Fact-checked by Grok 2 weeks ago

Consistency

Consistency is the of , , or uniformity among parts or features of a whole, often implying , reliability, or the absence of contradictions across related elements or over time. In logical and philosophical contexts, consistency refers to the property of a set of statements or beliefs where no contradictions arise, meaning it is possible for all elements to be true simultaneously without logical . For instance, a is consistent if there exists a model or under which all its axioms hold true. In , consistency encompasses theories and principles related to cognitive and behavioral alignment, where individuals are motivated to maintain congruence among their attitudes, beliefs, and actions to reduce psychological discomfort. Cognitive consistency theories, such as and , highlight how people strive for internal harmony in their mental states, viewing inconsistencies as sources of tension that drive or rationalization. This drive extends to , where consistency is seen as the absence of contradictions in , serving as a foundational for reliable ethical conduct. In and databases, consistency denotes the correctness and uniformity of states, ensuring that transactions adhere to predefined rules and that all replicas or copies of remain identical across systems. For example, in distributed systems, guarantees that all reads reflect the most recent write, while allows temporary discrepancies that resolve over time. In physics, consistency often manifests as dimensional consistency, a fundamental check where equations must balance in units (e.g., , time, ) on both sides to ensure physical validity. These varied applications underscore consistency's role as a core concept in ensuring reliability and across disciplines.

General Concepts

Definition and Historical Context

In formal logic, consistency is a core property of a that ensures its axioms do not lead to contradictory derivations. A is syntactically consistent if no formula φ exists such that both φ and its ¬φ can be derived as theorems from the axioms using the system's rules of . This definition emphasizes the syntactic aspect, focusing on what can be proved within the system itself without reference to external . In contrast, semantic consistency requires that the axioms have a model—an where all axioms are true—thus precluding any in that structure. These notions are equivalent in sound and complete systems like , where provability aligns with truth in all models. The historical development of consistency arose from foundational crises in during the late 19th and early 20th centuries, particularly in response to paradoxes that exposed flaws in informal reasoning about sets and . , discovered in 1901, exemplified this by showing that the assumption of a set containing all sets that do not contain themselves leads to a self-referential contradiction, undermining . This and similar paradoxes, such as those in Cantor's transfinite numbers, highlighted the need for precise axiomatic systems to avoid such inconsistencies and restore rigor to mathematical foundations. David Hilbert's program, outlined in the 1920s, formalized the quest for consistency by advocating finitary proofs—relying only on concrete, finite methods—to demonstrate that axiomatic systems for are free of contradictions, thereby justifying the use of ideal infinite processes. In the 1930s, this effort advanced through contributions from key figures: , who collaborated with Hilbert on foundational texts exploring consistency in logical systems, and , whose investigations into provability and models refined the criteria for consistent formalization. These developments established consistency as a central concern in , influencing subsequent axiomatic frameworks.

Types of Consistency

In mathematical logic, consistency can be categorized along two primary axes: syntactic versus semantic, and absolute versus relative. Syntactic consistency, also known as proof-theoretic consistency, holds for a formal theory if there is no derivation of a contradiction, such as both a formula ϕ and its negation ¬ϕ, within the system's axioms and rules of inference. Semantic consistency, or model-theoretic consistency, requires that the theory has a model—an interpretation in which all axioms are true—ensuring satisfiability without contradiction. For first-order logic, Gödel's completeness theorem establishes the equivalence between these notions: a theory is syntactically consistent if and only if it is semantically consistent. Absolute consistency refers to the outright provability of a theory's consistency within itself, without reliance on external assumptions. However, Gödel's second incompleteness theorem demonstrates that any consistent formal system capable of expressing basic arithmetic, such as Peano arithmetic, cannot prove its own consistency; thus, absolute consistency remains undecidable within sufficiently powerful systems. In contrast, relative consistency proofs establish that the consistency of one theory T implies the consistency of another theory S, often by embedding S into T or constructing a model of S from a model of T. These proofs can be syntactic, showing that a contradiction in S would yield one in T, or semantic, demonstrating model existence for S given T's model. A classic example is Gödel's relative consistency proof that the axioms of choice and the continuum hypothesis are consistent relative to Zermelo-Fraenkel set theory (ZF), via the constructible universe model. Similarly, the consistency of ZF is relative to ZF plus the existence of a strongly inaccessible cardinal, as the cumulative hierarchy up to such a cardinal yields a model of ZF. Beyond these distinctions, specialized notions of consistency arise in specific theories, such as ω-consistency in arithmetic systems like Peano arithmetic. A theory is ω-consistent if it is consistent and does not prove both an existential statement ∃x φ(x) and the negations ¬φ(n) for every standard natural number n, preventing the system from "falsely witnessing" its own existential claims through infinite descent. This stronger condition than mere consistency was used in Gödel's original proof of the first incompleteness theorem but was later weakened to simple consistency by Rosser's theorem. In arithmetic, ω-consistency ensures that the theory avoids proving false Σ₁ sentences, providing a safeguard against certain pathological proofs while remaining relevant for analyzing provability predicates.

