Cantor's theorem

Last updated
The cardinality of the set {x, y, z}, is three, while there are eight elements in its power set (3 < 2 = 8), here ordered by inclusion. Hasse diagram of powerset of 3.svg
The cardinality of the set {x, y, z}, is three, while there are eight elements in its power set (3 < 2 = 8), here ordered by inclusion.

In mathematical set theory, Cantor's theorem is a fundamental result which states that, for any set , the set of all subsets of known as the power set of has a strictly greater cardinality than itself.

Contents

For finite sets, Cantor's theorem can be seen to be true by simple enumeration of the number of subsets. Counting the empty set as a subset, a set with elements has a total of subsets, and the theorem holds because for all non-negative integers.

Much more significant is Cantor's discovery of an argument that is applicable to any set, and shows that the theorem holds for infinite sets also. As a consequence, the cardinality of the real numbers, which is the same as that of the power set of the integers, is strictly larger than the cardinality of the integers; see Cardinality of the continuum for details.

The theorem is named for German mathematician Georg Cantor, who first stated and proved it at the end of the 19th century. Cantor's theorem had immediate and important consequences for the philosophy of mathematics. For instance, by iteratively taking the power set of an infinite set and applying Cantor's theorem, we obtain an endless hierarchy of infinite cardinals, each strictly larger than the one before it. Consequently, the theorem implies that there is no largest cardinal number (colloquially, "there's no largest infinity").

Proof

Cantor's argument is elegant and remarkably simple. The complete proof is presented below, with detailed explanations to follow.

Theorem (Cantor)  Let be a map from set to its power set . Then is not surjective. As a consequence, holds for any set .

Proof

Consider the set . Suppose to the contrary that is surjective. Then there exists such that . But by construction, . This is a contradiction. Thus, cannot be surjective. On the other hand, defined by is an injective map. Consequently, we must have . Q.E.D.

By definition of cardinality, we have for any two sets and if and only if there is an injective function but no bijective function from to . It suffices to show that there is no surjection from to . This is the heart of Cantor's theorem: there is no surjective function from any set to its power set. To establish this, it is enough to show that no function that maps elements in to subsets of can reach every possible subset, i.e., we just need to demonstrate the existence of a subset of that is not equal to for any . (Recall that each is a subset of .) Such a subset is given by the following construction, sometimes called the Cantor diagonal set of : [1] [2]

This means, by definition, that for all , if and only if . For all the sets and cannot be equal because was constructed from elements of whose images under did not include themselves. For all either or . If then cannot equal because by assumption and by definition. If then cannot equal because by assumption and by the definition of .

Equivalently, and slightly more formally, we have just proved that the existence of such that implies the following contradiction:

Therefore, by reductio ad absurdum, the assumption must be false. [3] Thus there is no such that  ; in other words, is not in the image of and does not map onto every element of the power set of , i.e., is not surjective.

Finally, to complete the proof, we need to exhibit an injective function from to its power set. Finding such a function is trivial: just map to the singleton set . The argument is now complete, and we have established the strict inequality for any set that .

Another way to think of the proof is that , empty or non-empty, is always in the power set of . For to be onto, some element of must map to . But that leads to a contradiction: no element of can map to because that would contradict the criterion of membership in , thus the element mapping to must not be an element of meaning that it satisfies the criterion for membership in , another contradiction. So the assumption that an element of maps to must be false; and cannot be onto.

Because of the double occurrence of in the expression "", this is a diagonal argument. For a countable (or finite) set, the argument of the proof given above can be illustrated by constructing a table in which each row is labelled by a unique from , in this order. is assumed to admit a linear order so that such table can be constructed. Each column of the table is labelled by a unique from the power set of ; the columns are ordered by the argument to , i.e. the column labels are , ..., in this order. The intersection of each row and column records a true/false bit whether . Given the order chosen for the row and column labels, the main diagonal of this table thus records whether for each . The set constructed in the previous paragraphs coincides with the row labels for the subset of entries on this main diagonal where the table records that is false. [3] Each column records the values of the indicator function of the set corresponding to the column. The indicator function of coincides with the logically negated (swap "true" and "false") entries of the main diagonal. Thus the indicator function of does not agree with any column in at least one entry. Consequently, no column represents .

