \in \dom(f), there is at most one b \in B such that (a, b) \in f. This representation emphasizes the partial nature, as the graph may not include pairs for all elements of A. A partial function f: A \rightharpoonup B is a partial injection if it is injective on its domain, meaning that for all a_1, a_2 \in \dom(f), if f(a_1) = f(a_2), then a_1 = a_2. Similarly, f is a partial surjection if its image equals the codomain, i.e., \im(f) = B, ensuring every element of B is mapped to by some input in \dom(f). These properties extend the standard notions of injectivity and surjectivity to partial settings, focusing only on the defined portion of the function.

Composition and Restrictions

The composition of two partial functions f: A \rightharpoonup B and g: B \rightharpoonup C is the partial function g \circ f: A \rightharpoonup C defined by (g \circ f)(x) = g(f(x)) whenever x \in \dom(f) and f(x) \in \dom(g). The domain of the composition is given by \dom(g \circ f) = \{ x \in \dom(f) \mid f(x) \in \dom(g) \}, which is the subset of \dom(f) on which g is defined after applying f. Composition of partial functions is associative whenever it is defined: for partial functions f: A \rightharpoonup B, g: B \rightharpoonup C, and h: C \rightharpoonup D, the equality (h \circ g) \circ f = h \circ (g \circ f) holds on the common domain where all three compositions are defined, though the resulting partial function may not be total. The restriction of a partial function f: A \rightharpoonup B to a subset D \subseteq A is the partial function f|_D: A \rightharpoonup B such that \dom(f|_D) = D \cap \dom(f) and f|_D(x) = f(x) for all x \in \dom(f|_D). This operation preserves the codomain B and the values of f on the restricted domain. Two partial functions f: A \rightharpoonup B and g: A \rightharpoonup B (with the same input and output sets) are equal if they have the same domain and codomain, and f(x) = g(x) for all x \in \dom(f), or equivalently, if their graphs are identical.

Examples in Mathematics

Analytic Functions

In real analysis, partial functions arise naturally when mathematical operations are restricted to subsets of the real numbers where they are well-defined, ensuring the functions map to real values without invoking complex numbers or undefined expressions. These domain restrictions highlight the partial nature of the functions, distinguishing them from total functions defined everywhere on the reals. A classic example is the natural logarithm, denoted \ln: \mathbb{R} \rightharpoonup \mathbb{R}, which is defined only for positive real inputs, with domain \operatorname{dom}(\ln) = (0, \infty). It is undefined for non-positive reals because the integral representation \ln x = \int_1^x \frac{1}{t} \, dt requires the integrand to be defined over a path in the positive reals. Another fundamental partial function is the principal square root, \sqrt{\cdot}: \mathbb{R} \rightharpoonup \mathbb{R}, with domain \operatorname{dom}(\sqrt{\cdot}) = [0, \infty), as the square root of a negative number is not real-valued. This non-negativity requirement stems from the definition \sqrt{x} = y where y \geq 0 and y^2 = x for x \geq 0, preventing the function from being extended to all reals while preserving real outputs. Rational functions provide further illustrations of partiality through poles, where the denominator vanishes. For instance, the function f(x) = \frac{1}{x-1} has domain \mathbb{R} \setminus \{1\}, excluding x=1 to avoid division by zero, which introduces a vertical asymptote or pole at that point. Partial functions in real analysis are typically continuous on their domains, meaning that at every point in the domain, the limit of the function as the input approaches that point equals the function value there. This continuity holds for the natural logarithm on (0, \infty), the square root on [0, \infty), and rational functions on their domains excluding poles, as these restrictions avoid singularities within the defined sets. Such properties ensure that these functions behave predictably for analytical purposes, like differentiation or integration, within their restricted scopes.

