Fact-checked by Grok 2 weeks ago

Completeness

Completeness is a property or concept that appears in various fields of mathematics, logic, computer science, and other disciplines, often referring to the idea that a system, structure, or theory fully captures or satisfies certain criteria without gaps or omissions.) In mathematical logic, it specifically denotes a fundamental property of formal deductive systems, ensuring that every semantically valid formula—true in every possible model—is also syntactically provable from the system's axioms and rules of inference. This bridges semantics (meaning and truth in models) and syntax (formal proofs). A cornerstone result is Gödel's completeness theorem (1929), which states that in classical first-order logic, a sentence \phi is provable from a set of axioms \Gamma if and only if it is true in every model of \Gamma. Equivalently, every consistent set of first-order sentences has a model. The theorem, proved by Kurt Gödel in his doctoral dissertation, applies to classical first-order predicate logic (with or without equality) and resolves a question from David Hilbert and Wilhelm Ackermann's 1928 Grundzüge der theoretischen Logik. Modern proofs are non-constructive, relying on the compactness theorem, Zorn's lemma, and the axiom of choice to construct models from consistent theories. Completeness is often paired with soundness, where every provable formula is semantically valid; together, they ensure the system proves exactly the valid formulas. While first-order logic is complete, higher-order logics are not. This contrasts with Gödel's incompleteness theorems (1931), which show that sufficiently powerful formal systems for arithmetic cannot be both consistent and complete in deciding all sentences. Beyond , completeness has diverse meanings. In , it includes order-theoretic completeness (e.g., Dedekind-complete orders like the reals), metric completeness (Cauchy-complete spaces), and set-theoretic axioms like the completeness axiom for the reals. In , describes systems capable of simulating any , while identifies hard decision problems. Variations appear in (conceptual completeness), (harmonic completeness), statistics (statistical completeness), and (semantic completeness). These notions, while related in emphasizing wholeness, are formalized differently across domains, as explored in subsequent sections.

Philosophical and Conceptual Foundations

General Definition

Completeness, derived from the Latin completus, the past participle of complēre meaning "to fill up" or "to fulfill," refers to the state of being whole, entire, or without any missing parts. In philosophical terms, it embodies the idea of a system, structure, or body of knowledge that lacks gaps, deficiencies, or unresolved elements, achieving a form of wholeness or sufficiency that renders it self-contained and integral. This notion underscores the pursuit of totality, where all necessary components are present to ensure coherence and fulfillment, free from fragmentation or incompleteness. In everyday , completeness manifests intuitively through like a "complete ," which includes balanced nutrients to satisfy nutritional needs without omissions, or a "complete collection," such as a set encompassing all variants from a given , illustrating the absence of voids in an otherwise assembled whole. These examples highlight how completeness denotes not just quantity but qualitative sufficiency, where the entity meets its intended purpose or scope fully. This formalization of intuitive ideas about wholeness appears in disciplines like , where completeness ensures that all valid implications are captured within a system.

Historical Development

The concept of completeness originated in , particularly in 's metaphysics, where it was intertwined with teleological notions of purpose and fulfillment. Aristotle introduced the term entelecheia, denoting the realization or completeness of a thing's potential, such that an entity achieves its inherent end or when it fully actualizes its form. For instance, in living beings, completeness arises from the actualization of their natural capacities, marking the transition from potency to full being. This teleological framework positioned completeness not as mere wholeness but as the attainment of a purposeful , influencing subsequent philosophical inquiries into and . In medieval , thinkers like adapted and Christianized Aristotelian ideas, linking completeness to divine perfection and the hierarchical order of creation. viewed as the ultimate source of all perfections, where completeness signifies integrity and wholeness derived from divine essence, with creatures participating imperfectly in this divine completeness through their forms. In his synthesis of faith and reason, human beatitude—complete happiness—requires union with , the perfectly good being, underscoring completeness as a theological ideal beyond earthly attainment. This development elevated completeness from a metaphysical principle to a marker of spiritual fulfillment, bridging pagan philosophy with Christian doctrine in works like the . Epistemological interpretations of completeness emerged notably in Immanuel Kant's late 18th-century philosophy, where completeness emerged as an ideal of reason that remains unattainable in empirical knowledge. In the , Kant described the "Ideal of Pure Reason" as a systematic unity encompassing the totality of conditions, yet one that reason pursues as a regulative principle without ever fully realizing it in experience. This ideal represents completeness as the complete determination of objects through reason's categories, but its transcendence of sensory limits renders it an asymptotic goal rather than a concrete achievement. Concurrently, in mathematics, notions of completeness began to formalize in analysis, addressing gaps in real numbers through constructions like Dedekind cuts to ensure foundational rigor. The transition to 20th-century formal logic marked completeness as a central objective in , which sought to axiomatize all of with a complete, consistent, and decidable system. envisioned completeness as the property of a formal system capturing all mathematical truths within its proofs, aiming to secure against paradoxes through finitary methods. This ambition culminated in pivotal challenges from Kurt Gödel's 1930 announcement and 1931 publication of the incompleteness theorems, which demonstrated that sufficiently powerful formal systems cannot be both complete and consistent, thereby reshaping discussions on the limits of formal completeness. Paradoxically, Gödel's results intensified philosophical and logical explorations of completeness by highlighting its unattainability in comprehensive axiomatic frameworks.