Despite the simplicity of the above proof, it is rather difficult for an automated theorem prover to produce it. The main difficulty lies in an automated discovery of the Cantor diagonal set. Lawrence Paulson noted in 1992 that Otter could not do it, whereas Isabelle could, albeit with a certain amount of direction in terms of tactics that might perhaps be considered cheating. [2]

When A is countably infinite

Let us examine the proof for the specific case when is countably infinite. Without loss of generality, we may take , the set of natural numbers.

Suppose that is equinumerous with its power set . Let us see a sample of what looks like:

contains infinite subsets of , e.g. the set of all positive even numbers , along with the empty set .

Now that we have an idea of what the elements of are, let us attempt to pair off each element of with each element of to show that these infinite sets are equinumerous. In other words, we will attempt to pair off each element of with an element from the infinite set , so that no element from either infinite set remains unpaired. Such an attempt to pair elements would look like this:

Given such a pairing, some natural numbers are paired with subsets that contain the very same number. For instance, in our example the number 2 is paired with the subset {1, 2, 3}, which contains 2 as a member. Let us call such numbers selfish. Other natural numbers are paired with subsets that do not contain them. For instance, in our example the number 1 is paired with the subset {4, 5}, which does not contain the number 1. Call these numbers non-selfish. Likewise, 3 and 4 are non-selfish.

Using this idea, let us build a special set of natural numbers. This set will provide the contradiction we seek. Let be the set of all non-selfish natural numbers. By definition, the power set contains all sets of natural numbers, and so it contains this set as an element. If the mapping is bijective, must be paired off with some natural number, say . However, this causes a problem. If is in , then is selfish because it is in the corresponding set, which contradicts the definition of . If is not in , then it is non-selfish and it should instead be a member of . Therefore, no such element which maps to can exist.

Since there is no natural number which can be paired with , we have contradicted our original supposition, that there is a bijection between and .

Note that the set may be empty. This would mean that every natural number maps to a subset of natural numbers that contains . Then, every number maps to a nonempty set and no number maps to the empty set. But the empty set is a member of , so the mapping still does not cover .

Through this proof by contradiction we have proven that the cardinality of and cannot be equal. We also know that the cardinality of cannot be less than the cardinality of because contains all singletons, by definition, and these singletons form a "copy" of inside of . Therefore, only one possibility remains, and that is that the cardinality of is strictly greater than the cardinality of , proving Cantor's theorem.

Cantor's theorem and its proof are closely related to two paradoxes of set theory.

Cantor's paradox is the name given to a contradiction following from Cantor's theorem together with the assumption that there is a set containing all sets, the universal set . In order to distinguish this paradox from the next one discussed below, it is important to note what this contradiction is. By Cantor's theorem for any set . On the other hand, all elements of are sets, and thus contained in , therefore . [1]

Another paradox can be derived from the proof of Cantor's theorem by instantiating the function f with the identity function; this turns Cantor's diagonal set into what is sometimes called the Russell set of a given set A: [1]

The proof of Cantor's theorem is straightforwardly adapted to show that assuming a set of all sets U exists, then considering its Russell set RU leads to the contradiction:

This argument is known as Russell's paradox. [1] As a point of subtlety, the version of Russell's paradox we have presented here is actually a theorem of Zermelo; [4] we can conclude from the contradiction obtained that we must reject the hypothesis that RUU, thus disproving the existence of a set containing all sets. This was possible because we have used restricted comprehension (as featured in ZFC) in the definition of RA above, which in turn entailed that

Had we used unrestricted comprehension (as in Frege's system for instance) by defining the Russell set simply as , then the axiom system itself would have entailed the contradiction, with no further hypotheses needed. [4]

Despite the syntactical similarities between the Russell set (in either variant) and the Cantor diagonal set, Alonzo Church emphasized that Russell's paradox is independent of considerations of cardinality and its underlying notions like one-to-one correspondence. [5]

History

Cantor gave essentially this proof in a paper published in 1891 "Über eine elementare Frage der Mannigfaltigkeitslehre", [6] where the diagonal argument for the uncountability of the reals also first appears (he had earlier proved the uncountability of the reals by other methods). The version of this argument he gave in that paper was phrased in terms of indicator functions on a set rather than subsets of a set. [7] He showed that if f is a function defined on X whose values are 2-valued functions on X, then the 2-valued function G(x) = 1 f(x)(x) is not in the range of f.

Bertrand Russell has a very similar proof in Principles of Mathematics (1903, section 348), where he shows that there are more propositional functions than objects. "For suppose a correlation of all objects and some propositional functions to have been affected, and let phi-x be the correlate of x. Then "not-phi-x(x)," i.e. "phi-x does not hold of x" is a propositional function not contained in this correlation; for it is true or false of x according as phi-x is false or true of x, and therefore it differs from phi-x for every value of x." He attributes the idea behind the proof to Cantor.