Consistency in First-Order Logic

Notation and Syntax

In first-order logic, the standard notation includes a set of logical connectives to combine formulas: conjunction (\wedge), disjunction (\vee), negation (\neg), material implication (\to), and material equivalence (\leftrightarrow). Quantifiers are denoted by the universal quantifier (\forall) and the existential quantifier (\exists), which bind variables in their scope. The language also features variables (typically lowercase letters like x, y, z), constant symbols (lowercase letters like a, b, c or domain-specific names), function symbols (e.g., f, g with specified arity), and predicate symbols (e.g., P, Q with arity, representing relations). Equality is often included as a binary predicate =. The syntax defines terms recursively: a term is either a variable, a constant symbol, or a function symbol applied to terms of appropriate arity (e.g., f(t_1, \dots, t_n) where each t_i is a term). Atomic formulas consist of a predicate symbol applied to terms (e.g., P(t_1, \dots, t_n)) or equality between terms (t_1 = t_2). Well-formed formulas (wffs) are built recursively from atomic formulas using connectives and quantifiers: if \phi and \psi are wffs, then so are \neg \phi, (\phi \wedge \psi), (\phi \vee \psi), (\phi \to \psi), (\phi \leftrightarrow \psi); additionally, if x is a , then (\forall x \, \phi) and (\exists x \, \phi) are wffs. Parentheses ensure unambiguous parsing, with standard precedence rules (e.g., \neg binds tightest, followed by quantifiers, then \wedge and \vee, then \to and \leftrightarrow). Sentences are wffs with no free variables. Variables in a wff are either (occurring outside any quantifier ) or bound (within the of a matching \forall or \exists). For instance, in \forall x (P(x) \to Q(x)), x is bound, while in P(x) \to Q(y), both x and y are ; substituting a t for a x in a wff \phi yields \phi[t/x], provided no capture occurs (i.e., variables in t are renamed if necessary to avoid by quantifiers in \phi). This notation provides the formal basis for expressing and analyzing consistency in theories.

Formal Definition

In first-order logic, a T, consisting of a set of , is consistent if there exists no \phi such that both T \vdash \phi and T \vdash \neg \phi, where \vdash denotes syntactic derivability within the theory. Equivalently, T is consistent if it does not derive a , such as \bot or \phi \land \neg \phi for some \phi. A T is inconsistent if it is not consistent, meaning there exists some \phi with T \vdash \phi and T \vdash \neg \phi. In such a case, from the principles of , T derives every possible sentence \psi; that is, if T is inconsistent, then for any \psi, T \vdash \psi. This property, known as ex falso quodlibet (from falsehood, anything follows), renders an inconsistent trivial and incapable of distinguishing truths. Maximal consistent sets provide a useful extension of this notion: a consistent set T is maximal if no larger consistent set properly contains it, meaning T is closed under logical consequence and for every sentence \phi, exactly one of \phi or \neg \phi belongs to T. Every consistent theory can be extended to a maximal consistent set.

