Gödel's incompleteness theorems

Gödel's incompleteness theorems

In mathematical logic, Gödel's incompleteness theorems, proved by Kurt Gödel in 1931, are two theorems stating inherent limitations of all but the most trivial formal systems for arithmetic of mathematical interest. The theorems are of considerable importance to the philosophy of mathematics. They are widely regarded as showing that Hilbert's program to find a complete and consistent set of axioms for all of mathematics is impossible, thus giving a negative answer to Hilbert's second problem.

Background

In mathematical logic, a (formal) theory is a set of statements expressed in a particular formal language. Most theories of interest are infinite. Some statements in a theory are included without proof (these are the axioms of the theory), and others (the theorems) are included because they are implied by the axioms.

Because statements of a formal theory are written in symbolic form, it is possible to mechanically verify that a formal proof from a finite set of axioms is valid. This task, known as automatic proof verification, is closely related to automated theorem proving; the difference is that instead of constructing a new proof, the proof verifier simply checks that a provided formal proof (or, in some cases, instructions that can be followed to create a formal proof) is correct. This is not merely hypothetical; systems such as Isabelle are used today to formalize proofs and then check their validity.

Many theories of interest include an infinite set of axioms, however. In order to verify a formal proof when the set of axioms is infinite, it must be possible to determine whether a statement that is claimed to be an axiom is actually an axiom. This issue arises in first order theories of arithmetic, such as Peano arithmetic, because the principle of mathematical induction is expressed as an infinite set of axioms (an axiom schema).

A formal theory is said to be effectively generated if its set of axioms is a recursively enumerable set. This means that there is a computer program that, in principle, could enumerate all the axioms of the theory without listing any statements that are not axioms. This is equivalent to the ability to enumerate all the theorems of the theory (including the axioms) without enumerating any statements that are not theorems. For examples, each of the theories of Peano arithmetic and Zermelo-Fraenkel set theory has an infinite number of axioms and each is effectively generated.

In choosing a set of axioms, one goal is to be able to prove as many correct results as possible, without proving any incorrect results. A set of axioms is complete if, for any statement in the axioms' language, either that statement or its negation is provable from the axioms. A set of axioms is (simply) consistent if there is no statement so that both the statement and its negation are provable from the axioms. In the standard system of first-order logic, an inconsistent set of axioms will prove every statement in its language (this is sometimes called the principle of explosion), and is thus automatically complete. A set of axioms that is both complete and consistent, however, will prove as many theorems as can be proved without containing a contradiction.

First incompleteness theorem

Gödel's first incompleteness theorem, perhaps the single most celebrated result in mathematical logic, states that:

: "Any effectively generated theory capable of expressing elementary arithmetic cannot be both consistent and complete. In particular, for any consistent, effectively generated formal theory that proves certain basic arithmetic truths, there is an arithmetical statement that is true, [The word "true" is used disquotationally here; the Gödel sentence "G""T" states that no natural number with a certain property exists, and to say "G""T" is true simply means that, in fact, no number with that property exists (see Franzén 2005 pp. 28–33). It is also possible to read "G""T" is true" in the formal sense that it is a theorem of primitive recursive arithmetic that Con("T")→"G""T", where Con("T") is a canonical sentence asserting the consistency of "T" (Smoryński 1977 p. 840, Kikuchi and Tanaka 1994 p. 403) ] but not provable in the theory.

The resulting true but unprovable statement is often referred to as "the Gödel sentence" for the theory, although there are infinitely many other statements in the theory that share with the Gödel sentence the property of being true but not provable from the theory.

Roughly speaking, for each theory "T" the corresponding Gödel sentence "G" asserts: "G" cannot be proved within the theory "T". If "G" were provable under the axioms and rules of inference of "T", then "T" would have a theorem, "G", which effectively contradicts itself, and thus the theory "T" would be inconsistent. This means that if the theory "T" is consistent then "G" cannot be proved within it. This means that "G"'s claim about its own unprovability is correct; in this sense "G" is not only unprovable but true. Thus provability-within-the-theory-"T" is not the same as truth; the theory "T" is incomplete.

It is possible to define a larger theory T' that contains the whole of T, plus G as an additional axiom. In this case, G is indeed a theorem in T' (trivially so, since it is an axiom). However, there will be a new Gödel statement G' for T', showing that T' is also incomplete. Each theory has its own Gödel statement.

In order to prove the first incompleteness theorem, Gödel represented statements by numbers. Then the theory at hand, which is assumed to prove certain facts about numbers, also proves facts about its own statements. Questions about the provability of statements are represented as questions about the properties of numbers, which would be decidable by the theory if it were complete. In these terms, the Gödel sentence states that no natural number exists with a certain, strange property. A number with this property would encode a proof of the inconsistency of the theory. If there were such a number then the theory would be inconsistent, contrary to the consistency hypothesis. So, assuming the theory is consistent (as done in the theorem's hypothesis) there is no such number.

Meaning of the first incompleteness theorem

Gödel's first incompleteness theorem shows that any formal system that includes enough of the theory of the natural numbers is incomplete: it contains statements that are neither provably true nor provably false. Or one might say, no formal system which aims to define the natural numbers can actually do so, as there will be true number-theoretical statements which that system cannot prove. This has severe consequences for the program of logicism proposed by Gottlob Frege and Bertrand Russell, which aimed to define the natural numbers in terms of logic (Hellman 1981, p.451–468).

The existence of an incomplete system is in itself not particularly surprising. A system can be incomplete simply because not all the necessary axioms have been discovered. For example, Euclidean geometry without the parallel postulate is incomplete; it is not possible to prove or disprove the parallel postulate from the remaining axioms.

Gödel's theorem shows that, in theories that include a small portion of number theory, a complete and consistent finite list of axioms can "never" be created, or even an infinite list that can be enumerated by a computer program. Each time a new statement is added as an axiom, there are other true statements that still cannot be proved, even with the new axiom. If an axiom is ever added that makes the system complete, it does so at the cost of making the system inconsistent.