Arithmetic Operations

In arithmetic on natural numbers, partial functions commonly arise because certain operations, defined recursively via the successor function, fail to yield a result in the natural numbers for some inputs, reflecting the inductive structure of the domain. These operations illustrate how partiality captures the inherent limitations of the natural numbers under basic arithmetic, where totality would require extending to broader structures like the integers. A primary example is subtraction, formally defined as the partial binary operation -: \mathbb{N} \times \mathbb{N} \rightharpoondown \mathbb{N}, where for m, n \in \mathbb{N}, m - n is defined if and only if m \geq n, in which case m - n \in \mathbb{N}. The domain of this function is thus the set of pairs \{(m,n) \mid m \geq n \subseteq \mathbb{N} \times \mathbb{N}\}. This partiality stems from the fact that subtracting a larger number from a smaller one does not produce a natural number result, making the operation undefined in those cases. In recursive function theory, subtraction can be extended to a total "proper subtraction" by mapping undefined cases to 0, but the strict version remains partial to preserve the natural number codomain. Division provides another illustration, particularly when restricted to exact division as the partial operation /: \mathbb{N} \times (\mathbb{N} \setminus \{0\}) \rightharpoondown \mathbb{Q}, defined only for pairs (m, n) where n divides m evenly (i.e., there exists k \in \mathbb{Q} such that m = k \cdot n with no remainder). In this context, integer division is partial not just due to the exclusion of division by zero but also because remainders render the result non-integer for many inputs, limiting the domain to cases of perfect divisibility. This focus on exact division highlights discrete failure points in natural number arithmetic, distinct from approximate methods that might yield total functions over reals. The predecessor function, pred: \mathbb{N} \rightharpoondown \mathbb{N}, exemplifies unary partiality, where pred(0) is undefined, but for any positive natural number n+1, pred(n+1) = n, inverting the successor operation on its image. This definition aligns with the Peano axioms, where every nonzero natural number has a unique predecessor, but zero does not, ensuring the function's domain excludes 0. Such partiality in arithmetic operations directly relates to the Peano axioms, where functions are often specified through inductive (recursive) definitions that may not exhaustively cover the entire domain, leading to undefined values for base cases or non-inductive inputs. For instance, the recursive construction of operations like subtraction and predecessor relies on the successor axiom but terminates undefined when induction cannot proceed, as seen in the absence of a predecessor for 0. This inductive origin underscores how partial functions naturally emerge in foundational arithmetic without invoking total extensions.

Logical and Computational Structures

In domain theory, partial functions are modeled using domains equipped with a bottom element \bot, representing the undefined value, which allows for the semantic representation of non-terminating computations. To incorporate \bot into a set P, the lifted set P_\bot is formed as P \cup \{\bot\}, where P is given the discrete partial order (antichain) and \bot is the least element below all others. This construction ensures that partial functions from a domain D to E can be viewed as total functions into the lifted domain E_\bot, preserving the partial order via graph inclusion. Partial recursive functions, also known as μ-recursive functions, are the computable partial functions from \mathbb{N}^k to \mathbb{N} for some k, generated as the smallest class containing the basic functions—zero function, successor, and projections—and closed under composition, primitive recursion, and unbounded minimization. Composition combines existing partial recursive functions to form new ones, such as f(\mathbf{x}) = g(h_1(\mathbf{x}), \dots, h_m(\mathbf{x})); primitive recursion defines functions with a base case and recursive step, extended partially when the recursion may not terminate; and minimization seeks the smallest n such that a partial recursive function g(n, \mathbf{x}) = 0, remaining undefined if no such n exists. This class aligns with the Church-Turing thesis, positing that partial recursive functions capture all effectively computable partial functions, equivalent to those computable by Turing machines. A canonical example of a partial recursive function is the halting function h: \mathbb{N} \rightharpoonup \{0,1\}, where h(p) = 1 if the program with index p halts on input p, and h(p) is undefined otherwise. This function is partial recursive because a semi-decision procedure exists: simulate the program until it halts (returning 1) or diverges (undefined), but it cannot be extended to a total recursive function, as that would solve the undecidable halting problem. The domain of h, consisting of indices of halting programs, is recursively enumerable but not recursive. In Scott domains, which are algebraic complete partial orders (cpos) with a basis of compact elements, partial functions are ordered pointwise and required to be monotone to ensure continuity with respect to directed suprema. Monotonicity means that if x \leq y in the domain, then f(x) \leq f(y) whenever both are defined, inducing a partial order on the space of partial functions by graph inclusion or pointwise comparison. This structure supports denotational semantics for recursive programs, where fixed points of monotone operators yield the least solutions corresponding to computable approximations.