Ernst Zermelo has a theorem (which he calls "Cantor's Theorem") that is identical to the form above in the paper that became the foundation of modern set theory ("Untersuchungen über die Grundlagen der Mengenlehre I"), published in 1908. See Zermelo set theory.

Generalizations

Lawvere's fixed-point theorem provides for a broad generalization of Cantor's theorem to any category with finite products in the following way: [8] let be such a category, and let be a terminal object in . Suppose that is an object in and that there exists an endomorphism that does not have any fixed points; that is, there is no morphism that satisfies . Then there is no object of such that a morphism can parameterize all morphisms . In other words, for every object and every morphism , an attempt to write maps as maps of the form must leave out at least one map .

See also

Related Research Articles

In mathematics, a paracompact space is a topological space in which every open cover has an open refinement that is locally finite. These spaces were introduced by Dieudonné (1944). Every compact space is paracompact. Every paracompact Hausdorff space is normal, and a Hausdorff space is paracompact if and only if it admits partitions of unity subordinate to any open cover. Sometimes paracompact spaces are defined so as to always be Hausdorff.

<span class="mw-page-title-main">Cantor's diagonal argument</span> Proof in set theory

In set theory, Cantor's diagonal argument, also called the diagonalisation argument, the diagonal slash argument, the anti-diagonal argument, the diagonal method, and Cantor's diagonalization proof, was published in 1891 by Georg Cantor as a mathematical proof that there are infinite sets which cannot be put into one-to-one correspondence with the infinite set of natural numbers. Such sets are now known as uncountable sets, and the size of infinite sets is now treated by the theory of cardinal numbers which Cantor began.

<span class="mw-page-title-main">Fourier transform</span> Mathematical transform that expresses a function of time as a function of frequency

In physics, engineering and mathematics, the Fourier transform (FT) is an integral transform that takes as input a function and outputs another function that describes the extent to which various frequencies are present in the original function. The output of the transform is a complex-valued function of frequency. The term Fourier transform refers to both this complex-valued function and the mathematical operation. When a distinction needs to be made the Fourier transform is sometimes called the frequency domain representation of the original function. The Fourier transform is analogous to decomposing the sound of a musical chord into the intensities of its constituent pitches.

In mathematics, a base (or basis; pl.: bases) for the topology τ of a topological space (X, τ) is a family of open subsets of X such that every open set of the topology is equal to the union of some sub-family of . For example, the set of all open intervals in the real number line is a basis for the Euclidean topology on because every open interval is an open set, and also every open subset of can be written as a union of some family of open intervals.

In the mathematical discipline of set theory, forcing is a technique for proving consistency and independence results. Intuitively, forcing can be thought of as a technique to expand the set theoretical universe to a larger universe by introducing a new "generic" object .

In set theory, Zermelo–Fraenkel set theory, named after mathematicians Ernst Zermelo and Abraham Fraenkel, is an axiomatic system that was proposed in the early twentieth century in order to formulate a theory of sets free of paradoxes such as Russell's paradox. Today, Zermelo–Fraenkel set theory, with the historically controversial axiom of choice (AC) included, is the standard form of axiomatic set theory and as such is the most common foundation of mathematics. Zermelo–Fraenkel set theory with the axiom of choice included is abbreviated ZFC, where C stands for "choice", and ZF refers to the axioms of Zermelo–Fraenkel set theory with the axiom of choice excluded.

Freiling's axiom of symmetry is a set-theoretic axiom proposed by Chris Freiling. It is based on intuition of Stuart Davidson but the mathematics behind it goes back to Wacław Sierpiński.

In mathematics, the adele ring of a global field is a central object of class field theory, a branch of algebraic number theory. It is the restricted product of all the completions of the global field and is an example of a self-dual topological ring.

In mathematics, the Fourier inversion theorem says that for many types of functions it is possible to recover a function from its Fourier transform. Intuitively it may be viewed as the statement that if we know all frequency and phase information about a wave then we may reconstruct the original wave precisely.

In mathematics, a join-semilattice is a partially ordered set that has a join for any nonempty finite subset. Dually, a meet-semilattice is a partially ordered set which has a meet for any nonempty finite subset. Every join-semilattice is a meet-semilattice in the inverse order and vice versa.