It "is" possible to have a complete and consistent list of axioms that "cannot" be produced by a computer program (that is, the list is not computably enumerable). For example, one might take all true statements about the natural numbers to be axioms (and no false statements). But then there is no mechanical way to decide, given a statement about the natural numbers, whether it is an axiom or not.

Many logicians believe that Gödel's incompleteness theorems struck a fatal blow to David Hilbert's program towards a universal mathematical formalism which was based on "Principia Mathematica". The generally agreed-upon stance is that the second theorem is what specifically dealt this blow. However some believe it was the first, and others believe that neither did.

Relation to the liar paradox

The liar paradox is the sentence "This sentence is false." An analysis of the liar sentence shows that it cannot be true (for then, as it asserts, it is false), nor can it be false (for then, it is true). A Gödel sentence "G" for a theory "T" makes a similar assertion to the liar sentence, but with truth replaced by provability: "G" says "G" is not provable in the theory "T"." The analysis of the truth and provability of "G" is a formalized version of the analysis of the truth of the liar sentence.

It is not possible to replace "not provable" with "false" in a Gödel sentence because the predicate "Q is the Gödel number of a false formula" cannot be represented as a formula of arithmetic. This result, known as Tarski's undefinability theorem, was discovered independently by Gödel (when he was working on the proof of the incompleteness theorem) and by Alfred Tarski.

Original statement

The first incompleteness theorem first appeared as "Theorem VI" in his 1931 paper "On Formally Undecidable Propositions in Principia Mathematica and Related Systems I." In Gödel's original notation, it states::"The general result about the existence of undecidable propositions reads as follows:::"Theorem VI. "For every ω-consistent recursive class κ of" FORMULAS "there are recursive" CLASS SIGNS "r, such that neither v" Gen "r nor" Neg("v" Gen "r") "belongs to Flg(κ)" (where "v" is the FREE VARIABLE of "r"). [Here Flg("κ") represents the theory generated by κ and "v" Gen "r" is a particular formula in the language of arithmetic. "Flg" is from "Folgerungsmenge = set of consequences" and "Gen" is from "Generalisation = generalization" (cf Meltzer and Braithwaite 1962, 1992 edition:33-34).] (van Heijenoort 1967:607.)

Extensions of Gödel's original result

Gödel demonstrated the incompleteness of the theory of Principia Mathematica, a particular theory of arithmetic, but a parallel demonstration could be given for any effective theory of a certain expressiveness. Gödel commented on this fact in the introduction to his paper, but restricted the proof to one system for concreteness. In modern statements of the theorem, it is common to state the effectiveness and expressiveness conditions as hypotheses for the incompleteness theorem, so that it is not limited to any particular formal theory. The terminology used to state these conditions was not yet developed in 1931 when Gödel published his results.

Gödel's original statement and proof of the incompleteness theorem requires the assumption that the theory is not just consistent but "ω-consistent". A theory is ω-consistent if it is not ω-inconsistent, and is ω-inconsistent if there is a predicate "P" such that for every specific natural number "n" the theory proves ~"P"("n"), and yet the theory also proves that there exists a natural number "n" such that "P"("n"). That is, the theory says that a number with property "P" exists while denying that it has any specific value. The ω-consistency of a theory implies its consistency, but consistency does not imply ω-consistency. J. Barkley Rosser (1936) strengthened the incompleteness theorem by finding a variation of the proof (Rosser's trick) that only requires the theory to be consistent, rather than ω-consistent. This is mostly of technical interest, since all true formal theories of arithmetic (theories whose axioms are all true statements about natural numbers) are ω-consistent, and thus Gödel's theorem as originally stated applies to them. The stronger version of the incompleteness theorem that only assumes consistency, rather than ω-consistency, is now commonly known as Gödel's incompleteness theorem and as the Gödel–Rosser theorem.

Carl Hewitt (2008) has shown that Gödel's argument can be extended to certain sufficiently strong paraconsistent theories. These theories, which permit self-application of formulas (see Feferman 1984), are able to prove their own Gödel sentences, and are thus inconsistent.

econd incompleteness theorem

Gödel's second incompleteness theorem can be stated as follows:: "For any formal recursively enumerable (i.e., effectively generated) theory T including basic arithmetical truths and also certain truths about formal provability, T includes a statement of its own consistency if and only if T is inconsistent."This strengthens the first incompleteness theorem, because the statement constructed in the first incompleteness theorem does not directly express the consistency of the theory. The proof of the second incompleteness theorem is obtained, essentially, by formalizing the proof of the first incompleteness theorem within the theory itself.

A technical subtlety in the second incompleteness theorem is how to express the consistency of T as a formula in the language of T. There are many ways to do this, and not all of them lead to the same result. In particular, different formalizations of the claim that T is consistent may be inequivalent in T, and some may even be provable. For example, first-order arithmetic (Peano arithmetic or PA for short) can prove that the largest consistent subset of PA is consistent. But since PA is consistent, the largest consistent subset of PA is just PA, so in this sense PA "proves that it is consistent". What PA does not prove is that the largest consistent subset of PA is, in fact, the whole of PA. (The term "largest consistent subset of PA" is rather vague, but what is meant here is the largest consistent initial segment of the axioms of PA ordered according to some criteria, e.g., by "Gödel numbers", the numbers encoding the axioms as per the scheme used by Gödel mentioned above).