Formal Logic

Semantic Completeness

In formal logic, semantic completeness refers to the property of a logical system where every formula that is true in all models of the system is provable within that system. This ensures that the deductive apparatus captures all semantic validities, bridging the gap between truth in every interpretation and syntactic derivability. established this property for predicate logic in his 1929 dissertation, published in 1930. The states that for any set of sentences \Gamma and sentence \phi, \Gamma \models \phi if and only if \Gamma \vdash \phi, where \models denotes semantic entailment (truth in all models) and \vdash denotes syntactic provability. A high-level proof sketch involves showing that any consistent set of sentences \Gamma (one without ) can be extended to a maximally consistent set, which then corresponds to a model via the and a canonical construction assigning truth values based on provability; this model satisfies \Gamma, ensuring that semantic consistency aligns with syntactic consistency. Semantic completeness complements , which guarantees that every provable is true in all models (provability implies truth). Together, and completeness establish equivalence between provability and validity, meaning the system neither proves falsehoods nor misses truths. In propositional logic, semantic completeness holds via truth-table semantics: every , such as P \lor \neg P, is provable using axioms like and substitution. This extends to , where the applies to quantified s; for instance, the valid sentence \forall x (P(x) \to P(x)) is provable, and any consistent in the syntactic sense has a model interpreting predicates and domains accordingly. However, semantic completeness fails for higher-order logics, where quantification over predicates and functions allows expressing categoricity (e.g., uniquely characterizing the natural numbers up to ), precluding a complete effective axiomatization that captures all standard models without incompleteness. Similarly, in non-classical logics like , which rejects the and uses instead of classical models, the classical notion of semantic completeness does not hold, as provability requires constructive evidence rather than mere truth preservation across all possible worlds.

Syntactic Completeness

In formal logic, a is syntactically complete if for every \phi in its , either \phi or \neg \phi is provable from the theory's axioms. This property, also called deductive completeness, ensures that the theory decides every declarative statement expressible in its language, providing a maximal extension without . It emphasizes the exhaustiveness of the theory's deductive power for its specific content, distinct from the semantic completeness of the underlying proof system. Proof calculi such as Hilbert-style systems, which use a finite set of axiom schemas (e.g., for and ) with , and natural deduction systems, with introduction and elimination rules for connectives, achieve the syntactic derivations needed for semantic completeness in . Similarly, Gerhard Gentzen's (1934–1935), using sequents \Gamma \vdash \Delta and structural rules with cut-elimination, provides a syntactic framework that realizes completeness for classical and intuitionistic logics, advancing . In , the syntactic completeness of standard proof systems aligns with semantic completeness via Gödel's 1930 completeness theorem, which proves that every semantically valid formula is syntactically provable; however, this alignment breaks in incomplete theories such as Peano arithmetic, where Gödel's incompleteness results demonstrate the existence of undecidable sentences neither provable nor refutable within the system.

Mathematics

Order-Theoretic Completeness

In order theory, a partially ordered set (poset) is complete if every subset thereof has both a least upper bound, known as the supremum, and a greatest lower bound, known as the infimum. This property ensures that the poset possesses all possible joins and meets for arbitrary collections of elements, extending the finite case of lattices. A complete poset is thus inherently a lattice, as finite subsets also admit such bounds. A is a specific type of complete poset where the order structure supports these operations universally, often arising in algebraic and combinatorial contexts. Prominent examples include the power set of any set, ordered by , where the supremum of a collection of subsets is their and the infimum is their . Similarly, the collection of all open sets in a forms a under , with arbitrary unions serving as suprema and arbitrary intersections as infima, though the latter may require under infinite operations in certain topologies. For linear orders, a related notion is Dedekind completeness, where every nonempty subset that is bounded above possesses a supremum; this avoids gaps in the order, as formalized via Dedekind cuts. The real numbers , equipped with the standard order, exemplify a Dedekind-complete , distinguishing them from , which contain "gaps" like the cut defining \sqrt{2}. Order-theoretic completeness finds significant application in domain theory, where complete partial orders (cpos)—posets complete with respect to directed suprema—model the semantics of recursive computations in programming languages. These structures enable by interpreting programs as elements in , with fixed points of continuous functions representing solutions to recursive definitions via least fixed-point theorems. A key result in this framework is , which asserts that if every chain in a poset has an upper bound, then the poset contains a maximal element; this lemma is equivalent to the and underpins existence proofs in complete lattices, such as the construction of bases or maximal ideals.

