Cardinality

Last updated
The set
S
{\displaystyle S}
of all Platonic solids has 5 elements. Thus the cardinality of
S
{\displaystyle S}
is 5 or, in symbols,
|
S
|
=
5
{\displaystyle |S|=5}
. Platonic Solids Transparent.svg
The set of all Platonic solids has 5 elements. Thus the cardinality of is 5 or, in symbols, .

In mathematics, the cardinality of a set is a measure of the number of elements of the set. For example, the set contains 3 elements, and therefore has a cardinality of 3. Beginning in the late 19th century, this concept was generalized to infinite sets, which allows one to distinguish between different types of infinity, and to perform arithmetic on them. There are two approaches to cardinality: one which compares sets directly using bijections and injections, and another which uses cardinal numbers. [1] The cardinality of a set may also be called its size, when no confusion with other notions of size [2] is possible.

Contents

The cardinality of a set is usually denoted , with a vertical bar on each side; [3] this is the same notation as absolute value, and the meaning depends on context. The cardinality of a set may alternatively be denoted by , , , or .

History

A crude sense of cardinality, an awareness that groups of things or events compare with other groups by containing more, fewer, or the same number of instances, is observed in a variety of present-day animal species, suggesting an origin millions of years ago. [4] Human expression of cardinality is seen as early as 40000 years ago, with equating the size of a group with a group of recorded notches, or a representative collection of other things, such as sticks and shells. [5] The abstraction of cardinality as a number is evident by 3000 BCE, in Sumerian mathematics and the manipulation of numbers without reference to a specific group of things or events. [6]

From the 6th century BCE, the writings of Greek philosophers show hints of the cardinality of infinite sets. While they considered the notion of infinity as an endless series of actions, such as adding 1 to a number repeatedly, they did not consider the size of an infinite set of numbers to be a thing. [7] The ancient Greek notion of infinity also considered the division of things into parts repeated without limit. In Euclid's Elements , commensurability was described as the ability to compare the length of two line segments, a and b, as a ratio, as long as there were a third segment, no matter how small, that could be laid end-to-end a whole number of times into both a and b. But with the discovery of irrational numbers, it was seen that even the infinite set of all rational numbers was not enough to describe the length of every possible line segment. [8] Still, there was no concept of infinite sets as something that had cardinality.

To better understand infinite sets, a notion of cardinality was formulated circa 1880 by Georg Cantor, the originator of set theory. He examined the process of equating two sets with bijection, a one-to-one correspondence between the elements of two sets based on a unique relationship. In 1891, with the publication of Cantor's diagonal argument, he demonstrated that there are sets of numbers that cannot be placed in one-to-one correspondence with the set of natural numbers, i.e. uncountable sets that contain more elements than there are in the infinite set of natural numbers. [9]

Comparing sets

Bijective function from N to the set E of even numbers. Although E is a proper subset of N, both sets have the same cardinality. Aplicacion 2 inyectiva sobreyectiva04.svg
Bijective function from N to the set E of even numbers. Although E is a proper subset of N, both sets have the same cardinality.
N does not have the same cardinality as its power set P(N): For every function f from N to P(N), the set T = {n[?]N: n[?]f(n)} disagrees with every set in the range of f, hence f cannot be surjective. The picture shows an example f and the corresponding T; red: n[?]f(n)\T, blue:n[?]T\f(n). Diagonal argument powerset svg.svg
N does not have the same cardinality as its power set P(N): For every function f from N to P(N), the set T = {nN: nf(n)} disagrees with every set in the range of f, hence f cannot be surjective. The picture shows an example f and the corresponding T; red: nf(n)\T, blue:nT\f(n).

While the cardinality of a finite set is just the number of its elements, extending the notion to infinite sets usually starts with defining the notion of comparison of arbitrary sets (some of which are possibly infinite).

Definition 1: |A| = |B|

Two sets A and B have the same cardinality if there exists a bijection (a.k.a., one-to-one correspondence) from A to B, [10] that is, a function from A to B that is both injective and surjective. Such sets are said to be equipotent, equipollent, or equinumerous . This relationship can also be denoted AB or A ~ B.
For example, the set E = {0, 2, 4, 6, ...} of non-negative even numbers has the same cardinality as the set N = {0, 1, 2, 3, ...} of natural numbers, since the function f(n) = 2n is a bijection from N to E (see picture).
For finite sets A and B, if some bijection exists from A to B, then each injective or surjective function from A to B is a bijection. This is no longer true for infinite A and B. For example, the function g from N to E, defined by g(n) = 4n is injective, but not surjective, and h from N to E, defined by h(n) = n - (n mod 2) is surjective, but not injective. Neither g nor h can challenge |E| = |N|, which was established by the existence of f.