In the case of Peano arithmetic or any familiar explicitly axiomatized theory T, it is possible to define the consistency "Con(T)" of T in terms of the non-existence of a number with a certain property, as follows: "there does not exist an integer coding a sequence of sentences, such that each sentence is either one of the (canonical) axioms of T, a logical axiom, or an immediate consequence of preceding sentences according to the rules of inference of first-order logic, and such that the last sentence is a contradiction". However, for arbitrary T there is no canonical choice for Con(T).Fact|date=August 2008

The formalization of Con(T) depends on two factors: formalizing the notion of a sentence being derivable from a set of sentences and formalizing the notion of being an axiom of T. Formalizing derivability can be done in canonical fashion, so given an arithmetical formula A(x) defining a set of axioms, we can canonically form the predicate ProvA(P) which expresses that P is provable from the set of axioms defined by A(x). Using this predicate we can express Con(T) as "not ProvA('P and not-P')". Solomon Feferman showed that Gödel's second incompleteness theorem goes through when the formula A(x) is chosen so that it has the form "there exists a number n satisfying the decidable predicate P" for some P. In addition, ProvA(P) must satisfy the so-called Hilbert-Bernays provability conditions:

1. If T proves P, then T proves ProvA(P)

2. T proves 1., i.e. T proves that if T proves P, then T proves ProvA(P)

3. T proves that if T proves that (P implies Q) then T proves that provability of P implies provability of Q

Gödel's second incompleteness theorem also implies that a theory T1 satisfying the technical conditions outlined above can't prove the consistency of any theory T2 which proves the consistency of T1. This is because then T1 can prove that if T2 proves the consistency of T1, then T1 is in fact consistent. For the claim that T1 is consistent has form "for all numbers n, n has the decidable property of not being a code for a proof of contradiction in T1". If T1 were in fact inconsistent, then T2 would prove for some n that n is the code of a contradiction in T1. But if T2 also proved that T1 is consistent, i.e., there is no such n, it would itself be inconsistent. We can carry out this reasoning in T1 and conclude that if T2 is consistent, then T1 is consistent. Since by second incompleteness theorem, T1 does not prove its consistency, it can't prove the consistency of T2 either.

This easy corollary of the second incompleteness theorem shows that there is no hope of proving e.g. the consistency of first-order arithmetic using finitistic means provided we accept that finitistic means are correctly formalized in a theory the consistency of which is provable in PA. It's generally accepted that the theory of primitive recursive arithmetic (PRA) is an accurate formalization of finitistic mathematics, and PRA is provably consistent in PA. Thus PRA can't prove the consistency of PA. This is generally seen to show that Hilbert's program, which is to use "ideal" mathematical principles to prove "real" (finitistic) mathematical statements by showing that the "ideal" principles are consistent by finitistically acceptable principles, can't be carried out.

This corollary is actually what makes the second incompleteness theorem epistemically relevant. As Georg Kreisel remarked, it would actually provide no interesting information if a theory T proved its consistency. This is because inconsistent theories prove everything, including their consistency. Thus a consistency proof of T in T would give us no clue as to whether T really is consistent; no doubts about T's consistency would be resolved by such a consistency proof. The interest in consistency proofs lies in the possibility of proving the consistency of a theory T in some theory T' which is in some sense less doubtful than T itself, e.g. weaker than T. For most naturally occurring T and T', such as T = Zermelo-Fraenkel set theory and T' = primitive recursive arithmetic, the consistency of T' is provable in T, and thus T' can't prove the consistency of T by the above corollary of the second incompleteness theorem.

The consistency of first-order arithmetic has been proved assuming that a certain ordinal called ε0 is wellfounded. See Gentzen's consistency proof.

Original statement of Gödel’s Theorem XI

While contemporary usage calls it the “Second incompleteness Theorem”, in the original Gödel presented it as his “Theorem XI”. It is stated thus (in the following, “Section 2” is where his Theorem VI appears, and "P" is Gödel’s abbreviation for the system obtained by adding the Peano axioms to the logical system of Principia Mathematica.:”The results of Section 2 have a surprising consequence concerning a consistency proof for the system "P" (and its extensions), which can be stated as follows:::”Theorem XI. "Let κ be any recursive consistent63 class of" FORMULAS; "then the" SENTENTIAL FORMULA "stating that κ is consistent is not κ"-PROVABLE; in particular, the consistency of "P" is not provable in "P",64 provided "P" is consistent (in the opposite case, of course, every proposition is provable [in P] )". (Brackets in original added by Gödel “to help the reader”, translation and typography in van Heijenoort 1967:614):::”63 “κ is consistent” (abbreviated by “Wid(κ)”) is defined as thus: Wid(κ)≡ ("Ex")(Form("x") & ~Bewκ("x"))." :::::(Note: In the original "Bew" has a negation-“bar” written over it, indicated here by ~. “Wid” abbreviates “Widerspruchfreiheit = consistency”, “Form” abbreviates “Formel = formula”, “Bew” abbreviates “Beweisbar = provable” (translations from Meltzer and Braithwaite 1962, 1996 edition:33-34) ) :::”64 This follows if we substitute the empty class of FORMULAS for κ.”

Examples of undecidable statements

:"See also: List of statements undecidable in ZFCThere are two distinct senses of the word "undecidable" in mathematics and computer science. The first of these is the proof-theoretic sense used in relation to Gödel's theorems, that of a statement being neither provable nor refutable in a specified deductive system. The second sense, which will not be discussed here, is used in relation to computability theory and applies not to statements but to decision problems, which are countably infinite sets of questions each requiring a yes or no answer. Such a problem is said to be undecidable if there is no computable function that correctly answers every question in the problem set (see undecidable problem).

Because of the two meanings of the word undecidable, the term independent is sometimes used instead of undecidable for the "neither provable nor refutable" sense. The usage of "independent" is also ambiguous, however. Some use it to mean just "not provable", leaving open whether an independent statement might be refuted.

Undecidability of a statement in a particular deductive system does not, in and of itself, address the question of whether the truth value of the statement is well-defined, or whether it can be determined by other means. Undecidability only implies that the particular deductive system being considered does not prove the truth or falsity of the statement. Whether there exist so-called "absolutely undecidable" statements, whose truth value can never be known or is ill-specified, is a controversial point in the philosophy of mathematics.

The combined work of Gödel and Paul Cohen has given two concrete examples of undecidable statements (in the first sense of the term): The continuum hypothesis can neither be proved nor refuted in ZFC (the standard axiomatization of set theory), and the axiom of choice can neither be proved nor refuted in ZF (which is all the ZFC axioms "except" the axiom of choice). These results do not require the incompleteness theorem. Gödel proved in 1940 that neither of these statements could be disproved in ZF or ZFC set theory. In the 1960s, Cohen proved that neither is provable from ZF, and the continuum hypothesis cannot be proven from ZFC.

In 1973, the Whitehead problem in group theory was shown to be undecidable, in the first sense of the term, in standard set theory.