Basic Properties and Results

In first-order logic, a theory T is consistent if and only if it does not prove the falsum \bot, meaning there exists no deduction from T to a contradiction. This syntactic characterization ensures that the theory avoids deriving both a formula \phi and its negation \neg \phi for any \phi. A fundamental property is the , which establishes that for any theory T, \phi, and \psi, T \cup \{\phi\} \vdash \psi if and only if T \vdash \phi \to \psi. This equivalence facilitates reasoning about hypothetical assumptions and preserves the structure of proofs across theory extensions. From the follows the preservation of consistency under non-refuting extensions: if T is consistent and T \not\vdash \neg \phi, then T \cup \{\phi\} is also consistent. Equivalently, T \cup \{\phi\} proves \bot only if T already derives \neg \phi, ensuring that adding a compatible with T maintains the theory's . Every consistent extends to a complete consistent , often termed a prime , via applied to the of consistent extensions of T. In this construction, the maximal element is complete, deciding every formula in the language, and remains consistent. Such extensions underpin results like Henkin's theorem, which constructs models from consistent .

Henkin's Theorem

Henkin's theorem, a cornerstone of , asserts that every consistent set of sentences in has a model, thereby establishing the semantic completeness of the proof system. Formally, if \Gamma is a consistent set of closed formulas in a first-order language, then there exists a structure \mathcal{M} such that \mathcal{M} \models \phi for all \phi \in \Gamma. This result forms a key part of the broader completeness theorem, linking syntactic consistency directly to the existence of a satisfying interpretation. The proof relies on the Henkin construction, which extends a given consistent to a maximal consistent set enriched with es for existential quantifiers. Beginning with a consistent set \Gamma, one enumerates all closed sentences in the and iteratively extends \Gamma by adding sentences while preserving consistency, yielding a maximal consistent set T_0. To address existential commitments, new constant symbols (es) are introduced for each sentence of the form \exists x \, \psi(x) in T_0; specifically, a fresh constant c is added along with \psi(c), ensuring the extension remains consistent. This process is repeated transfinitely, incorporating infinitely many such es if necessary, to produce a "Henkin " T that is maximal consistent and satisfies a witness condition: for every existential sentence in T, its witness instantiation is also in T. From this enriched , a model is constructed by taking the set of all introduced constants as the and defining interpretations based on the truth in T. Leon Henkin introduced this construction in his PhD thesis, providing a more intuitive and general alternative to Kurt Gödel's original proof of , which relied on the and canonical models without explicit witness handling. Henkin's approach proved particularly adaptable for extensions to higher-order logics and non-classical systems. A detailed proof follows in the subsequent section.

Proof Sketch for Henkin's Theorem

The proof of Henkin's theorem relies on constructing a model for a consistent in , assuming the language is countable to ensure the process terminates without cycles. The first step enumerates all sentences of the language as \sigma_1, \sigma_2, \dots, and iteratively extends the original consistent \Gamma by adding witness constants for existential quantifiers. Specifically, for each sentence \sigma_n of the form \exists x \, \phi(x), a new constant c_n is introduced, and the \phi(c_n) is added to the extension, ensuring the theory remains consistent at each finite stage. The second step extends this countable chain to a maximal consistent theory T_m using the , which guarantees that T_m is consistent, closed under , and includes witnesses for all existentials in the original theory, preserving consistency throughout the process. In the third step, a is constructed from T_m, with the domain consisting of all (including the new constants) in the expanded language, interpretations defined via the theory's sentences (e.g., t_1 = t_2 holds if T_m \vdash t_1 = t_2, and R(t_1, \dots, t_k) holds if T_m \vdash R(t_1, \dots, t_k)), forming a term model that satisfies the axioms by . Finally, satisfaction is verified by showing that for any sentence \sigma in the original theory, if \sigma \in T_m, then it is true in the model by the construction of interpretations; conversely, the maximality of T_m ensures that any true sentence in the model is provable from T_m, thus from \Gamma, establishing semantic consistency.