Metric and Topological Completeness

In , completeness is defined in terms of , which capture the notion of that "should" converge based on their internal distances. A (X, d) is complete if every in X converges to a point within X. Here, a \{x_n\} is if, for every \epsilon > 0, there exists N \in \mathbb{N} such that d(x_m, x_n) < \epsilon for all m, n > N. This property ensures that the space is "closed" under limits of such sequences, making it suitable for where convergence is essential. Incomplete spaces, by contrast, contain that fail to converge internally, highlighting gaps in the structure. A classic example of a complete metric space is the set of real numbers \mathbb{R} equipped with the standard Euclidean metric d(x, y) = |x - y|, where every Cauchy sequence converges to a real number. In contrast, the rational numbers \mathbb{Q} with the same metric form an incomplete space; for instance, the sequence x_n = \sum_{k=1}^n \frac{1}{k^2} is Cauchy in \mathbb{Q} but converges to \frac{\pi^2}{6}, which is irrational and thus outside \mathbb{Q}. To remedy incompleteness, any metric space can be completed by constructing its Cauchy completion, which embeds the original space densely into a complete metric space. This completion is formed by taking the set of equivalence classes of Cauchy sequences in the original space, where two sequences \{x_n\} and \{y_n\} are equivalent if \lim_{n \to \infty} d(x_n, y_n) = 0, and equipping this quotient set with a metric defined by d([\ \{x_n\}\ ], [\ \{y_n\}\ ]) = \lim_{n \to \infty} d(x_n, y_n). The resulting space is complete, and the original space is isometric to a dense subspace. The provides a profound characterization of s, asserting that any complete metric space is non-meager (or of second category) in itself, meaning it cannot be expressed as a countable union of nowhere dense sets. A set is nowhere dense if its has empty interior. This theorem implies that complete metric spaces have a "rich" topological structure, resisting decomposition into "small" sets. In applications to function spaces, such as the space of continuous functions on a compact set with the sup , the underpins key results like the (Banach-Steinhaus theorem), which states that a bounded family of bounded linear operators on a is uniformly bounded. These spaces, being complete metric spaces, leverage the theorem to ensure the existence of continuous selections or to prove openness of mappings between them. The concept of completeness extends beyond metric spaces to the more general framework of uniform spaces, where replace to define . A is complete if every Cauchy (or ) converges in the space; a is Cauchy if it is contained in every after some point. This generalizes completeness, as every induces a uniform structure, but uniform spaces allow for non-metrizable topologies, such as those on product spaces. Complete uniform spaces inherit desirable properties like those in the metric case, including applications in and analysis. A particularly significant class of complete metric spaces is that of Polish spaces, defined as separable complete metric spaces (i.e., those admitting a countable dense subset). Polish spaces are central to descriptive , where they serve as the ambient spaces for studying Borel and analytic sets, enabling the classification of definable subsets under Polish topology and measurability. Examples include \mathbb{R}^n, the \ell^2, and the \mathbb{N}^\mathbb{N}, all of which support powerful regularity results for sets and functions.

Set-Theoretic Completeness

In Zermelo-Fraenkel with the (ZFC), the is not stated as a primitive axiom but emerges as a theorem derived from the axiom and the axiom schema of separation. The axiom asserts that for any set A, there exists its \mathcal{P}(A), the set of all subsets of A, while separation allows the formation of subsets defined by formulas within a given set. Together, these axioms enable the construction of the real numbers \mathbb{R} as the set of all s of the rational numbers \mathbb{Q}, where a is a of \mathbb{Q} into two non-empty subsets L and R such that all elements of L are less than all elements of R, and L has no greatest element. This construction ensures that \mathbb{R} satisfies the : every non-empty subset of \mathbb{R} that is bounded above has a least upper bound in \mathbb{R}. introduced this construction in to rigorously define the real numbers and establish their completeness, motivated by the need to fill gaps in that prevent certain geometric intuitions, such as the of lines. By relying solely on the of sets via separation and , Dedekind proved that the resulting \mathbb{R} is a complete , where completeness guarantees the of suprema without invoking limits or sequences directly. This set-theoretic approach demonstrates that completeness follows from the foundational axioms of set , providing a uniform basis for . Historically, Dedekind's work framed completeness as the "" for the reals, distinguishing it from earlier informal notions of in and , such as those in or Cauchy, which lacked rigorous handling of irrationals. This shift resolved debates in 19th-century about whether should be an or a ; Dedekind argued it as an axiomatic property of the reals derived from set partitions, avoiding reliance on completeness alone and emphasizing the absence of gaps in the . Paul Cohen's 1963 forcing technique reveals the flexibility of completeness-like properties in by constructing models of ZFC where certain structural features vary, such as the cardinality of the continuum relative to the reals' completeness. Forcing extends a ground model by adding a generic filter through a partial order, producing independence results like the negation of the while preserving ZFC axioms, including the completeness of \mathbb{R}. This shows that while core completeness is provable in ZFC, broader properties involving infinite sets and their bounds can differ across models. A central tool in forcing is the complete Boolean algebra, which generalizes partial orders to ensure that arbitrary suprema and infima exist for sets of conditions, facilitating the construction of generic extensions. In this framework, dense subsets of the Boolean algebra correspond to forcing conditions that "force" statements to hold, guaranteeing that the generic filter intersects every dense set and thus maintains a form of generic completeness in the extended model, where new sets are added without violating foundational completeness.