In 1977, Paris and Harrington proved that the Paris-Harrington principle, a version of the Ramsey theorem, is undecidable in the first-order axiomatization of arithmetic called Peano arithmetic, but can be proven to be true in the larger system of second-order arithmetic. Kirby and Paris later showed Goodstein's theorem, a statement about sequences of natural numbers somewhat simpler than the Paris-Harrington principle, to be undecidable in Peano arithmetic.

Kruskal's tree theorem, which has applications in computer science, is also undecidable from Peano arithmetic but provable in set theory. In fact Kruskal's tree theorem (or its finite form) is undecidable in a much stronger system codifying the principles acceptable on basis of a philosophy of mathematics called predicativism. The related but more general graph minor theorem (2003) has consequences for computational complexity theory.

Gregory Chaitin produced undecidable statements in algorithmic information theory and proved another incompleteness theorem in that setting. Chaitin's theorem states that for any theory that can represent enough arithmetic, there is an upper bound "c" such that no specific number can be proven in that theory to have Kolmogorov complexity greater than "c". While Gödel's theorem is related to the liar paradox, Chaitin's result is related to Berry's paradox.

Limitations of Gödel's theorems

The conclusions of Gödel's theorems only hold for the formal theories that satisfy the necessary hypotheses. Not all axiom systems satisfy these hypotheses, even when these systems have models that include the natural numbers as a subset. For example, there are first-order axiomatizations of Euclidean geometry and real closed fields that do not meet the hypotheses of Gödel's theorems. The key fact is that these axiomatizations are not expressive enough to define the set of natural numbers or develop basic properties of the natural numbers.

Gödel's theorems only apply to consistent theories. In first-order logic, because of the principle of explosion, an inconsistent theory proves "every" formula in its language.

Gödel's theorems only apply to recursively enumerable theories. If all true statements about naturals are taken as axioms for a theory, then this theory is a consistent, complete extension of Peano arithmetic for which none of Gödel's theorems hold, because this theory is not recursively enumerable.

Gödel's theorems only show that the consistency of certain theories cannot be proved from the axioms of those theories themselves. It does not show that the consistency cannot be proved from other (consistent) axioms. For example, the consistency of the Peano arithmetic can be proved in Zermelo-Frankel set theory (ZFC). In 1936, Gerhard Gentzen proved the consistency of Peano arithmetic using a formal system which was more powerful in certain aspects than arithmetic, but less powerful than common axiom system for systems of set theory.

Relationship with computability

The incompleteness theorem is closely related to several results about undecidable sets in recursion theory.

Stephen Cole Kleene (1943) presented a proof of Gödel's incompleteness theorem using basic results of computability theory. One such result shows that the halting problem is unsolvable: there is no computer program that can correctly determine, given a program "P" as input, whether "P" eventually halts when run with no input. Kleene showed that the existence of a complete effective theory of arithmetic with certain consistency properties would force the halting problem to be decidable, a contradiction. Expositions of this proof at the undergraduate level are given by Charlesworth (1980) and Hopcroft and Ullman (1979).

Franzén (2005, p. 73) explains how Matiyasevich's solution to Hilbert's 10th problem can be used to obtain a proof to Gödel's first incompleteness theorem. Matiyasevich proved that there is no algorithm that, given a multivariate polynomial p(x1, x2,...,xk) with integer coefficients, determines whether there is an integer solution to the equation "p" = 0. Because polynomials with integer coefficients, and integers themselves, are directly expressible in the language of arithmetic, if a multivariate integer polynomial equation "p" = 0 does have a solution in the integers then any sufficiently strong theory of arithmetic "T" will prove this. Moreover, if the theory "T" is ω-consistent, then it will never prove that some polynomial equation has a solution when in fact there is no solution in the integers. Thus, if "T" were complete and ω-consistent, it would be possible to algorithmically determine whether a polynomial equation has a solution by merely enumerating proofs of "T" until either "p" has a solution" or "p" has no solution" is found, in contradiction to Matiyasevich's theorem.

Smorynski (1977, p. 842) shows how the existence of recursively inseparable sets can be used to prove the first incompleteness theorem. This proof is often extended to show that systems such as Peano arithmetic are essentially undecidable (see Kleene 1967, p. 274).

Proof sketch for the first theorem

Throughout the proof we assume a formal system is fixed and satisfies the necessary hypotheses. The proof has three essential parts. The first part is to show that statements can be represented by natural numbers, known as Gödel numbers, and that properties of the statements can be detected by examining their Gödel numbers. This part culminates in the construction of a formula expressing the idea that a statement is provable in the system. The second part of the proof is to construct a particular statement that, essentially, says that it is unprovable. The third part of the proof is to analyze this statement to show that it is neither provable nor disprovable in the system.

Arithmetization of syntax

The main problem in fleshing out proof described above is that it seems at first that in order to construct a statement "p" that is equivalent to "p"cannot be proved", "p" would have to somehow contain a reference to "p", which could easily give rise to an infinite regress. Gödel's ingenious trick, which was later used by Alan Turing in his work on the Entscheidungsproblem, is to represent statements as numbers, which is often called the arithmetization of syntax. This allows a self-referential formula to be constructed in a way that avoids any infinite regress of definitions.

To begin with, every formula or statement that can be formulated in our system gets a unique number, called its Gödel number. This is done in such a way that it is easy to mechanically convert back and forth between formulas and Gödel numbers. It is similar, for example, to the way English sentences are encoded as sequences (or "strings") of numbers using ASCII: such a sequence is considered as a single (if potentially very large) number. Because our system is strong enough to reason about "numbers", it is now also possible to reason about "formulas" within the system.

A formula "F"("x") that contains exactly one free variable "x" is called a "statement form" or "class-sign". As soon as "x" is replaced by a specific number, the statement form turns into a "bona fide" statement, and it is then either provable in the system, or not. For certain formulas one can show that for every natural number n, F(n) is true if and only if it can be proven (the precise requirement in the original proof is weaker, but for the proof sketch this will suffice). In particular, this is true for every specific arithmetic operation between a finite number of natural numbers, such as "2*3=6".