Advanced Applications

Category Theory Perspective

In category theory, partial functions can be formalized using the concept of a partial map classifier, particularly within the framework of toposes. In a topos \mathcal{E}, the partial map classifier is an object \Sigma equipped with a morphism \top: 1 \to \Sigma (where $1 is the terminal object), such that for any objects A, B \in \mathcal{E}, partial maps A \rightharpoonup B correspond bijectively to total maps A \to \Sigma^B. This structure generalizes the subobject classifier \Omega by incorporating partiality through a subterminal object that represents truth values extended with an "undefined" element, allowing the domain of definition to be specified via characteristic maps. A monad T on a category \mathcal{C} serves as a partial map classifier if every partial map f: X \rightharpoonup Y is represented as a total map \hat{f}: X \to TY, where the domain is encoded in the structure of T. For instance, the maybe monad (or option monad) on the category of sets, defined as T X = 1 + X with unit \eta_X: X \to TX injecting elements and multiplication \mu_X: T(TX) \to TX collapsing nested options, models partial functions via its Kleisli category \mathcal{C}_T. In this Kleisli category, morphisms X \to Y are maps X \to TY in \mathcal{C}, representing partial functions, and composition is given by f ; g = \mu_Y \circ T g \circ f, which handles undefinedness through the monadic bind operation, ensuring that undefined inputs propagate appropriately. Partial functions can also be represented as spans in a category with pullbacks. A partial function f: A \rightharpoonup B is encoded as a span A \leftarrow D \to B, where D \subseteq A is the domain of definition, the left leg is the inclusion of the domain, and the right leg is the total function restricted to D. Two such spans are equivalent if related by an isomorphism over A and B, yielding the category \mathbf{Par}(\mathcal{C}) of partial maps, which extends \mathcal{C} by allowing undefinedness via the intermediate domain object. This span construction is functorial and preserves the relational nature of partiality. Within these frameworks, total functions—those defined on the entire domain—form a full subcategory of the category of partial functions. In restriction categories, which axiomatize partial maps via idempotent restrictions \overline{f} satisfying f = f \circ \overline{f} and related identities, total maps are precisely those with \overline{f} = \mathrm{id}, embedding the original category as a full subcategory. This distinction highlights strictness, where total maps compose strictly without introducing partiality, contrasting with the broader partial map category.

Algebraic Structures

In algebraic structures, partial functions play a crucial role by allowing operations and mappings that are defined only on subsets of the domain, enabling the study of incomplete or restricted systems while preserving local algebraic properties. This approach extends classical total functions to settings like magmas and quasigroups, where full totality may not hold or is unnecessary. Partial homomorphisms are mappings between algebraic structures that preserve operations only where both the operation and the map are defined on their arguments. In a quasigroup (Q, \cdot), a partial homomorphism f: S \to Q' from a subset S \subseteq Q \times Q satisfies f(a \cdot b) = f(a) \cdot' f(b) whenever a, b \in \mathrm{dom}(f) and a \cdot b \in S. For instance, in free Steiner quasigroups, which arise in combinatorial designs like Steiner triple systems, partial homomorphisms extend base maps recursively across levels of the free construction, ensuring preservation of the unique pairing property where defined. Such maps are essential in quasigroup theory for analyzing embeddings and endomorphisms, particularly in relation to partial Latin squares, where a partial Latin square of order n on symbols Q defines a partial quasigroup operation on Q via the filled entries, allowing homomorphisms to model symmetries or completions. In magmas, partial inverse functions address elements without full inverses by defining an inverse operation only where it exists and satisfies the required properties locally. A partial magma (M, \cdot) equips a set M with a partial binary operation, and a partial inverse for an element a \in M is a function i: D \to M (with D \subseteq M) such that a \cdot i(b) = b and i(b) \cdot a = b whenever defined, focusing on subsets where cancellativity or solvability holds. This construction is vital in universal algebra for studying non-associative structures like loops or general magmas, where total inverses fail, but partial ones facilitate quotient formations or extensions without assuming global properties. For example, in a partial magma derived from domain restrictions of a total magma, the partial inverse preserves local identities where the operation is closed. Free constructions in universal algebra often incorporate partial operations to generate minimal structures satisfying given identities on specified domains. The free partial algebra over a set X in a variety K is the quotient of the term algebra by the congruence induced by equations in K, with operations defined only on terms where the identities apply, yielding a universal object for homomorphisms from X. In universal algebra, partial operations generate such free structures by extending total algebras via coproducts, where the free extension adds generators with partial evaluations, as seen in equational theories for monoids or lattices. This allows modeling incomplete systems, like partially defined term operations in logic, while ensuring the structure is freely generated up to isomorphism. Partial functions induce congruences in algebraic structures by defining equivalence relations that respect partial operations through their domains. On a partial algebra (A, F), a congruence \theta is an equivalence relation on A such that if a \sim b (mod \theta) and the operation F is defined on (a_1, \dots, a_n), then it is defined on (b_1, \dots, b_n) and F(a_1, \dots, a_n) \sim F(b_1, \dots, b_n) (mod \theta). In partial lattices, for example, a partial function like a restricted join \vee induces a congruence E via the two-point extension, where the quotient inherits partial operations as intersections with the original domain, preserving lattice properties where defined. This framework enables quotient constructions for partial structures, analogous to total cases but restricted to compatible domains.