Theoretical Computer Science

Turing Completeness

A is Turing-complete if it can simulate any and thus compute any partial , representing the standard notion of universal computability. This property captures the expressive power required to perform all effectively calculable procedures, as formalized in the Church-Turing thesis, which posits that any function computable by an is computable by a . The concept originated in Alan Turing's 1936 paper "On Computable Numbers, with an Application to the ," where he introduced the as an abstract model of mechanical computation to address Hilbert's . A consists of an infinite tape divided into cells, each holding a from a ; a read-write head that scans one cell at a time; a of states, including an initial state; and a finite table of transitions that specify, for each state- pair, the symbol to write, the direction to move the head (left, right, or none), and the next state. Computation proceeds deterministically by repeatedly applying these transitions until reaching a halting state or continuing indefinitely, enabling the machine to compute functions on natural numbers or simulate human calculation processes. The Church-Turing thesis links Turing machines to effective computability, asserting their equivalence to other models like Alonzo Church's and Stephen Kleene's mu-recursive functions, all of which define the same class of partial recursive functions. For instance, , introduced by Church in 1936, achieves through term abstraction and application, allowing simulation of any via encoding of states and tape operations. Similarly, mu-recursive functions, built from basic functions (zero, successor, projection) via composition, primitive recursion, and minimization, are equivalent to computations, as proven by Kleene in 1936 and formalized in Turing's 1937 analysis. While delineates the boundaries of standard computation, concepts of explore models purportedly exceeding these limits, such as accelerating Turing machines that solve undecidable problems like the in finite time; however, such models remain theoretical and do not alter the thesis for physically realizable effective computation.

Completeness

In , a is complete for a such as if it belongs to the class and every other problem in the class can be reduced to it using a polynomial-time , meaning there exists a polynomial-time that maps yes-instances of the original problem to yes-instances of the complete problem and no-instances to no-instances. This notion captures the hardest problems within the class, as solving a complete problem efficiently would allow efficient solutions to all problems in the class via the reductions. Completeness under polynomial-time reductions provides a way to study the relative difficulty of problems without solving them directly, highlighting structural properties of . The foundational result establishing completeness in this context is the Cook-Levin theorem, which proves that the (SAT)—the task of determining whether there exists a truth assignment that satisfies a given in —is NP-complete. Introduced by in 1971, the theorem demonstrates that SAT is in NP (via a nondeterministic polynomial-time verifier) and that any problem in NP reduces to SAT in polynomial time, using a construction that simulates the nondeterministic computation of the original problem as a . This breakthrough, independently discovered by around the same time, provided the first example of an NP-complete problem and sparked the systematic study of completeness across combinatorial problems. Subsequent work expanded the roster of NP-complete problems, illustrating the breadth of this hardness. For instance, the 3-SAT problem, a restricted version of SAT where each contains exactly three literals, is NP-complete, as shown by Richard Karp in 1972 through a from general SAT. Similarly, the —deciding whether a given undirected contains a of specified size—is NP-complete, again via Karp's reductions from SAT, underscoring how completeness permeates and optimization. These examples demonstrate that is not isolated to logical satisfiability but extends to diverse domains, implying that efficient algorithms for one would yield them for the others. The implications of for the P versus NP question are profound: if any NP-complete problem admits a polynomial-time , then P = NP, as all problems in NP would then be solvable in polynomial time through the . This equivalence holds because the preserve polynomial-time solvability. To explore relativized versions of this question, separations reveal limitations of proof techniques; in 1975, Theodore Baker, John Gill, and Robert Solovay constructed relative to which P = NP and others where P ≠ NP, showing that non-relativizing methods are necessary to resolve the problem. Such results emphasize that completeness notions can vary relative to additional computational power provided by , aiding in understanding barriers to proving P ≠ NP.

Applications in Other Disciplines

Music Theory

In music theory, completeness refers to a compositional structure, such as a , that exhausts all distinct pitches within a given without repetition, ensuring structural balance in atonal music. This property is foundational to , where a complete set—typically comprising all twelve pitches—serves as the basis for deriving melodic, harmonic, and textural elements, promoting equality among notes and avoiding tonal hierarchy. The concept of completeness emerged in the early as a response to the increasing dissonance in late , particularly following Richard Wagner's innovations, which blurred distinctions between and pushed toward chromatic saturation. , seeking to extend this trend while maintaining organizational rigor, developed the in the , formalizing completeness as a core principle for atonal composition. In this method, a composer creates a —a specific ordering of the twelve pitches—used as the source material; all subsequent material derives from transformations of this complete row to ensure every pitch appears with equal frequency across the piece. Key examples of completeness in serial music include the prime form of the , its inversion (upside-down version), (backward reading), and retrograde-inversion (backward and upside-down), each preserving the full set of pitches. These forms allow composers to generate varied yet unified sections, as seen in Schoenberg's own works like the Suite for , Op. 25 (1923), where the row's completeness underpins motivic development. Schoenberg's students Alban Berg and Anton Webern applied completeness extensively in their compositions; Berg integrated it into more lyrical, tonal-inflected structures in operas like Wozzeck (1925), while Webern achieved stark economy through fragmented presentations of complete rows in pieces such as the Six Bagatelles for String Quartet, Op. 9 (1913, pre-twelve-tone but anticipatory). The principle later extended beyond pitch to other parameters, including rhythm and dynamics, in total serialism pioneered by composers like Olivier Messiaen and Milton Babbitt, where ordered series ensure completeness across durations, intensities, and timbres for multidimensional coherence.