Statement forms themselves are not statements and therefore cannot be proved or disproved. But every statement form "F"("x") can be assigned with a Gödel number which we will denote by G("F"). The choice of the free variable used in the form "F"("x") is not relevant to the assignment of the Gödel number G("F").

Now comes the trick: The notion of provability itself can also be encoded by Gödel numbers, in the following way. Since a proof is a list of statements which obey certain rules, we can define the Gödel number of a proof. Now, for every statement "p", we may ask whether a number "x" is the Gödel number of its proof. The relation between the Gödel number of "p" and "x", the Gödel number of its proof, is an arithmetical relation between two numbers. Therefore there is a statement form Bew("x") that uses this arithmetical relation to state that a Gödel number of a proof of "x" exists::Bew("y") = ∃ "x" ( "y" is the Gödel number of a formula and "x" is the Gödel number of a proof of the formula encoded by "y").

The name Bew is short for "beweisbar", the German word for "provable". An important feature of Bew is that if a statement "p" is provable in the system then Bew(G("p")) is also provable. This is because any proof of "p" would have a corresponding Gödel number, the existence of which causes Bew(G("p")) to be satisfied.

Diagonalization

The next step in the proof is to obtain a statement that says it is unprovable. Although Gödel constructed this statement directly, the existence of at least one such statement follows from the diagonal lemma, which says that for any sufficiently strong formal system and any statement form "F" there is a statement "p" such that the system proves:"p" ↔ "F"(G("p")).We obtain "p" by letting "F" be the negation of Bew("x"); thus "p" roughly states that its own Gödel number is the Gödel number of an unprovable formula.

The statement "p" is not literally equal to ~Bew(G("p")); rather, "p" states that if a certain calculation is performed, the resulting Gödel number will be that of an unprovable statement. But when this calculation is performed, the resulting Gödel number turns out to be the Gödel number of "p" itself. This is similar to the following sentence in English::", when preceded by itself in quotes, is unprovable.", when preceded by itself in quotes, is unprovable.This sentence does not directly refer to itself, but when the stated transformation is made the original sentence is obtained as a result, and thus this sentence asserts its own unprovability. The proof of the diagonal lemma employs a similar method.

Proof of independence

We will now assume that our axiomatic system is ω-consistent.We let "p" be the statement obtained in the previous section.

If "p" were provable, then Bew(G("p")) would be provable, as argued above. But "p" asserts the negation of Bew(G("p")). Thus our system would be inconsistent, proving both a statement and its negation. This contradiction shows that "p" cannot be provable.

If the negation of "p" were provable, then Bew(G("p")) would be provable (because "p" was constructed to be equivalent to the negation of Bew(G("p"))). However, for each specific number "x", "x" cannot be the Gödel number of the proof of "p", because "p" is not provable (from the previous paragraph). Thus on one hand the system proves there is a number with a certain property (that it is the Gödel number of the proof of "p"), but on the other hand, for every specific number "x", we can prove that it does not have this property. This is impossible in an ω-consistent system. Thus the negation of "p" is not provable.

Thus the statement "p" is undecidable: it can neither be proved nor disproved within the system.

It should be noted that "p" is not provable (and thus true) in every consistent system. The assumption of ω-consistency is only required for the negation of "p" to be not provable. Thus:
*In an ω-consistent formal system, we may prove neither "p" nor its negation, and so "p" is undecidable.
*In a consistent formal system we may either have the same situation, or we may prove the negation of "p"; In the later case, we have a statement ("not "p") which is false but provable.

Note that if one tries to "add the missing axioms" in order to avoid the undecidability of the system, then one has to add either "p" or "not "p" as axioms. But then the definition of "being a Gödel number of a proof" of a statement changes. which means that the statement form Bew("x") is now different. Thus when we apply the diagonal lemma to this new form Bew, we obtain a new statement "p", different from the previous one, which will be undecidable in the new system if it is ω-consistent.

Proof via Berry's paradox

George Boolos (1989) sketches an alternative proof of the first incompleteness theorem that uses Berry's paradox rather than the liar paradox to construct a true but unprovable formula. A similar proof method was independently discovered by Saul Kripke (Boolos 1998, p. 383). Boolos's proof proceeds by constructing, for any computably enumerable set "S" of true sentences of arithmetic, another sentence which is true but not contained in "S". This gives the first incompleteness theorem as a corollary. According to Boolos, this proof is interesting because it provides a "different sort of reason" for the incompleteness of effective, consistent theories of arithmetic (Boolos 1998, p. 388).

Proof sketch for the second theorem

The main difficulty in proving the second incompleteness theorem is to show that various facts about provability used in the proof of the first incompleteness theorem can be formalized within the system using a formal predicate for provability. Once this is done, the second incompleteness theorem essentially follows by formalizing the entire proof of the first incompleteness theorem within the system itself.

Let "p" stand for the undecidable sentence constructed above, and assume that the consistency of the system can be proven from within the system itself. We have seen above that if the system is consistent, then "p" is not provable. The proof of this implication can be formalized within the system, and therefore the statement "p" is not provable", or "not "P"("p")" can be proven in the system.

But this last statement is equivalent to "p" itself (and this equivalence can be proven in the system), so "p" can be proven in the system. This contradiction shows that the system must be inconsistent.

Discussion and implications

The incompleteness results affect the philosophy of mathematics, particularly versions of formalism, which use a single system formal logic to define their principles. One can paraphrase the first theorem as saying, "we can never find an all-encompassing axiomatic system which is able to prove "all" mathematical truths, but no falsehoods."

On the other hand, from a strict formalist perspective this paraphrase would be considered meaningless because it presupposes that mathematical "truth" and "falsehood" are well-defined in an absolute sense, rather than relative to each formal system.

The following rephrasing of the second theorem is even more unsettling to the foundations of mathematics:

:"If an axiomatic system can be proven to be consistent and complete from within itself, then it is inconsistent."

Therefore, in order to establish the consistency of a system S, one needs to use some other "more powerful" system T, but a proof in T is not completely convincing unless T's consistency has already been established without using S.