Geometric Constructions

In differential geometry, partial functions play a fundamental role in defining local coordinate systems on manifolds through the concept of charts. A chart on an n-dimensional manifold M is a partial homeomorphism \phi: U \dashrightarrow \mathbb{R}^n, where U is an open subset of M and \phi maps U bijectively onto an open subset of Euclidean space, providing local coordinates for points in U. This partiality arises because the domain is restricted to U, allowing the manifold's global structure to be pieced together from these local Euclidean approximations without requiring a single global coordinate system. An atlas on M is a collection of charts \{(U_i, \phi_i)\}_{i \in I} such that the union of the U_i covers M, enabling a consistent description of the manifold's topology and smoothness. The transition maps between charts, defined on the overlaps V_{ij} = \phi_i(U_i \cap U_j), are partial functions \phi_j \circ \phi_i^{-1}: V_{ij} \dashrightarrow \mathbb{R}^n that must be homeomorphisms to ensure compatibility; for a smooth atlas, these maps are required to be smooth diffeomorphisms, guaranteeing that the manifold inherits a differentiable structure from the Euclidean charts. This framework allows partial functions to glue local pieces into a coherent global object, as the restrictions on domains prevent inconsistencies across the manifold. Partial functions extend naturally to fiber bundles, where local trivializations serve as charts that reveal the bundle's structure over open subsets of the base. For a fiber bundle (E, p, B, F) with projection p: E \to B and fiber F, a local trivialization over an open set U \subset B is a partial homeomorphism \psi: p^{-1}(U) \dashrightarrow U \times F that commutes with the projection, identifying the preimage p^{-1}(U) with the product space locally while preserving the fiber over each point in U. These trivializations, like manifold charts, are partial because they are defined only over local opens, and their compatibility on overlaps is ensured by transition functions that are partial bundle isomorphisms, maintaining the bundle's topological or smooth structure. The compatibility condition for these partial functions is crucial for defining smooth or topological structures on manifolds and bundles. On manifold overlaps U_i \cap U_j, the partial transition maps must align to induce a well-defined Jacobian or derivative, ensuring that vector fields, differentials, and other geometric objects are independent of chart choice; similarly, in fiber bundles, compatible partial sections or projections over trivialization overlaps preserve the bundle's global twisting, such as in the case of non-trivial line bundles. This reliance on partial functions underscores their role in constructing geometric objects where global definitions are impossible or impractical.