Statistics and Probability

In statistics, completeness is a property of a statistic that ensures it captures all relevant information about the parameter without redundancy in unbiased estimation. A statistic T is said to be complete for a family of distributions parameterized by \theta if, for every measurable function g, the condition E_\theta [g(T)] = 0 for all \theta in the parameter space implies that P_\theta (g(T) = 0) = 1 for all \theta. This property prevents the existence of non-trivial unbiased estimators of zero based on T, thereby guaranteeing the uniqueness of certain estimators when combined with sufficiency. When a is both and complete, it plays a central role in improving through the . The theorem asserts that if \hat{\theta} is an unbiased of a \theta and T is a , then the E[\hat{\theta} \mid T] is also unbiased for \theta and has variance less than or equal to that of \hat{\theta}. If T is additionally complete, this Rao–Blackwellized is the unique uniformly (UMVUE) of \theta, as established by the Lehmann–Scheffé theorem. This combination minimizes information loss in data reduction for parameter estimation. The Lehmann–Scheffé theorem formalizes the connection between completeness, sufficiency, and optimal estimation: if T is a complete and W is any unbiased of a g(\theta), then E[W \mid T] is the unique unbiased of g(\theta) that is a of T and achieves the minimum variance among all unbiased estimators. This result ensures that complete sufficient statistics provide the foundation for deriving optimal point estimators in models. A prominent class of distributions where complete minimal sufficient statistics arise is the . For independent and identically distributed observations from a full-rank with density f(x \mid \theta) = h(x) c(\theta) \exp(\eta(\theta) \cdot t(x)), the natural sufficient statistic T = \sum t(X_i) is complete under mild conditions on the parameter space, such as it containing an . For example, in the distribution N(\mu, \sigma^2) with both parameters unknown, the pair consisting of the and sample variance forms a complete minimal . Similarly, for the with \lambda, the sample sum is a complete . Basu's theorem further highlights the utility of completeness by linking it to independence properties in inference. The theorem states that if T is a complete sufficient statistic and V is an ancillary statistic (whose distribution does not depend on \theta), then T and V are independent. This independence simplifies the construction of tests and estimators by separating parameter-dependent and parameter-free components of the data. In applications, complete sufficient statistics facilitate the development of optimal procedures in hypothesis testing and confidence intervals. For hypothesis testing, the Lehmann–Scheffé theorem implies that uniformly most powerful unbiased tests, when they exist, can be based on complete sufficient statistics, ensuring minimal power loss. In confidence intervals, UMVUEs derived via Rao–Blackwellization on complete sufficient statistics yield intervals with the shortest expected length among unbiased alternatives, providing precise while minimizing information loss from the data.

Linguistics and Semantics

In linguistics, a key property of a grammar, aligned with descriptive adequacy in , is its ability to generate all and only the well-formed of a given , ensuring full coverage of its structural and interpretive possibilities without over- or under-generation. The grammar must accurately capture the native speaker's intuitions about and meaning for the language's set. Historically, distinguished between —the idealized knowledge of a language's rules that enables speakers to produce and understand all possible sentences—and , the actual, imperfect use of language influenced by factors like memory or context. Competence embodies this property by positing an innate, fully specified system that accounts for the language's expressive range, as articulated in Chomsky's foundational work on . This idealization draws from formal systems in , providing a tie to precise meaning coverage in . Montague grammar advances this through formal semantics, using lambda calculus to map syntactic structures compositionally onto meanings, thereby ensuring completeness in interpretation for specified language fragments. Richard Montague's approach treats natural language fragments as formal languages with fully specified syntax and semantics, where every syntactic combination yields a corresponding semantic value via lambda abstraction and application. For instance, in analyzing quantified noun phrases like "every man runs," the grammar assigns types (e.g., individual or truth-value denotations) and uses lambda terms to guarantee that all well-formed expressions receive exhaustive, non-ambiguous interpretations without gaps. Categorial grammars exemplify this property by classifying linguistic expressions as complete (e.g., basic categories like nouns) or incomplete (function categories like verbs seeking arguments), enabling the derivation of all valid sentence structures through function-argument application. In frameworks like the Lambek calculus, such grammars achieve coverage for fragments by proving that every permissible combination aligns with semantic compositionality, as seen in analyses of scope ambiguities in sentences like "a chases every ." These systems extend Montague's method, providing bidirectional that covers both surface order and underlying relations without extraneous derivations. In , this focus on complete coverage informs algorithms designed to process utterances, ensuring that all valid syntactic-semantic analyses are recovered for input strings matching the . For example, chart parsers in categorial frameworks maintain completeness by exhaustively exploring derivations while avoiding , which is crucial for applications like or systems handling full, context-independent sentences. This emphasis on full expressive capacity supports robust by prioritizing grammars that mirror the full range of human language use.