Consistency in Arithmetic and Set Theory

Peano Arithmetic

Peano arithmetic (PA) is the canonical axiomatization of the natural numbers, providing a formal for within . It consists of a with symbols for zero (0), the successor function (S), addition (+), multiplication (×), logical connectives, quantifiers, and variables ranging over natural numbers. The axioms of PA fall into two categories: a of basic axioms defining the structure of natural numbers and operations, and an infinite axiom schema for . The basic axioms are as follows:
  • ∀x (x ≠ 0 → ∃y (x = S(y))) (every is either or a successor).
  • ∀x ¬(S(x) = 0) ( is not the successor of any ).
  • ∀x ∀y (S(x) = S(y) → x = y) (the is injective).
  • ∀x (x + 0 = x) and ∀x ∀y (x + S(y) = S(x + y)) (recursive definition of ).
  • ∀x (x × 0 = 0) and ∀x ∀y (x × S(y) = (x × y) + x) (recursive definition of ).
These axioms ensure the form an ordered with a distinguished and successor operation. The schema states that for any φ(x) in the language of , (\phi(0) \land \forall x \, (\phi(x) \to \phi(S(x)))) \to \forall x \, \phi(x), capturing the principle that properties true of zero and preserved under successor hold for all natural numbers. This schema generates infinitely many axioms, one for each formula φ, making PA recursively axiomatizable. Regarding consistency, PA is known to be consistent relative to stronger formal systems, such as primitive recursive arithmetic extended by transfinite induction up to the ordinal ε₀, as established by Gerhard Gentzen in 1936. However, Kurt Gödel's second incompleteness theorem demonstrates that if PA is consistent, then the statement of PA's consistency, denoted Con(PA) and formalizing the assertion that 0 = 1 is not provable in PA, cannot be proved within PA itself. This theorem applies to any consistent extension of PA capable of representing basic recursion theory, rendering the absolute consistency of PA undecidable within the system. In the context of first-order logic, the consistency of PA thus refers to the absence of any derivation of a contradiction, such as 0 = 1, from its axioms. A related but stronger notion is ω-consistency, introduced by Gödel to strengthen the assumption in his original proof of the first incompleteness theorem. A theory is ω-consistent if, for any formula ψ(x), it does not prove both ∀x ψ(x) and ∃x ¬ψ(x). This condition prevents the theory from proving a universal statement while also proving its negation for each specific instance, ensuring with respect to the of arithmetic. ω-consistency implies ordinary consistency but is strictly stronger; notably, while the of is ω-consistent, there exist consistent extensions of that are not ω-consistent, such as certain non-standard models where infinite descending chains appear. This distinction is crucial for results like the original formulation of Gödel's first incompleteness theorem, which requires ω-consistency to guarantee the existence of undecidable sentences.

Zermelo-Fraenkel Set Theory

Zermelo-Fraenkel set theory (ZF) provides a foundational axiomatic framework for mathematics, comprising nine axioms that govern the formation and properties of sets. These include the axiom of extensionality, which states that two sets are equal if they have the same elements; the axiom of the empty set, asserting the existence of a set with no elements; the axiom of pairing, allowing the formation of a set containing any two given sets; the axiom of union, which ensures the set of all elements of sets in a given collection exists; the axiom of power set, guaranteeing the existence of the set of all subsets of a given set; the axiom schema of separation (or specification), asserting that for any existing set and first-order formula, the subset of elements satisfying the formula exists; the axiom of infinity, positing the existence of an infinite set; the axiom schema of replacement, enabling the substitution of elements via definable functions to form new sets; and the axiom of foundation (or regularity), preventing infinite descending membership chains. The development of ZF began with Ernst Zermelo's 1908 axiomatization, which included most of these axioms except replacement and foundation, aimed at resolving paradoxes like Russell's while enabling the . Abraham Fraenkel refined the system in 1922 by introducing the to overcome limitations in separation, ensuring the theory could handle transfinite recursions more robustly. A notable variant is the von Neumann–Bernays–Gödel (NBG) set theory, formalized in the 1930s and 1940s, which extends ZF by treating classes as primitive alongside sets, providing a conservative extension equivalent in power to ZF but facilitating proofs involving proper classes. The consistency of ZF hinges on the of a model satisfying all its axioms; by the completeness theorem, this is equivalent to no formal being provable within the system. Relative consistency proofs demonstrate that if a stronger theory is consistent, then so is ZF—for instance, the of a strongly κ implies that the cumulative hierarchy V_κ forms a model of ZF, as κ's uncountability, regularity, and strong limit properties ensure closure under the axioms. No absolute proof of ZF's consistency exists due to Gödel's second incompleteness theorem, but these relative results bolster confidence in its coherence. A key result concerning ZF's consistency involves the (CH), which posits that there is no cardinal between the cardinality of the natural numbers and the continuum. proved in 1940 that the consistency of ZF implies the consistency of ZF + CH, via the constructible universe L, an inner model where CH holds. Complementing this, established in 1963, using forcing, that the consistency of ZF + CH implies the consistency of ZF + ¬CH, rendering CH independent of ZF.