Partial Functions in Function Spaces

Spaces of Partial Functions

The set of all partial functions from a set A to a set B, commonly denoted A \rightharpoonup B or \mathrm{PF}(A, B), comprises all right-unique binary relations R \subseteq A \times B, where for each a \in A, there is at most one b \in B such that (a, b) \in R. This collection forms a subset of the power set \mathcal{P}(A \times B), as each partial function corresponds to its graph, a functional subset of the Cartesian product. For finite sets with |A| = m and |B| = n, the cardinality of A \rightharpoonup B equals (n + 1)^m, reflecting the n + 1 choices per element of A (undefined or mapped to one of n elements in B). In the infinite case, under cardinal arithmetic, this cardinality is (|B| + 1)^{|A|}. When restricted to partial functions from a set A to itself, the collection A \rightharpoonup A becomes a monoid under the operation of composition, where for partial functions f, g: A \rightharpoonup A, the composition f \circ g is defined at x \in A by (f \circ g)(x) = f(g(x)) whenever g(x) is defined and f is defined at g(x); otherwise, it is undefined. The identity element of this monoid is the total identity function \mathrm{id}_A: A \to A, which maps every element to itself. This structure, known as the monoid of partial transformations or the full partial transformation monoid on A, plays a significant role in semigroup theory and combinatorics on words. Pointwise operations can be defined on spaces of partial functions when B carries additional algebraic structure, such as that of a ring or vector space, provided the domains align appropriately. For instance, if f, g: A \rightharpoonup B share the same domain D \subseteq A and B is a ring, the pointwise sum is the partial function h: A \rightharpoonup B with domain D given by h(x) = f(x) + g(x) for x \in D, and similarly for multiplication h(x) = f(x) \cdot g(x). More generally, the domain of such an operation is the intersection of the domains of f and g, enabling the construction of algebraic structures like partial rings or modules over these function spaces. These operations preserve the partial nature and facilitate applications in constructive analysis and measure theory. In topological contexts, spaces of partial continuous functions between topological spaces X and Y—such as those with closed domains—can be endowed with topologies like the generalized compact-open topology, where convergence is evaluated subbasis elements involving compact subsets of X and open sets in Y. Under suitable conditions, such as X being locally compact Hausdorff and Y completely regular, certain subsets of these function spaces exhibit compactness, analogous to the compact-open topology on total continuous functions; for example, equicontinuous families of partial continuous functions on compact domains may form compact sets. These properties underpin developments in generalized topology and proximity spaces.

Topological and Metric Properties

In topological spaces, a partial function f: X \dashv Y between topological spaces (X, \tau_X) and (Y, \tau_Y) is defined to be continuous if its restriction to the domain \dom(f) \subseteq X (equipped with the subspace topology induced from \tau_X) is a continuous map from \dom(f) to Y. Equivalently, f is continuous if for every open set V \in \tau_Y, the preimage f^{-1}(V) = \{ x \in \dom(f) \mid f(x) \in V \} is open in the subspace topology on \dom(f). This relative openness condition ensures that continuity is preserved under the natural embedding of the partial function into total functions on the domain. An alternative characterization arises when viewing the partial function as a relation or via its graph. In certain settings, such as when X and Y are Hausdorff and \dom(f) is compact, the graph G(f) = \{ (x, f(x)) \mid x \in \dom(f) \} \subseteq X \times Y (with the product topology) is closed if and only if f is continuous and \dom(f) is closed in X. This closed-graph property is particularly useful in settings where partial functions are treated as relations, as the continuity of the relation aligns with the closedness of the graph under the assumption of a closed domain. To endow spaces of partial functions with a metric structure, one common approach extends the codomain to include an "undefined" value, often represented using the extended real line [-\infty, \infty] where undefined points map to \infty (or -\infty). For partial functions f, g: X \dashv \mathbb{R} on a metric space (X, d_X), a metric on the space of such functions can be defined by first extending them to total functions \tilde{f}, \tilde{g}: X \to [-\infty, \infty] with \tilde{f}(x) = \infty if x \notin \dom(f), and then using d(f, g) = \sup_{x \in X} | \tilde{f}(x) - \tilde{g}(x) |, where the absolute value is adapted for extended reals (e.g., via a suitable metric on [-\infty, \infty], such as d_\infty(a, b) = |e^a - e^b| for finite a, b with adjustments for infinities). This supremum metric over the entire domain captures differences on overlapping defined regions while penalizing domain mismatches through infinite distances. Alternatively, for functions with potentially differing domains, the distance can be restricted to the intersection \dom(f) \cap \dom(g), yielding d(f, g) = \sup_{x \in \dom(f) \cap \dom(g)} |f(x) - g(x)| if the intersection is nonempty, with infinite distance otherwise; this induces a pseudometric on the space of partial functions, useful for uniform convergence on common domains. In measure theory, a partial function f: X \dashv Y between measurable spaces (X, \mathcal{A}_X, \mu) and (Y, \mathcal{A}_Y) is measurable if \dom(f) \in \mathcal{A}_X (i.e., the domain is measurable) and the restriction f|_{\dom(f)}: \dom(f) \to Y is a measurable function with respect to the subspace sigma-algebra on \dom(f). This ensures that preimages under f of measurable sets in Y remain measurable in X, with the undefined points outside \dom(f) not affecting the property. Equivalently, f can be viewed as a total measurable function to the disjoint union Y \sqcup \{\bot\}, where \bot denotes undefined, and the preimage of \{\bot\} is the complement of \dom(f), which must then be measurable. This definition aligns with standard extensions in computable measure theory, where partial measurability preserves integrals over domains.