At first, Gödel's theorems seemed to leave some hope—it was thought that it might be possible to produce a general algorithm that indicates whether a given statement is undecidable or not, thus allowing mathematicians to bypass the undecidable statements altogether. However, the negative answer to the Entscheidungsproblem, obtained in 1936, showed that no such algorithm exists.

There are some who hold that a statement that is unprovable within a deductive system may be quite provable in a metalanguage. And what cannot be proven in "that" metalanguage can likely be proven in a meta-metalanguage, recursively, ad infinitum, in principle. By invoking such a system of typed metalanguages, along with an axiom of Reducibility — which by an inductive assumption applies to the entire stack of languages — one may, for all practical purposes, overcome the obstacle of incompleteness.

Note that Gödel's theorems only apply to "sufficiently strong" axiomatic systems. "Sufficiently strong" means that the theory contains enough arithmetic to carry out the coding constructions needed for the proof of the first incompleteness theorem. Essentially, all that is required are some basic facts about addition and multiplication as formalized, e.g., in Robinson arithmetic Q. There are even weaker axiomatic systems that are consistent and complete, for instance Presburger arithmetic which proves every true first-order statement involving only addition.

The axiomatic system may consist of infinitely many axioms (as first-order Peano arithmetic does), but for Gödel's theorem to apply, there has to be an effective algorithm which is able to check proofs for correctness. For instance, one might take the set of all first-order sentences which are true in the standard model of the natural numbers. This system is complete; Gödel's theorem does not apply because there is no effective procedure that decides if a given sentence is an axiom. In fact, that this is so is a consequence of Gödel's first incompleteness theorem.

Another example of a specification of a theory to which Gödel's first theorem does not apply can be constructed as follows: order all possible statements about natural numbers first by length and then lexicographically, start with an axiomatic system initially equal to the Peano axioms, go through your list of statements one by one, and, if the current statement cannot be proven nor disproven from the current axiom system, add it to that system. This creates a system which is complete, consistent, and sufficiently powerful, but not computably enumerable.

Gödel himself only proved a technically slightly weaker version of the above theorems; the first proof for the versions stated above was given by J. Barkley Rosser in 1936.

In essence, the proof of the first theorem consists of constructing a statement "p" within a formal axiomatic system that can be given a meta-mathematical interpretation of:

:"p" = "This statement cannot be proven in the given formal theory"

As such, it can be seen as a modern variant of the Liar paradox, although unlike the classical paradoxes it's not really paradoxical.

If the axiomatic system is consistent, Gödel's proof shows that "p" (and its negation) cannot be proven in the system. Therefore "p" is true ("p" claims to be not provable, and it is not provable) yet it cannot be formally proved in the system. If the axiomatic system is ω-consistent, then the negation of "p" cannot be proven either, and so "p" is undecidable. In a system which is not ω-consistent (but consistent), either we have the same situation, or we have a false statement which can be proven (namely, the negation of "p").

Adding "p" to the axioms of the system would not solve the problem: there would be another Gödel sentence for the enlarged theory. Theories such as Peano arithmetic, for which any computably enumerable consistent extension is incomplete, are called essentially incomplete.

Minds and machines

Authors including J. R. Lucas have debated what, if anything, Gödel's incompleteness theorems imply about human intelligence. Much of the debate centers on whether the human mind is equivalent to a Turing machine, or by the Church-Turing thesis, any finite machine at all. If it is, and if the machine is consistent, then Gödel's incompleteness theorems would apply to it.

Hilary Putnam (1960) suggested that while Gödel's theorems cannot be applied to humans, since they make mistakes and are therefore inconsistent, it may be applied to the human faculty of science or mathematics in general. If we are to believe that it is consistent, then either we cannot prove its consistency, or it cannot be represented by a Turing machine.

Postmodernism and continental philosophy

Appeals are sometimes made to the incompleteness theorems to support by analogy ideas which go beyond mathematics and logic. For instance, Régis Debray applies it to politics. ["The secret takes the form of a logical law, an extension of Gödel's theorem: "There can be no organised system without closure and no system can be closed by elements internal to that system alone".Debray, R. "Critique of Political Reason, quoted in Sokal and Bricmont's "Fashionable Nonsense".] A number of authors have commented, mostly negatively, on such extensions and interpretations, including Torkel Franzén, Alan Sokal and Jean Bricmont, Ophelia Benson, Jacques Bouveresse, and Jeremy Stangroom. The last two quote [In their "Why Truth Matters"] biographer Rebecca Goldstein ["The Proof and paradox of Kurt Gödel"] commenting on the disparity between Gödel's avowed Platonism and the anti-realist uses to which his ideas are put by humanist intellectuals.

Theories of everything and physics

Stanley Jaki, followed much later by Stephen Hawking and others, argue that (an analogous argument to) Gödel's theorem implies that even the most sophisticated formulation of physics will be incomplete, and that therefore there can never be an ultimate theory that can be formulated as a finite number of principles, known for certain as "final". [Stanley Jaki [http://pirate.shu.edu/~jakistan/JakiGodel.pdf "A Late Awakening to Gödel in Physics"] ] [Stephen Hawking [http://www.damtp.cam.ac.uk/strings02/dirac/hawking/ "Gödel and the end of physics"] ]

ee also

* Consistency proof
* Gödel's completeness theorem
* Löb's Theorem
* Logicism
* Münchhausen Trilemma
* Non-standard arithmetic
* Self-verifying theories
* Tarski's indefinability theorem
* "Minds, Machines and Gödel"

Notes

References

Articles by Gödel

* 1931, "Über formal unentscheidbare Sätze der Principia Mathematica und verwandter Systeme, I." "Monatshefte für Mathematik und Physik 38": 173-98.
** Hirzel, Martin, 2000, " [http://www.research.ibm.com/people/h/hirzel/papers/canon00-goedel.pdf On formally undecidable propositions of Principia Mathematica and related systems I.] ". A modern translation by the author.
* 1951, "Some basic theorems on the foundations of mathematics and their implications" in Solomon Feferman, ed., 1995. "Collected works / Kurt Gödel, Vol. III". Oxford University Press: 304-23.