Definition 2: |A||B|

A has cardinality less than or equal to the cardinality of B, if there exists an injective function from A into B.

Definition 3: |A| < |B|

A has cardinality strictly less than the cardinality of B, if there is an injective function, but no bijective function, from A to B.
For example, the set N of all natural numbers has cardinality strictly less than its power set P(N), because g(n) = { n } is an injective function from N to P(N), and it can be shown that no function from N to P(N) can be bijective (see picture). By a similar argument, N has cardinality strictly less than the cardinality of the set R of all real numbers. For proofs, see Cantor's diagonal argument or Cantor's first uncountability proof.

If |A||B| and |B||A|, then |A| = |B| (a fact known as Schröder–Bernstein theorem). The axiom of choice is equivalent to the statement that |A||B| or |B||A| for every A, B. [11] [12]

Cardinal numbers

In the above section, "cardinality" of a set was defined functionally. In other words, it was not defined as a specific object itself. However, such an object can be defined as follows.

The relation of having the same cardinality is called equinumerosity, and this is an equivalence relation on the class of all sets. The equivalence class of a set A under this relation, then, consists of all those sets which have the same cardinality as A. There are two ways to define the "cardinality of a set":

  1. The cardinality of a set A is defined as its equivalence class under equinumerosity.
  2. A representative set is designated for each equivalence class. The most common choice is the initial ordinal in that class. This is usually taken as the definition of cardinal number in axiomatic set theory.

Assuming the axiom of choice, the cardinalities of the infinite sets are denoted

For each ordinal , is the least cardinal number greater than .

The cardinality of the natural numbers is denoted aleph-null (), while the cardinality of the real numbers is denoted by "" (a lowercase fraktur script "c"), and is also referred to as the cardinality of the continuum. Cantor showed, using the diagonal argument, that . We can show that , this also being the cardinality of the set of all subsets of the natural numbers.

The continuum hypothesis says that , i.e. is the smallest cardinal number bigger than , i.e. there is no set whose cardinality is strictly between that of the integers and that of the real numbers. The continuum hypothesis is independent of ZFC, a standard axiomatization of set theory; that is, it is impossible to prove the continuum hypothesis or its negation from ZFC—provided that ZFC is consistent. For more detail, see § Cardinality of the continuum below. [13] [14] [15]

Finite, countable and uncountable sets

If the axiom of choice holds, the law of trichotomy holds for cardinality. Thus we can make the following definitions:

Infinite sets

Our intuition gained from finite sets breaks down when dealing with infinite sets. In the late 19th century Georg Cantor, Gottlob Frege, Richard Dedekind and others rejected the view that the whole cannot be the same size as the part. [16] [ citation needed ] One example of this is Hilbert's paradox of the Grand Hotel. Indeed, Dedekind defined an infinite set as one that can be placed into a one-to-one correspondence with a strict subset (that is, having the same size in Cantor's sense); this notion of infinity is called Dedekind infinite. Cantor introduced the cardinal numbers, and showed—according to his bijection-based definition of size—that some infinite sets are greater than others. The smallest infinite cardinality is that of the natural numbers ().

Cardinality of the continuum

One of Cantor's most important results was that the cardinality of the continuum () is greater than that of the natural numbers (); that is, there are more real numbers R than natural numbers N. Namely, Cantor showed that (see Beth one) satisfies:

(see Cantor's diagonal argument or Cantor's first uncountability proof).

The continuum hypothesis states that there is no cardinal number between the cardinality of the reals and the cardinality of the natural numbers, that is,

However, this hypothesis can neither be proved nor disproved within the widely accepted ZFC axiomatic set theory, if ZFC is consistent.

Cardinal arithmetic can be used to show not only that the number of points in a real number line is equal to the number of points in any segment of that line, but that this is equal to the number of points on a plane and, indeed, in any finite-dimensional space. These results are highly counterintuitive, because they imply that there exist proper subsets and proper supersets of an infinite set S that have the same size as S, although S contains elements that do not belong to its subsets, and the supersets of S contain elements that are not included in it.