References

  1. [1]
    partial function in nLab
    ### Summary of Partial Functions from nLab
  2. [2]
    Partial Function - an overview | ScienceDirect Topics
    A partial function is a relation where each input has a unique output, defined on a subset of a set, with a domain of elements that have corresponding outputs.
  3. [3]
    Partial Function -- from Wolfram MathWorld
    A partial function is a function that is not total.Missing: definition | Show results with:definition
  4. [4]
    [PDF] Chapter 2 Relations, Functions, Partial Functions
    A function takes inputs to outputs, with each input having a unique output. A partial function is not defined for all input values.
  5. [5]
    [PDF] Set Theory for Computer Science
    A partial function from X to Y is a relation f ⊆ X × Y for which. ∀x, y, y0 ... (a, b) = f(a0,b0). We obtain. (fA(a),fB(b)) = (fA(a0),fB(b0)). Whence ...
  6. [6]
    Java Software Development with Event B
    ... function, and we write f: A → B. If f is defined for some values of A, we say that f is a partial function, and we write f: A ⇸ B. If f is a function such ...
  7. [7]
    [PDF] Glossary of mathematical notation and terminology
    'f(x)↑' means 'there is no y ∈ Y with (x, y) ∈ f' (and is read 'f(x) is undefined'). An n-ary partial function from X to Y is just a partial function from ...<|control11|><|separator|>
  8. [8]
    tag removed - Definition of Function - MathOverflow
    Jul 3, 2010 · When we use the Bourbaki definition of function as a triple (domain, codomain, graph), then two functions are usually defined to be equal iff ...
  9. [9]
    partial function in nLab
    Jan 28, 2024 · 1. Idea. A partial function f : A → B f: A \to B is like a function from A to B except that f ( x ) may not be defined for every element x of A ...
  10. [10]
    Partial functions and total functions - Applied Mathematics Consulting
    Dec 6, 2021 · A partial function is a relation in which for every a in A there is at most one pair (a, b) in the relation. A “multi-valued function” is ...
  11. [11]
    [PDF] THE AXIOM OF CHOICE
    A any function f with domain J ⊆ I such that f(i) ∈ Ai for all i ∈ J. The set F of partial choice functions on A can be partially ordered by inclusion ...
  12. [12]
    On extensions of partial functions - ScienceDirect.com
    Nov 1, 2007 · The problem of extending partial functions is considered from the general viewpoint. Some aspects of this problem are illustrated by ...
  13. [13]
    [PDF] Partial Functions Exercises - Cimat
    A program designed to evaluate a function may not produce the correct value of the function for all elements in the domain of this function.<|separator|>
  14. [14]
    [PDF] Review of Math
    ○ If f is a partial injection, we write: f ∈ S ↣̸ T ... Can an integer array “int[] a” be modelled/formalized as a partial surjection (i.e., a ∈ Z ↠̸ Z)?.
  15. [15]
    [PDF] Event B: Sets, Relations, Functions, ArithmeticMany slides borrowed ...
    Injection: if f (x) = f (y), then x = y. Partial injection. S 7↣ T. Total injection. S ↣ T. Surjection: f ∈ S ↔ T, ran(f ) = T. Partial surjection. S 7↠ T.Missing: mathematics | Show results with:mathematics
  16. [16]
    Practical Foundations of Mathematics - Paul Taylor
    The poset of partial functions Agreement of total functions is all or nothing, but two partial functions (Definition 1.3.1(b)) may agree on the intersection ...
  17. [17]
    9.2 The natural logarithm
    1 The natural logarithm ln(x) is an antiderivative of 1/x, given by lnx=∫x11tdt. ◻. Figure 9.2.1 gives a geometric interpretation of ...