Model-Theoretic Perspectives

Models and Consistency

In , a is semantically consistent if there exists a , or model, that satisfies all the in the . This contrasts with syntactic consistency, which requires that no is derivable from the using the system. A model consists of a non-empty together with interpretations for the language's constants, functions, and predicates such that every holds true in the under Tarski's semantic definition of truth. Herbrand models provide a specific semantic characterization for theories in clausal form, bridging syntactic clauses to semantic satisfiability. For a set of clauses in with only universal quantifiers, the Herbrand universe is the set of all ground terms generated from the function symbols, and a is a subset of this universe that makes the ground instances of the clauses true. Herbrand's theorem states that a clausal has a model it has a Herbrand model, linking the of clauses to propositional satisfiability over the Herbrand base. By Gödel's completeness theorem, syntactic and semantic consistency coincide for first-order logic: a theory is consistent in the syntactic sense if and only if it is semantically consistent, meaning it possesses a model. This equivalence ensures that the existence of a proof of contradiction from the theory is equivalent to the absence of any satisfying structure. The proof constructs a model from a maximally consistent set of sentences, as in Henkin's method, confirming that consistent theories are always satisfiable. Consistent theories may admit finite models, such as the axioms defining a finite cyclic group of order three, or exclusively infinite models, as in the theory of dense linear orders without endpoints, axiomatized by the sentences expressing totality, antisymmetry, , (∀x ∀y (x < y → ∃z (x < z ∧ z < y))), and the absence of minimal or maximal elements. In the latter case, no finite structure satisfies the density axiom, yet the theory remains consistent with models like the rational numbers under the standard order.

Compactness Theorem Applications

The compactness theorem in first-order logic states that a set of sentences \Gamma is satisfiable if and only if every finite subset of \Gamma is satisfiable. Equivalently, for a theory T, T is consistent if and only if every finite subset of T is consistent. This finite character property implies that logical entailment from an infinite theory reduces to finite subtheories: if every finite subset T_n \models \phi, then T \models \phi. A proof of the compactness theorem can be obtained using maximal consistent sets. Starting from a consistent theory T, apply to extend T to a maximal consistent set \Delta, which is complete in the sense that for every sentence \psi, exactly one of \psi or \neg\psi belongs to \Delta. The ensures such a \Delta defines a model via the canonical interpretation where terms are interpreted as equivalence classes under provable equivalence in \Delta. Alternatively, compactness follows from the existence of ultraproducts: if every finite subset of T is satisfiable, construct an ultraproduct of models of these subsets over a non-principal , yielding a model of the full theory T. One key application is the construction of non-standard models of arithmetic. Consider the theory of Peano arithmetic PA augmented with new constants c_n for each standard natural number n, and axioms c_m > c_n for all m > n. Every finite subset of these axioms is satisfiable in the standard model \mathbb{N} by interpreting the finitely many constants appropriately, so by compactness, the full extended theory has a model \mathcal{M}. In \mathcal{M}, the c_n form an infinite increasing sequence without upper bound, yielding non-standard elements larger than all standard numerals. Compactness underpins the Löwenheim-Skolem theorems. The downward Löwenheim-Skolem theorem states that every consistent in a countable has a countable model; it is proved using the Tarski-Vaught of elementary substructures. The upward Löwenheim-Skolem theorem states that if T has an infinite model of \kappa, then it has models of any \lambda \geq \kappa by adding \lambda many new distinct constants with axioms asserting their mutual distinctness and applying to the extended .