Translations, during his lifetime, of Gödel’s paper into English

None of the following agree in all translated words and in typography. The typography is a serious matter, because Gödel expressly wished to emphasize “those metamathematical notions that had been defined in their usual sense before . . ."(van Heijenoort 1967:595). Three translations exist. Of the first John Dawson states that: “The Meltzer translation was seriously deficient and received a devastating review in the "Journal of Symbolic Logic"; ”Gödel also complained about Braithwaite’s commentary (Dawson 1997:216). “Fortunately, the Meltzer translation was soon supplanted by a better one prepared by Elliott Mendelson for Martin Davis’s anthology "The Undecidable" . . . he found the translation “not quite so good” as he had expected . . . [but due to time constraints he] agreed to its publication” (ibid). (In a footnote Dawson states that “he would regret his compliance, for the published volume was marred throughout by sloppy typography and numerous misprints” (ibid)). Dawson states that “The translation that Gödel favored was that by Jean van Heijenoort”(ibid). For the serious student another version exists as a set of lecture notes recorded by Stephen Kleene and J. B. Rosser "during lectures given by Gödel at to the Institute for Advanced Study during the spring of 1934" (cf commentary by Davis 1965:39 and beginning on p. 41); this version is titled "On Undecidable Propositions of Formal Mathematical Systems". In their order of publication:

*B. Meltzer (translation) and R. B. Braithwaite (Introduction), 1962. "On Formally Undecidable Propositions of Principia Mathematica and Related Systems", Dover Publications, New York (Dover edition 1992), ISBN 0-486-66980-7 (pbk.) This contains a useful translation of Gödel's German abbreviations on pp.33-34. As noted above, typography, translation and commentary is suspect. Unfortunately, this translation was reprinted with all its suspect content by :*Stephen Hawking editor, 2005. "God Created the Integers: The Mathematical Breakthroughs That Changed History", Running Press, Philadelphia, ISBN-10: 0-7624-1922-9. Gödel’s paper appears starting on p. 1097, with Hawking’s commentary starting on p. 1089.
*Martin Davis editor, 1965. "The Undecidable: Basic Papers on Undecidable Propositions, Unsolvable problems and Computable Functions", Raven Press, New York, no ISBN. Godel’s paper begins on page 5, preceded by one page of commentary.
*Jean van Heijenoort editor, 1967, 3rd edition 1967. "From Frege to Gödel: A Source Book in Mathematical Logic, 1979-1931", Harvard University Press, Cambridge Mass., ISBN 0-674-32449-8 (pbk). van Heijenoort did the translation. He states that “Professor Gödel approved the translation, which in many places was accommodated to his wishes.”(p. 595). Gödel’s paper begins on p. 595; van Heijenoort’s commentary begins on p. 592.

*Martin Davis editor, 1965, ibid. "On Undecidable Propositions of Formal Mathematical Systems." A copy with Gödel's corrections of errata and Gödel's added notes begins on page 41, preceded by two pages of Davis's commentary. Until Davis included this in his volume this lecture existed only as mimeographed notes.

Articles by others

*George Boolos, 1989, "A New Proof of the Gödel Incompleteness Theorem", "Notices of the American Mathematical Society" v. 36, pp. 388–390 and p. 676, reprinted in in Boolos, 1998, "Logic, Logic, and Logic", Harvard Univ. Press. ISBN 0 674 53766 1
*Arthur Charlesworth, 1980, "A Proof of Godel's Theorem in Terms of Computer Programs," "Mathematics Magazine", v. 54 n. 3, pp. 109-121. [http://links.jstor.org/sici?sici=0025-570X%28198105%2954%3A3%3C109%3AAPOGTI%3E2.0.CO%3B2-1&size=LARGE&origin=JSTOR-enlargePage JStor]
* Solomon Feferman, 1984, [http://links.jstor.org/sici?sici=0022-4812(198403)49%3A1%3C75%3ATUTTI%3E2.0.CO;2-D Toward Useful Type-Free Theories, I] , "Journal of Symbolic Logic", v. 49 n. 1, pp. 75–111.
* Geoffrey Hellman, "How to Gödel a Frege-Russell: Gödel's Incompleteness Theorems and Logicism." Noûs, Vol. 15, No. 4, Special Issue on Philosophy of Mathematics. (Nov., 1981), pp. 451–468.
*David Hilbert, 1900, " [http://aleph0.clarku.edu/~djoyce/hilbert/problems.html#prob2 Mathematical Problems.] " English translation of a lecture delivered before the International Congress of Mathematicians at Paris, containing Hilbert's statement of his Second Problem.
* Carl Hewitt, 2008, [http://hewitt-seminars.blogspot.com/2008/03/large-scale-organizational-computing.html Large-scale Organizational Computing requires Unstratified Reflection and Strong Paraconsistency] , "Coordination, Organizations, Institutions, and Norms in Agent Systems III", Springer-Verlag.
* | year=1994 | journal=Notre Dame Journal of Formal Logic | issn=0029-4527 | volume=35 | issue=3 | pages=403–412
* Stephen Cole Kleene, 1943, "Recursive predicates and quantifiers," reprinted from "Transactions of the American Mathemaical Society", v. 53 n. 1, pp. 41–73 in Martin Davis 1965, "The Undecidable" (loc. cit.) pp. 255–287.
*Hilary Putnam, 1960, "Minds and Machines" in Sidney Hook, ed., "Dimensions of Mind: A Symposium". New York University Press. Reprinted in Anderson, A. R., ed., 1964. "Minds and Machines". Prentice-Hall: 77.
* John Barkley Rosser, 1936, "Extensions of some theorems of Gödel and Church," reprinted from the "Journal of Symbolic Logic" vol. 1 (1936) pp. 87-91, in Martin Davis 1965, "The Undecidable" (loc. cit.) pp. 230-235.
* John Barkley Rosser, 1939, "An Informal Exposition of proofs of Gödel's Theorem and Church's Theorem", Reprinted from the "Journal of Symbolic Logic", vol. 4 (1939) pp. 53-60, in Martin Davis 1965, "The Undecidable" (loc. cit.) pp. 223-230
* C. Smoryński, "The incompleteness theorems", in J. Barwise, ed., "Handbook of Mathematical Logic", North-Holland 1977, pp. 821&ndsash;866.
* Jean van Heijenoort, 1963. "Gödel's Theorem" in Edwards, Paul, ed., "Encyclopedia of Philosophy, Vol. 3". Macmillan: 348-57.
* Richard Zach, 2005, "Paper on the incompleteness theorems" in Grattan-Guiness, I., ed., "Landmark Writings in Western Mathematics". Elsevier: 917-25.