  18. [18]
    [PDF] Rational Functions - Math
    The implied domain of a rational function is the set of all real numbers except for the roots of the denominator. That's because it doesn't make sense to divide ...
  19. [19]
    [PDF] Continuous Functions - UC Davis Math
    A function f : A → R is continuous on a set B ⊂ A if it is continuous at every point in B, and continuous if it is continuous at every point of its domain A.
  20. [20]
    Recursive Functions - Stanford Encyclopedia of Philosophy
    Apr 23, 2020 · The recursive functions are a class of functions on the natural numbers studied in computability theory, a branch of contemporary ...
  21. [21]
    Numbers | SpringerLink
    9.1.2.3 Subtraction. It is worthwhile to start by noting that subtraction on the natural numbers is a partial function, in that n−m is not defined when m is ...
  22. [22]
    [PDF] The Euclidean algorithm on the natural numbers N = f0,1,...g can be ...
    A j= E = w () den(A,E) = w. Partiality introduces some complications which deserve notice. For exam- ple, if we view subtraction as a partial function on N, ...
  23. [23]
    P-RAM vs. RP-RAM - ScienceDirect
    Aug 15, 2017 · ∸ P-RAM [ + , [ ∸ ] , [ × ] , ← , [ → ] , Bool ] = PSPACE . Here, “/”, known as exact division, is a weaker form of division.
  24. [24]
    [PDF] chapter 1: the peano axioms - Summer 2019 Edition
    We define the predecessor function π : N+ → N as follows: given n ∈ N+ we define πn to be the unique predecessor of n. 9. We do not use ! by itself to mean “ ...
  25. [25]
    [PDF] 22.1 Representability of Functions in a Formal Theory
    Apr 9, 2009 · • The predecessor function p inverts the successor function ... Theorem: All µ-recursive functions are representable in Peano Arithmetic.
  26. [26]
    [PDF] Domain Theory
    • The set [X*Y ] of partial functions between sets X and Y ordered by graph ... lifted by mapping the new bottom element of D⊥ to the new bottom element of E⊥.
  27. [27]
    [PDF] rec.1 The Halting Problem
    The halting problem in general is the problem of deciding, given the specifica- tion e (e.g., program) of a computable function and a number n, whether the.
  28. [28]
    [PDF] Domain Theory: An Introduction - Rice University
    Exercise 2.13: Show also that partial order of monotonic functions mapping D0 to E0 (using the pointwise ordering) is isomorphic to the partial order of ...
  29. [29]
    [PDF] Complete Partial Orders, PCF, and Control
    Definition 3. Let (D,≤) and (D0,≤0) be partial orders. A function f : D → D0 is called monotonic if, for all x,y ∈ D, if x ≤ y, then f(x) ≤0 f(y).
  30. [30]
    [PDF] Partiality and Container Monads
    A monad T on a category C is a partial map classifier if every partial map f : X*Y in C, to be thought of as a total map from a certain subset of X into Y , is ...
  31. [31]
    [PDF] Kleisli Categories
    Apr 7, 2018 · The category SetOption corresponds to sets with partial functions. The category SetM corresponds to sets with enumerably-non-deterministic ...<|control11|><|separator|>
  32. [32]
    [PDF] a 2-categories companion - UChicago Math
    (j) Par consists of sets and partial functions. A partial function from X to Y is a diagram X D → Y in Set, where D is the domain of definition of the partial ...
  33. [33]
    [PDF] Restriction categories I: categories of partial maps
    A more basic example of a restriction category is the category of sets and partial functions.
  34. [34]