Advanced Topics

Relative Consistency

Relative consistency proofs establish the consistency of a theory T by reducing it to the consistency of a base or stronger theory S, typically through an interpretation of T within S or by constructing a model of T from a model of S. If S is consistent, this implies T cannot prove a , as any such proof in T would translate to one in S. This approach avoids direct absolute consistency proofs, which are often unattainable due to limitations in formal systems, and has been central to for securing the foundations of through relative means. A classic example is David Hilbert's demonstration of the relative consistency of with respect to . In his foundational work, Hilbert showed that the axioms of can be interpreted within the framework of , specifically by coordinatizing points with real numbers constructed arithmetically, such that any contradiction in geometry would yield one in . This relative consistency result, achieved around 1900, underscored the potential of as a secure base for geometric reasoning, assuming the consistency of Peano arithmetic (). In , provided a seminal relative consistency proof by introducing the constructible universe L, showing that Zermelo-Fraenkel set theory with the (ZF + V = L), along with the (AC) and the generalized (GCH), is consistent relative to ZF. constructed L as a definable inner model of any satisfying ZF, where V = L holds, and demonstrated that L satisfies ZF, AC, and GCH. This establishes that if ZF is consistent, then ZF + V = L + AC + GCH is consistent as well. This 1938–1940 result not only secured the consistency of these additional axioms relative to ZF but also introduced the inner model approach central to descriptive set theory and forcing techniques. Complementing Gödel's results, developed the method of forcing in 1963 to prove the relative consistency of the negations of and GCH with ZF. By constructing generic extensions of models of ZF, Cohen showed that if ZF is consistent, then so are ZF + ¬ and ZF + ¬CH, establishing the independence of these axioms from ZF. , which sought finitary consistency proofs for infinite theories using only finite methods, faced fundamental limitations revealed by Gödel's second incompleteness theorem in 1931. Gödel demonstrated that any consistent capable of expressing basic arithmetic, such as , cannot prove its own consistency using only finitary means internal to the system, undermining the program's goal of absolute finitary justifications. This shift prompted explorations beyond strict , as in Gerhard Gentzen's 1936 proof of 's consistency relative to extended by up to the ordinal \varepsilon_0. Gentzen formalized cut-elimination in and used ordinal-based induction to bound proof lengths, showing that no infinite descent to contradiction is possible, thus establishing Con() relative to this stronger but still constructive system.

Incompleteness and Consistency

Gödel's first incompleteness theorem states that any consistent capable of expressing basic arithmetic, such as Peano arithmetic (), is incomplete, meaning there exist true statements in the language of the system that cannot be proved or disproved within it. This theorem implies that such systems cannot prove all arithmetical truths, highlighting inherent limitations in formal axiomatizations of . The second incompleteness theorem, also established by Gödel, asserts that if a T is consistent and sufficiently powerful to interpret arithmetic (as does), then the consistency of T, denoted Con(T), cannot be proved within T itself. This result follows from the first theorem by showing that a proof of Con(T) in T would allow a derivation of the Gödel sentence (the undecidable statement from the first theorem), leading to a if T is consistent. Consequently, for strong theories like , absolute consistency proofs within finitary methods are impossible, as any such proof would require resources beyond the system's own capabilities. These theorems have profound implications for of , precluding finitary consistency proofs for theories strong enough to formalize and necessitating alternative approaches for establishing relative consistency. One such method is , pioneered by , which proves the consistency of relative to a weaker incorporating up to the ordinal ε₀. In this framework, proofs are analyzed by assigning ordinals to their complexity, demonstrating termination below ε₀, thereby providing a non-finitary but rigorous relative consistency result without invoking the full strength of . The second incompleteness theorem extends naturally to set theory, particularly (ZF), since ZF interprets arithmetic via its axioms for natural numbers. Thus, if ZF is consistent, it cannot prove its own consistency, Con(ZF), within the system. This application underscores the theorem's broad reach, applying to any consistent axiomatizable extension of , including foundational systems like ZF.