The first of these results is apparent by considering, for instance, the tangent function, which provides a one-to-one correspondence between the interval (−½π, ½π) and R (see also Hilbert's paradox of the Grand Hotel).

The second result was first demonstrated by Cantor in 1878, but it became more apparent in 1890, when Giuseppe Peano introduced the space-filling curves, curved lines that twist and turn enough to fill the whole of any square, or cube, or hypercube, or finite-dimensional space. These curves are not a direct proof that a line has the same number of points as a finite-dimensional space, but they can be used to obtain such a proof.

Cantor also showed that sets with cardinality strictly greater than exist (see his generalized diagonal argument and theorem). They include, for instance:

  • the set of all subsets of R, i.e., the power set of R, written P(R) or 2R
  • the set RR of all functions from R to R

Both have cardinality

(see Beth two).

The cardinal equalities and can be demonstrated using cardinal arithmetic:

Examples and properties

Union and intersection

If A and B are disjoint sets, then

From this, one can show that in general, the cardinalities of unions and intersections are related by the following equation: [17]

See also

Related Research Articles

In mathematics, specifically set theory, the continuum hypothesis is a hypothesis about the possible sizes of infinite sets. It states that

there is no set whose cardinality is strictly between that of the integers and the real numbers,

In mathematics, a set is countable if either it is finite or it can be made in one to one correspondence with the set of natural numbers. Equivalently, a set is countable if there exists an injective function from it into the natural numbers; this means that each element in the set may be associated to a unique natural number, or that the elements of the set can be counted one at a time, although the counting may never finish due to an infinite number of elements.

<span class="mw-page-title-main">Cardinal number</span> Size of a possibly infinite set

In mathematics, a cardinal number, or cardinal for short, is what is commonly called the number of elements of a set. In the case of a finite set, its cardinal number, or cardinality is therefore a natural number. For dealing with the case of infinite sets, the infinite cardinal numbers have been introduced, which are often denoted with the Hebrew letter (aleph) marked with subscript indicating their rank among the infinite cardinals.

In mathematics, especially in order theory, the cofinality cf(A) of a partially ordered set A is the least of the cardinalities of the cofinal subsets of A.

In mathematics, an uncountable set, informally, is an infinite set that contains too many elements to be countable. The uncountability of a set is closely related to its cardinal number: a set is uncountable if its cardinal number is larger than aleph-null, the cardinality of the natural numbers.

<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.

In mathematics, transfinite numbers or infinite numbers are numbers that are "infinite" in the sense that they are larger than all finite numbers. These include the transfinite cardinals, which are cardinal numbers used to quantify the size of infinite sets, and the transfinite ordinals, which are ordinal numbers used to provide an ordering of infinite sets. The term transfinite was coined in 1895 by Georg Cantor, who wished to avoid some of the implications of the word infinite in connection with these objects, which were, nevertheless, not finite. Few contemporary writers share these qualms; it is now accepted usage to refer to transfinite cardinals and ordinals as infinite numbers. Nevertheless, the term transfinite also remains in use.

<span class="mw-page-title-main">Aleph number</span> Infinite cardinal number

In mathematics, particularly in set theory, the aleph numbers are a sequence of numbers used to represent the cardinality of infinite sets that can be well-ordered. They were introduced by the mathematician Georg Cantor and are named after the symbol he used to denote them, the Hebrew letter aleph.

In mathematics, two sets or classes A and B are equinumerous if there exists a one-to-one correspondence (or bijection) between them, that is, if there exists a function from A to B such that for every element y of B, there is exactly one element x of A with f(x) = y. Equinumerous sets are said to have the same cardinality (number of elements). The study of cardinality is often called equinumerosity (equalness-of-number). The terms equipollence (equalness-of-strength) and equipotence (equalness-of-power) are sometimes used instead.

In set theory, a regular cardinal is a cardinal number that is equal to its own cofinality. More explicitly, this means that is a regular cardinal if and only if every unbounded subset has cardinality . Infinite well-ordered cardinals that are not regular are called singular cardinals. Finite cardinal numbers are typically not called regular or singular.

In mathematics, particularly in set theory, the beth numbers are a certain sequence of infinite cardinal numbers, conventionally written , where is the Hebrew letter beth. The beth numbers are related to the aleph numbers, but unless the generalized continuum hypothesis is true, there are numbers indexed by that are not indexed by .

In set theory, the cardinality of the continuum is the cardinality or "size" of the set of real numbers , sometimes called the continuum. It is an infinite cardinal number and is denoted by or .