    [PDF] A Course in Universal Algebra - Department of Mathematics
    Free algebras are discussed in great detail—we use them to derive the existence of simple algebras, the rules of equational logic, and the important Mal'cev.
  35. [35]
    [PDF] Endomorphisms of free Steiner quasigroups - arXiv
    Feb 24, 2025 · We start with f0 = f. Recall that A = S0 and that a · b 6∈ S0 if a, b ∈ S0 and, therefore, f0 is a partial homomorphism ...
  36. [36]
    [PDF] loop transversal codes - Jonathan DH Smith's
    A quasigroup Q ... Suppose that. T carries a loop structure (T,∗,0) given by an isomorphism (4.5) such that s : (T,+) →. (An−k,+) is a partial homomorphism.
  37. [37]
    [PDF] Partially-Static Data as Free Extension of Algebras
    We present a foundational view of partially-static data structures as free extensions of algebras for suitable equational theories, i.e. the coproduct of an ...
  38. [38]
    None
    ### Summary of Congruences on Partial Lattices and Partial Functions
  39. [39]
    Fibre Bundles | SpringerLink
    Download chapter PDF. Preliminaries on Homotopy Theory. Preliminaries on Homotopy Theory. Dale Husemoller. Pages 1-8. The General Theory of Fibre Bundles. Front ...
  40. [40]
    Number of partial functions between two sets - Math Stack Exchange
    Feb 1, 2013 · Number of partial functions between two sets ... The number of partial functions A→B is (1+|B|)|A|. Now either this is a well-known formula, or I ...<|control11|><|separator|>
  41. [41]
    Ehresmann semigroups whose categories are EI and their ...
    Nov 1, 2021 · ... s ⁎ m ) ⁎ and by the right. The monoid of partial functions. Recall that PT n denotes the monoid of all partial functions on the set ...
  42. [42]
    [PDF] Families of Sets in Constructive Measure Theory - arXiv
    Jul 8, 2022 · ... partial functions ... (pointwise) operations on F(X). Let. (A, iA,f) and (B,iB,g) be real-valued partial functions and ∗ be any of the ...
  43. [43]
    topology.partial - mathlib3 docs - Lean community
    In this file we prove properties of filter.ptendsto etc in topological spaces. We also introduce pcontinuous , a version of continuous for partially defined ...
  44. [44]
    ct.category theory - Continuous relations? - MathOverflow
    Aug 22, 2014 · ... A⇸B defines two adjunctions R∗⊣R−1 and R−1⊣R!, where R∗( ... NOTATION (set theory):. A relation (from X to Y) is an ordered triple (X ...
  45. [45]
    Partial Metrics, Valuations, and Domain Theory - ResearchGate
    Aug 6, 2025 · In this paper we develop some connections between the partial metrics of Matthews and the topological aspects of domain theory.
  46. [46]
    [PDF] On S-Finite Measures and Kernels - arXiv
    Oct 3, 2018 · We define a measurable partial function X⇀Y as a measurable function X → Y + {⊥}. A kernel from X to Y is a map X ×ΣY → [0, ∞] which is ...
  47. [47]
    [PDF] REPRESENTATIONS OF MEASURABLE SETS IN COMPUTABLE ...
    Aug 19, 2014 · Measure theory is a fundament of modern analysis. In particular, computable measure theory is a fundament of computable analysis.
  48. [48]
    [PDF] The fixed-point property for represented spaces - Hal-Inria
    Definition 1.1. A represented space X has the fixed-point property if every continuous multifunction h : X ⇒ X has a fixed-point. This ...<|control11|><|separator|>
  49. [49]
    Metric fixed point theory and partial impredicativity - Journals
    Apr 10, 2023 · ... fixed point theorem for both Baire and Borel functions is ... A lower semi-continuous partial function V : X → R is coded by a ...