Books about the theorems

* Domeisen, Norbert, 1990. "Logik der Antinomien". Bern: Peter Lang. 142 S. 1990. ISBN 3-261-04214-1. [http://www.emis.de/cgi-bin/zmen/ZMATH/en/quick.html?first=1&maxdocs=3&type=html&an=0724.03003&format=complete Zentralblatt MATH]
* Torkel Franzén, 2005. "Gödel's Theorem: An Incomplete Guide to its Use and Abuse". A.K. Peters. ISBN 1568812388 MathSciNet|2007d:03001
* Douglas Hofstadter, 1979. "". Vintage Books. ISBN 0465026850. 1999 reprint: ISBN 0465026567. MathSciNet|80j:03009
* Douglas Hofstadter, 2007. "I Am a Strange Loop". Basic Books. ISBN-13: 9780465030781. ISBN 0465030785. MathSciNet|2008g:00004
* Stanley Jaki, OSB, 2005. "The drama of the quantities". [http://www.realviewbooks.com/ Real View Books.]
* J.R. Lucas, FBA, 1970. "The Freedom of the Will". Clarendon Press, Oxford, 1970.
* Ernest Nagel, James Roy Newman, Douglas R. Hofstadter, 2002 (1958). "Gödel's Proof", revised ed. ISBN 0814758169. MathSciNet|2002i:03001
* Rudy Rucker, 1995 (1982). "Infinity and the Mind: The Science and Philosophy of the Infinite". Princeton Univ. Press. MathSciNet|84d:03012
* Smith, Peter, 2007. " [http://www.godelbook.net/ An Introduction to Gödel's Theorems.] " Cambridge University Press. [http://www.ams.org/mathscinet/search/publdoc.html?arg3=&co4=AND&co5=AND&co6=AND&co7=AND&dr=all&pg4=AUCN&pg5=AUCN&pg6=PC&pg7=ALLF&pg8=ET&s4=Smith%2C%20Peter&s5=&s6=&s7=&s8=All&yearRangeFirst=&yearRangeSecond=&yrop=eq&r=2&mx-pid=2384958 MathSciNet]
*Raymond Smullyan, 1991. "Godel's Incompleteness Theorems". Oxford Univ. Press.
*—, 1994. "Diagonalization and Self-Reference". Oxford Univ. Press. MathSciNet|96c:03001
* Hao Wang, 1997. "A Logical Journey: From Gödel to Philosophy". MIT Press. ISBN 0262231891 MathSciNet|97m:01090

Miscellaneous References

*John W. Dawson, Jr., 1997. "Logical Dilemmas: The Life and Work of Kurt Gödel", A.K. Peters, Wellesley Mass, ISBN 1-56881-256-6.
* John Hopcroft and Jeffrey Ullman 1979, "Introduction to Automata theory", Addison-Wesley, ISBN 0-201-02988-X
*Goldstein, Rebecca, 2005, Incompleteness - The Proof & Paradox of Kurt Godel.

External links

*Stanford Encyclopedia of Philosophy: " [http://plato.stanford.edu/entries/goedel/ Kurt Gödel] " -- by Juliette Kennedy.
*MacTutor biographies:
** [http://www-groups.dcs.st-and.ac.uk/~history/Mathematicians/Godel.html Kurt Gödel.]
** [http://www-groups.dcs.st-and.ac.uk/~history/Mathematicians/Gentzen.html Gerhard Gentzen.]


Wikimedia Foundation. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Gödel's completeness theorem — is a fundamental theorem in mathematical logic that establishes a correspondence between semantic truth and syntactic provability in first order logic. It was first proved by Kurt Gödel in 1929. A first order formula is called logically valid if… …   Wikipedia

  • Gödel, Escher, Bach — Gödel, Escher, Bach: an Eternal Golden Braid   …   Wikipedia

  • Gödel's proof — may refer to: *Gödel s incompleteness theorems *Gödel s ontological proof …   Wikipedia

  • Gödel's theorem — may refer to: *Gödel s incompleteness theorems *Gödel s completeness theorem …   Wikipedia

  • Gödel number — In mathematical logic, a Gödel numbering is a function that assigns to each symbol and well formed formula of some formal language a unique natural number called its Gödel number. The concept was first used by Kurt Gödel for the proof of his… …   Wikipedia

  • Gödel'scher Unvollständigkeitssatz — Der gödelsche Unvollständigkeitssatz ist einer der wichtigsten Sätze der modernen Logik. Er beschäftigt sich mit der Ableitbarkeit von Aussagen in formalen Theorien. Der Satz zeigt die Grenzen der formalen Systeme ab einer bestimmten Mächtigkeit… …   Deutsch Wikipedia

  • Gödel's theorem — n. either of two theorems published by the mathematician Kurt Gödel in 1931 that prove all mathematical systems are incomplete in that their truth or consistency can only be proved using a system of a higher order: also called Gödel s proof or… …   Universalium

  • Gödel's theorem — n. either of two theorems published by the mathematician Kurt Gödel in 1931 that prove all mathematical systems are incomplete in that their truth or consistency can only be proved using a system of a higher order: also called Gödel s proof or… …   English World dictionary

  • Gödel, Kurt — born April 28, 1906, Brünn, Austria Hungary died Jan. 14, 1978, Princeton, N.J., U.S. Austrian born U.S. mathematician and logician. He began his career on the faculty of the University of Vienna, where he produced his groundbreaking proof (see… …   Universalium

  • Gödel metric — The Gödel metric is an exact solution of the Einstein field equations in which the stress energy tensor contains two terms, the first representing the matter density of a homogeneous distribution of swirling dust particles, and the second… …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”