References

  1. [1]
    Gödel's Completeness Theorem -- from Wolfram MathWorld
    Gödel's Completeness Theorem: If T is a set of axioms in a first-order language, and a statement p holds for any structure M satisfying T, then p can be ...
  2. [2]
    [PDF] Gödel and the metamathematical tradition - andrew.cmu.ed
    In his 1929 doctoral dissertation, Gödel proved the completeness theorem for first-order logic, clarifying a relation- ship between semantic and syntactic ...
  3. [3]
    [PDF] g¨odel's completeness and incompleteness theorems
    Theorem 3.1. Any consistent set of formulas cannot be complete, in particular, for every consistent set of formulas there is a statement that is neither ...
  4. [4]
    Complete - Etymology, Origin & Meaning
    Originating in the late 14th century from Old French and Latin completus, "complete" means perfect, finished, or to make something whole or accomplished.
  5. [5]
    COMPLETENESS definition | Cambridge English Dictionary
    the quality of being whole or perfect and having nothing missing. For the sake of completeness, I should also mention two other minor developments.
  6. [6]
  7. [7]
    Fundamentality - Stanford Encyclopedia of Philosophy
    Jul 21, 2018 · So, one could distinguish between absolute completeness and restricted completeness. It is easy to see how the restricted notion of ...
  8. [8]
    [PDF] Philosophy of Logic The Concept of Completeness
    Aug 23, 2012 · 1929. “¨Uber die Vollständigkeit des Logikkalküls,” Gödel's doc- toral thesis at the University of Vienna, translation by S. Bauer-Mengelberg.
  9. [9]
    [PDF] Teleology in Aristotle's Practical Philosophy - PhilArchive
    The 'complete' or 'perfect'. (teleia) actualisation of the form is the. 'goal' (telos) of living beings, which have an internal 'drive' (hormê) to actualise.
  10. [10]
    Aristotle's Ontology of Change on JSTOR
    Aristotle's concepts of potency (dunamis), activity (energeia), and being-complete or fulfillment (entelecheia) are among the greatest contributions of change ...
  11. [11]
    [PDF] Final Causality in the Thought of Thomas Aquinas - Purdue e-Pubs
    And because perfection denotes completeness, integrity, or wholeness, it ... their goodness, and each of their perfections—and this we call a God.231. In ...
  12. [12]
    Thomas Aquinas: Moral Philosophy
    Aquinas believes that we can never achieve complete or final happiness in this life. For him, final happiness consists in beatitude, or supernatural union with ...
  13. [13]
    Thomas Aquinas - Stanford Encyclopedia of Philosophy
    Dec 7, 2022 · Thomas Aquinas (ca. 1225–1274). The greatest figure of thirteenth-century Europe in the two preeminent sciences of the era, philosophy and theology.Life and Works · Cognitive Theory · Will and Freedom · EthicsMissing: completeness | Show results with:completeness
  14. [14]
    Kant's Account of Reason - Stanford Encyclopedia of Philosophy
    Sep 12, 2008 · Kant's philosophy focuses on the power and limits of reason. Two questions are central. In his theoretical philosophy, Kant asks whether reasoning can give us ...Theoretical reason: reason's... · Practical reason: morality and...Missing: unattainable | Show results with:unattainable
  15. [15]
    Critique of Pure Reason (Ideal) - Stanford University
    They contain a certain completeness to which no possible empirical know- ledge ever attains. In them reason aims only at a systematic unity, to which it seeks ...
  16. [16]
    [PDF] Completeness and Categoricity: - 19th Century Axiomatics to
    Jul 10, 2001 · Quite distinct but equally important are several notions of completeness for mathematical theories T. Logicians today are accustomed to talking.
  17. [17]
    Hilbert's Program - Stanford Encyclopedia of Philosophy
    Jul 31, 2003 · It calls for a formalization of all of mathematics in axiomatic form, together with a proof that this axiomatization of mathematics is consistent.
  18. [18]
    Gödel's Incompleteness Theorems
    Nov 11, 2013 · Gödel's two incompleteness theorems are among the most important results in modern logic, and have deep implications for various issues.
  19. [19]
    How Gödel's Proof Works - Quanta Magazine
    Jul 14, 2020 · His incompleteness theorems destroyed the search for a mathematical theory of everything. Nearly a century later, we're still coming to ...
  20. [20]
    Classical Logic - Stanford Encyclopedia of Philosophy
    Sep 16, 2000 · A logic consists of a formal or informal language together with a deductive system and/or a model-theoretic semantics.
  21. [21]
    Die Vollständigkeit der Axiome des logischen Funktionenkalküls
    Apr 30, 2005 · Die Vollständigkeit der Axiome des logischen Funktionenkalküls ... Article PDF. Download to read the full article text. Use our pre-submission ...
  22. [22]
    The Model-Theoretic Argument and the Completeness Theorem
    The Löwenheim-Skolem Theorem states that if aa set of FOL sentences has an infinite model, it has a model whose domain is countably infinite.
  23. [23]
    Kurt Gödel - Stanford Encyclopedia of Philosophy
    Feb 13, 2007 · The Second Incompleteness Theorem shows that the consistency of arithmetic cannot be proved in arithmetic itself. Thus Gödel's theorems ...
  24. [24]
    Second-order and Higher-order Logic
    Aug 1, 2019 · Later in §9.1 we shall modify the semantics so that a Completeness Theorem can be proved.
  25. [25]
    Intuitionistic Logic - Stanford Encyclopedia of Philosophy
    Sep 1, 1999 · Intuitionistic logic encompasses the general principles of logical reasoning which have been abstracted by logicians from intuitionistic mathematics.4. Basic Proof Theory · 5. Basic Semantics · 6. Additional Topics And...<|control11|><|separator|>