References

  1. [1]
  2. [2]
    Propositional Logic | Internet Encyclopedia of Philosophy
    For example, two statements are said to be consistent when it is possible for both to be true, and are said to be inconsistent when it is not possible for both ...
  3. [3]
    Formal Logic - Philosophy 160 (002) - UMSL
    Definition: a set of sentences is logically consistent if and only if it is possible for all the members of that set to be true. A consistent set of sentences ...
  4. [4]
    Consistency theory - APA Dictionary of Psychology
    Apr 19, 2018 · a class of social psychological theory holding that people are chiefly motivated by a desire to maintain congruence or consistency among ...Missing: philosophy computing
  5. [5]
    Encyclopedia of Social Psychology - Cognitive Consistency
    The term consistency refers to consistency across cognitions, meaning that cognitions should be in agreement, symmetrical, balanced, or ...Missing: philosophy computing
  6. [6]
    Consistency and Ethics - Santa Clara University
    Consistency—the absence of contradictions—has sometimes been called the hallmark of ethics. Ethics is supposed to provide us with a guide for moral living, ...Missing: computing | Show results with:computing
  7. [7]
    Data Consistency vs Data Integrity: Similarities and Differences | IBM
    What is data consistency? Data consistency refers to the state of data in which all copies or instances are the same across all systems and databases.
  8. [8]
    What is Database Consistency? Definition & FAQs | ScyllaDB
    Consistency in database systems refers to the requirement that any given database transaction must change affected data only in allowed ways.
  9. [9]
    1.5: Dimensional Analysis - Physics LibreTexts
    Aug 11, 2021 · By the definition of dimensional consistency, we need to check that each term in a given equation has the same dimensions as the other terms ...
  10. [10]
    Proof Theory > A. Formal Axiomatics: Its Evolution and ...
    In 1899 and 1900a,b, Hilbert formulated a (quasi-) syntactic notion of consistency: no finite number of logical steps lead (from given axioms) to a ...
  11. [11]
    Consistency Proof - an overview | ScienceDirect Topics
    A consistency proof is defined as a demonstration that a logical system does not lead to contradictions, often confirmed through the construction of a model or ...
  12. [12]
    Consistency | EPFL Graph Search
    The lack of contradiction can be defined in either semantic or syntactic terms. The semantic definition states that a theory is consistent if it has a model ...
  13. [13]
    [PDF] INTRODUCTION TO LOGIC Formalisation in Predicate Logic
    A set Γ of L -sentences is semantically consistent if and only if Γ is syntactically consistent. In order to show that a set of sentences is semantically or.
  14. [14]
    Russell's paradox - Stanford Encyclopedia of Philosophy
    Dec 18, 2024 · In response to Russell's paradox, David Hilbert also expanded his program of building a consistent, axiomatic foundation for mathematics that ...
  15. [15]
    Hilbert's Program - Stanford Encyclopedia of Philosophy
    Jul 31, 2003 · This proposal incorporated Hilbert's ideas from 1904 regarding direct consistency proofs, his conception of axiomatic systems, and also the ...
  16. [16]
    Kurt Gödel - Stanford Encyclopedia of Philosophy
    Feb 13, 2007 · Among his mathematical achievements at the decade's close is the proof of the consistency of both the Axiom of Choice and Cantor's Continuum ...
  17. [17]
    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.
  18. [18]
    [PDF] Some Easy Relative Consistency Proofs
    Two methods of attack: Syntactic : Show that if T2 leads to a contradiction, then so does T1. Semantic : Show that if T1 has a model, then so does T2. By Gödel ...
  19. [19]
    Provability Logic - Stanford Encyclopedia of Philosophy
    Apr 2, 2003 · An arithmetical theory \(T\) is defined to be \(\omega\)-consistent if and only if for all formulas A with free variable \(x\), \(T\vdash \neg\ ...
  20. [20]
    [PDF] First-Order Logic: Syntax and Semantics
    Feb 28, 2020 · Figure 1 gives the grammar for the syntax of first-order logic, which shows how these symbols are ... notation for dealing with variables.
  21. [21]
    [PDF] First-Order Logic - Syntax, Semantics, Resolution