In functional analysis, a branch of mathematics, the Borel functional calculus is a functional calculus, which has particularly broad scope. Thus for instance if T is an operator, applying the squaring function ss2 to T yields the operator T2. Using the functional calculus for larger classes of functions, we can for example define rigorously the "square root" of the (negative) Laplacian operator −Δ or the exponential

In statistics and probability theory, a point process or point field is a collection of mathematical points randomly located on a mathematical space such as the real line or Euclidean space. Point processes can be used for spatial data analysis, which is of interest in such diverse disciplines as forestry, plant ecology, epidemiology, geography, seismology, materials science, astronomy, telecommunications, computational neuroscience, economics and others.

In mathematical logic, the theory of infinite sets was first developed by Georg Cantor. Although this work has become a thoroughly standard fixture of classical set theory, it has been criticized in several areas by mathematicians and philosophers.

In mathematics, a π-system on a set is a collection of certain subsets of such that

In commutative algebra, an element b of a commutative ring B is said to be integral over a subring A of B if b is a root of some monic polynomial over A.

In mathematics, a filter on a set is a family of subsets such that:

  1. and
  2. if and , then
  3. If , and , then

In mathematics, the FBI transform or Fourier–Bros–Iagolnitzer transform is a generalization of the Fourier transform developed by the French mathematical physicists Jacques Bros and Daniel Iagolnitzer in order to characterise the local analyticity of functions on Rn. The transform provides an alternative approach to analytic wave front sets of distributions, developed independently by the Japanese mathematicians Mikio Sato, Masaki Kashiwara and Takahiro Kawai in their approach to microlocal analysis. It can also be used to prove the analyticity of solutions of analytic elliptic partial differential equations as well as a version of the classical uniqueness theorem, strengthening the Cauchy–Kowalevski theorem, due to the Swedish mathematician Erik Albert Holmgren (1872–1943).

In mathematics, a polyadic space is a topological space that is the image under a continuous function of a topological power of an Alexandroff one-point compactification of a discrete space.

In mathematics, particularly in functional analysis and convex analysis, the Ursescu theorem is a theorem that generalizes the closed graph theorem, the open mapping theorem, and the uniform boundedness principle.

<span class="mw-page-title-main">Ultrafilter on a set</span> Maximal proper filter

In the mathematical field of set theory, an ultrafilter on a set is a maximal filter on the set In other words, it is a collection of subsets of that satisfies the definition of a filter on and that is maximal with respect to inclusion, in the sense that there does not exist a strictly larger collection of subsets of that is also a filter. Equivalently, an ultrafilter on the set can also be characterized as a filter on with the property that for every subset of either or its complement belongs to the ultrafilter.

References

  1. 1 2 3 4 Abhijit Dasgupta (2013). Set Theory: With an Introduction to Real Point Sets. Springer Science & Business Media. pp. 362–363. ISBN   978-1-4614-8854-5.
  2. 1 2 Lawrence Paulson (1992). Set Theory as a Computational Logic (PDF). University of Cambridge Computer Laboratory. p. 14.
  3. 1 2 Graham Priest (2002). Beyond the Limits of Thought. Oxford University Press. pp. 118–119. ISBN   978-0-19-925405-7.
  4. 1 2 Heinz-Dieter Ebbinghaus (2007). Ernst Zermelo: An Approach to His Life and Work . Springer Science & Business Media. pp.  86–87. ISBN   978-3-540-49553-6.
  5. Church, A. [1974] "Set theory with a universal set." in Proceedings of the Tarski Symposium. Proceedings of Symposia in Pure Mathematics XXV, ed. L. Henkin, Providence RI, Second printing with additions 1979, pp. 297−308. ISBN   978-0-8218-7360-1. Also published in International Logic Review 15 pp. 11−23.
  6. Cantor, Georg (1891), "Über eine elementare Frage der Mannigfaltigskeitslehre", Jahresbericht der Deutschen Mathematiker-Vereinigung (in German), 1: 75–78, also in Georg Cantor, Gesammelte Abhandlungen mathematischen und philosophischen Inhalts, E. Zermelo, 1932.
  7. A. Kanamori, "The Empty Set, the Singleton, and the Ordered Pair", p.276. Bulletin of Symbolic Logic vol. 9, no. 3, (2003). Accessed 21 August 2023.
  8. F. William Lawvere; Stephen H. Schanuel (2009). Conceptual Mathematics: A First Introduction to Categories . Cambridge University Press. Session 29. ISBN   978-0-521-89485-2.