  26. [26]
    Proof Theory - Stanford Encyclopedia of Philosophy
    Aug 13, 2018 · Hilbert's approach raised fascinating metamathematical questions—from semantic completeness through mechanical decidability to syntactic ...
  27. [27]
    [PDF] CHAPTER 5 Hilbert Proof Systems: Completeness of Classical ...
    The Hilbert proof systems are systems based on a language with implication and contain a Modus Ponens rule as a rule of inference. They are usually called ...
  28. [28]
    A note on the completeness proof for natural deduction - Project Euclid
    The completeness of the system is now easily shown. Let a be some arbitrary formula which cannot be proved. Then the assumption -a is undischargeable in the one ...
  29. [29]
    [PDF] completeness in modal logic - UChicago Math
    Notice that weak completeness is an exact converse of our soundness definition, while strong completeness applies to whole sets of formulas at once.
  30. [30]
    (PDF) The completeness theorem of Gödel - ResearchGate
    Aug 9, 2025 · In this part, we present the completeness theorem of first order logic proved first by Gödel in 1929. We give a sketch of the proof due to ...
  31. [31]
    [PDF] arXiv:1907.08050v2 [math.CO] 20 Oct 2024
    Oct 20, 2024 · A complete lattice is a poset L such that every subset X of L has a meet (greatest lower bound) 기 X and join (least upper bound) V X. The ...
  32. [32]
    Chapter 2 Ordered Sets and Complete Lattices - profs.scienze.univr.it
    In particular, the image of a (complete) lattice under an order-isomorphism is a (complete) lattice. Further, an order-isomorphism preserves and ⊥ when these ...
  33. [33]
    [PDF] Chapter 5. Lattices, closure operators, and Galois connections.
    If T is a topological space, show that the open sets in T, partially ordered by inclusion, form a complete lattice. Describe the meet and join operations ( ...
  34. [34]
    Dedekind completion in nLab
    Aug 19, 2025 · While Dedekind completeness was traditionally described in the context of the real numbers, it can be stated for any strict linear order, ...
  35. [35]
    [PDF] Denotational Semantics - University of Cambridge
    Domain theory makes use of partially ordered sets satisfying certain completeness properties. The definition of a partial order is recalled on Slide 15. D ...
  36. [36]
    [PDF] Zorn's lemma and some applications - Keith Conrad
    Zorn's lemma is not intuitive, but it is logically equivalent to more intuitively plausible statements in set theory like the Axiom of Choice, which says every ...
  37. [37]
    Complete Metric Space -- from Wolfram MathWorld
    A complete metric space is a metric space in which every Cauchy sequence is convergent. Examples include the real numbers with the usual metric.
  38. [38]
    [PDF] Cauchy's Construction of R - UCSD Math
    The sequence 3,3.1,3.14,3.142,3.1416,3.14159,... is Cauchy but does not converge to a rational number; hence, there must be some irrational number to which it ...
  39. [39]
    [PDF] The Completion of a Metric Space - Northwestern Math Department
    Given a metric space (X, d), the completion of X with respect to d is the set X of classes of Cauchy sequences in X with the metric ¯d defined above. Example.
  40. [40]
    Baire category theorem - PlanetMath
    Mar 22, 2013 · The Baire category theorem Mathworld Planetmath is often stated as “no non-empty complete metric space is of first category”.
  41. [41]
    245B, Notes 9: The Baire category theorem and its Banach space ...
    Feb 1, 2009 · The Baire category theorem is equivalent to the claim that in a complete metric space, the countable intersection of open dense sets remain dense.
  42. [42]
    [PDF] Completeness of uniform spaces
    Defi: A uniform space is complete if overy Cauchy net converges. Props: A closed subspace of a complete space is complete. A complete subspace of a Hdf uniform ...<|separator|>
  43. [43]
    [PDF] Introduction to descriptive set theory - Mathematics and Statistics
    Oct 16, 2025 · Definition 1.1. A topological space is called Polish if it is separable and completely metrizable. (i.e. admits a complete compatible metric).
  44. [44]
    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.
  45. [45]
    [PDF] Project Gutenberg's Essays on the Theory of Numbers, by Richard ...
    ON CONTINUITY AND IRRATIONAL NUMBERS, and ON THE NATURE AND. MEANING OF NUMBERS. By R. Dedekind. From the German by W. W.. Beman. Pages, 115.
  46. [46]
    THE DISCOVERY OF FORCING - jstor
    Oct 29, 2001 · In my own work on CH, I never was able to successfully analyze proofs as a combinatorial "game" played with symbols on paper. Therefore, I begin ...<|separator|>
  47. [47]
    The Church-Turing Thesis (Stanford Encyclopedia of Philosophy)
    Jan 8, 1997 · The Church-Turing thesis concerns the concept of an effective or systematic or mechanical method, as used in logic, mathematics and computer science.The Case for the Church... · The Church-Turing Thesis and...
  48. [48]
    Turing machines - Stanford Encyclopedia of Philosophy
    Sep 24, 2018 · Independently of Turing, Emil Post (Post 1936) and Alonzo Church (Church 1936) gave a different but logically equivalent formulation (see Sec.
  49. [49]
  50. [50]
    Recursive Functions - Stanford Encyclopedia of Philosophy
    Apr 23, 2020 · The latter notion came first historically and was introduced by Turing (1939) and in an equivalent form by Kleene (1943). However it was ...The General Recursive... · The Origins of Recursive... · The Partial Recursive...
  51. [51]
    [PDF] Completeness - Computational Complexity: A Modern Approach