    ≻C = (≻L)mul, the multi-set extension of the literal ordering ≻L. Notation: ≻ also for ≻L and ≻C. Ruzica Piskac. First-Order Logic - Syntax, Semantics, ...
  22. [22]
    [PDF] First-Order Logic in a Nutshell Syntax: Formulae, Formal Proofs, and ...
    Syntax: Formulae, Formal Proofs, and Consistency. Like any other written language, First-Order Logic is based on an alphabet, which consists of the following ...
  23. [23]
    [PDF] First-order Logic
    specific definition of consistency of sets of sentences might differ, but like ⊢, we want consistency to coincide with its semantic counterpart, satisfiability.<|control11|><|separator|>
  24. [24]
    [PDF] 5 Deduction in First-Order Logic - UCLA Department of Mathematics
    Theorem 5.6 (Deduction Theorem). Let Γ be a set of sentences, let A be a sentence, and let B be a formula. If Γ ∪ {A} ` ...
  25. [25]
    [PDF] The Completeness Theorem - Open Logic Project Builds
    The completeness theorem is one of the most fundamental results about logic. It comes in two formulations, the equivalence of which we'll prove. In its.Missing: prime | Show results with:prime
  26. [26]
    [PDF] The Deduction Theorem
    To begin thinking about the Deduction Theorem, consider first the concept of validity in classical propositional logic. In this setting, we have compound ...
  27. [27]
    [PDF] LOGIC I 1. The Completeness Theorem 1.1. On consequences and ...
    We show below, using Zorn's Lemma, that every consistent theory can be extended to a complete consistent theory. Recall that Zorn's Lemma states that ...
  28. [28]
    first-order theory - PlanetMath.org
    Mar 22, 2013 · (Tarski) Every consistent theory T T is included in a complete theory. Proof : Use Zorn's lemma on the set of consistent theories that include ...
  29. [29]
  30. [30]
    [PDF] Henkin's Method and the Completeness Theorem
    Henkin's method is a construction that made the completeness theorem widely understood, and is the standard way to prove completeness.
  31. [31]
    [PDF] The Completeness of the First-Order Functional Calculus
    The Completeness of the First-Order Functional Calculus. Author(s): Leon Henkin. Source: The Journal of Symbolic Logic, Vol. 14, No. 3 (Sep., 1949), pp. 159 ...
  32. [32]
    [PDF] Completeness for First-order Logic I
    Suppose Σ is a consistent set, then by the previous lemma, there exists T ⊇ Σ such that T is a consistent Henkin. Theory. Then, by Lindenbaum's lemma, T can ...
  33. [33]
    [PDF] On Herbrand's Theorem - UCSD Math
    This paper discusses the famous theorem of Herbrand, which is one of the central theorems of proof-theory. The theorem called “Herbrand's theorem” in modern-.
  34. [34]
    [PDF] Herbrand logic - Computer Science
    Aug 28, 2009 · We have employed the name Herbrand logic, to reflect its first-order syntax and. Herbrand semantics. This paper's original form was a website ...
  35. [35]
    [PDF] MODEL THEORY Contents 1. First order logic 2 1.1. Structures 2 1.2 ...
    b) The theory of dense linear orders without endpoints is not ω- stable. ... A complete countable theory Φ with only infinite models is κ-categorical iff ...<|separator|>
  36. [36]
    The completeness and compactness theorems of first-order logic
    Apr 10, 2009 · In summary: to create a countable model from a consistent first-order theory, one first replaces the equality sign {=} (if any) by a binary ...
  37. [37]
    [PDF] com.1 The Compactness Theorem - Open Logic Project Builds
    The compactness theorem states that if each finite subset of a set of sentences is satisfiable, the entire set is satisfiable—even if the set itself is ...
  38. [38]
    [PDF] The compactness and Löwenheim–Skolem theorems
    Oct 30, 2012 · A simple application of the theorem is showing that given the consistency of some theory, non-standard models of that theory exists. This is ...
  39. [39]
    [PDF] Non-Standard Models of Arithmetic - UChicago Math
    May 1, 2004 · Use the Compactness Theorem to show consistency of. ΦP and Löwenheim - Skolem to get a countable model. ... non-standard model of arithmetic ...Missing: applications | Show results with:applications
  40. [40]
    Is Arithmetic Consistent? - MathPages
    Hilbert proved the consistency of Euclidean geometry by assuming the consistency of the theory of real numbers. This is an example of a relative consistency ...
  41. [41]