    Jan 8, 2007 · The following theorem provides us with our first natural NP-complete problems: Theorem 2.10 (Cook-Levin Theorem [?, ?]) Let SAT be the ...
  52. [52]
    [PDF] Cook 1971 - Department of Computer Science, University of Toronto
    Theorem 1: If a set S of strings is accepted by some nondeterministic Turing machine within polynomial time, then S is P-reducible to { DNF tautologies}.
  53. [53]
    [PDF] REDUCIBILITY AMONG COMBINATORIAL PROBLEMS
    REDUCIBILITY AMONG COMBINATORIAL PROBLEMS. +. Richard M. Karp. University of California at Berkeley. Abstract: A large class of computational problems involve ...
  54. [54]
    Relativizations of the P = ? N P Question - SIAM Publications Library
    N P Question. Authors: Theodore Baker, John Gill, and Robert SolovayAuthors Info & Affiliations ... On the P vs NP problem over reals with integer oracle.
  55. [55]
    34.1 Twelve-Tone Technique
    All twelve notes of the chromatic scale must occur · No note can be repeated in the series until the other 11 notes of the chromatic scale have occurred ( ...
  56. [56]
    Serialism – Twentieth- and Twenty-First-Century Music
    The Austrian composer and theorist Arnold Schoenberg invented twelve-tone music in the 1920s to provide a sense of logic analogous to that of harmonic function ...
  57. [57]
    [PDF] The Mathematics of Twelve Tone Music
    Having created a Tone Row and utilized it to establish a Twelve Tone Matrix, all the tools for composing a piece of music according to the twelve tone technique.
  58. [58]
    [PDF] the development of schoenberg's twelve-tone technique from opus ...
    Chamber Symphony will be shown to contain certain elements that may be analyzed in a more or less "traditional" man ner, it retains many ideas and elements ...
  59. [59]
    [PDF] Twelve-tone Serialism: Exploring the Works of Anton Webern
    To look at the birth of atonality, however, we cannot look at Wagner, even though he began the process of emancipating the dissonance. We must turn our ...<|control11|><|separator|>
  60. [60]
    [PDF] POST–WORLD WAR II CONTEXTS - UCI Music Department
    ... rhythm, dynamics, timbre, articula- tion, and texture. This effort, referred to variously as Integral Serialism, Total. Serialism, and General Serialism ...
  61. [61]
    Completeness, Ancillarity, and Basu's Theorem - Stat 210a
    A statistic T ( X ) is complete for a family of distributions if no nontrivial function of T can have expectation zero for every distribution in the family.
  62. [62]
    [PDF] 5. Completeness and sufficiency 5.1. Complete statistics. Definition ...
    5. Completeness and sufficiency 5.1. Complete statistics. Definition 5.1. A statistic T is called complete if Eg(T) = 0 for all.
  63. [63]
    [PDF] STA732 Statistical Inference - Lecture 05: Rao-Blackwell Theorem
    • Rao-Blackwell theorem allows us to improve an estimator based on sufficient statistics T. • If unbiased estimator exists, complete sufficient statistics T.
  64. [64]
    [PDF] 18 The Exponential Family and Statistical Applications
    This is in the form of a one parameter Exponential family with the natural sufficient statistic. T(X) = T(X1,X2) = X1 + X2. Example 18.5. (Gamma Distribution).
  65. [65]
    [PDF] Sufficiency, Minimal Sufficiency and the Exponential Family of ...
    If T is a complete sufficient statistic for the family of distributions P ... Poisson, and normal classes are all exponential families. 4.
  66. [66]
    Exponential Families - UC Berkeley Statistics
    Aug 29, 2023 · Example (Poisson, continued): As we showed above, in the Poisson exponential family the sufficient statistic is T ( X ) = X , the natural ...
  67. [67]
    On Statistics Independent of a Complete Sufficient Statistic - jstor
    Definition 1: Any measurable transformation T of the sample space (S, J?) onto a measurable space (v_?, ?8) is called a statistic. The probability measures on &.<|control11|><|separator|>
  68. [68]
    [PDF] Chapter 7 Point Estimation - Arizona Math
    Use the Rao-Blackwell Theorem to construct an unbiased estimator d∗(T(X)) of g(θ), where T(X) is a complete and sufficient statistic. 3. Construct d∗(T(X)) ...
  69. [69]
    [PDF] ASPECTS OF THE THEORY OF SYNTAX - Colin Phillips |
    generative grammar have been undertaken only quite recently. They seem to offer extremely rich and varied possibilities for study in all aspects of grammar.
  70. [70]
    What is the difference between syntactic and semantic completeness?
    Jul 9, 2017 · A theory T is semantically complete if, in every interpretation, every true formula ϕ is provable. I do not understand how syntactic ...Meaning of completeness in logic - Mathematics Stack ExchangeWhat is the difference between Completeness and Soundness in ...More results from math.stackexchange.com
  71. [71]
    [PDF] R. Montagues "English as a Formal Language" - Sascha Brawer
    [Montague 1970] Montague, Richard: English as a Formal Language. In: Visentini et al. [Hrsg.]: Linguaggi nella Società e nella Tecnica. Milano: Edizioni di ...
  72. [72]
    Categorial Grammar - an overview | ScienceDirect Topics
    'Categorial grammar' refers to approaches to grammar in which expressions are classified according to their completeness or incompleteness.
  73. [73]
    [PDF] The Logic of Categorial Grammars: Lecture Notes - HAL Inria
    May 19, 2006 · Abstract: These lecture notes present categorial grammars as deductive systems, and include detailed proofs of their main properties.
  74. [74]
    [PDF] Computational Linguistics: Parsing
    Oct 5, 2018 · ▷ Completeness: A parsing algorithm is complete if it returns every possible analysis of every string, given the grammar provided ...