In mathematics, set-theoretic topology is a subject that combines set theory and general topology. It focuses on topological questions that are independent of Zermelo–Fraenkel set theory (ZFC).

In set theory, Cichoń's diagram or Cichon's diagram is a table of 10 infinite cardinal numbers related to the set theory of the reals displaying the provable relations between these cardinal characteristics of the continuum. All these cardinals are greater than or equal to , the smallest uncountable cardinal, and they are bounded above by , the cardinality of the continuum. Four cardinals describe properties of the ideal of sets of measure zero; four more describe the corresponding properties of the ideal of meager sets.

This article contains a discussion of paradoxes of set theory. As with most mathematical paradoxes, they generally reveal surprising and counter-intuitive mathematical results, rather than actual logical contradictions within modern axiomatic set theory.

Pocket set theory (PST) is an alternative set theory in which there are only two infinite cardinal numbers, ℵ0 and c. The theory was first suggested by Rudy Rucker in his Infinity and the Mind. The details set out in this entry are due to the American mathematician Randall M. Holmes.

In the mathematical field of set theory, the continuum means the real numbers, or the corresponding (infinite) cardinal number, denoted by . Georg Cantor proved that the cardinality is larger than the smallest infinity, namely, . He also proved that is equal to , the cardinality of the power set of the natural numbers.

<span class="mw-page-title-main">Ordinal number</span> Generalization of "n-th" to infinite cases

In set theory, an ordinal number, or ordinal, is a generalization of ordinal numerals aimed to extend enumeration to infinite sets.

In the mathematical discipline of set theory, a cardinal characteristic of the continuum is an infinite cardinal number that may consistently lie strictly between , and the cardinality of the continuum, that is, the cardinality of the set of all real numbers. The latter cardinal is denoted or . A variety of such cardinal characteristics arise naturally, and much work has been done in determining what relations between them are provable, and constructing models of set theory for various consistent configurations of them.

In mathematics a group is a set together with a binary operation on the set called multiplication that obeys the group axioms. The axiom of choice is an axiom of ZFC set theory which in one form states that every set can be wellordered.

References

  1. Weisstein, Eric W. "Cardinal Number". MathWorld .
  2. Such as length and area in geometry. – A line of finite length is a set of points that has infinite cardinality.
  3. "Cardinality | Brilliant Math & Science Wiki". brilliant.org. Retrieved 2020-08-23.
  4. Cepelewicz, Jordana Animals Count and Use Zero. How Far Does Their Number Sense Go? , Quanta, August 9, 2021
  5. "Early Human Counting Tools". Math Timeline. Retrieved 2018-04-26.
  6. Duncan J. Melville (2003). Third Millennium Chronology Archived 2018-07-07 at the Wayback Machine , Third Millennium Mathematics. St. Lawrence University.
  7. Allen, Donald (2003). "The History of Infinity" (PDF). Texas A&M Mathematics. Archived from the original (PDF) on August 1, 2020. Retrieved Nov 15, 2019.
  8. Kurt Von Fritz (1945). "The Discovery of Incommensurability by Hippasus of Metapontum". The Annals of Mathematics.
  9. Georg Cantor (1891). "Ueber eine elementare Frage der Mannigfaltigkeitslehre" (PDF). Jahresbericht der Deutschen Mathematiker-Vereinigung. 1: 75–78.
  10. 1 2 "Infinite Sets and Cardinality". Mathematics LibreTexts. 2019-12-05. Retrieved 2020-08-23.
  11. Friedrich M. Hartogs (1915), Felix Klein; Walther von Dyck; David Hilbert; Otto Blumenthal (eds.), "Über das Problem der Wohlordnung", Mathematische Annalen , Leipzig: B. G. Teubner, 76 (4): 438–443, doi:10.1007/bf01458215, ISSN   0025-5831, S2CID   121598654
  12. Felix Hausdorff (2002), Egbert Brieskorn; Srishti D. Chatterji; et al. (eds.), Grundzüge der Mengenlehre (1. ed.), Berlin/Heidelberg: Springer, p. 587, ISBN   3-540-42224-2 - Original edition (1914)
  13. Cohen, Paul J. (December 15, 1963). "The Independence of the Continuum Hypothesis". Proceedings of the National Academy of Sciences of the United States of America. 50 (6): 1143–1148. Bibcode:1963PNAS...50.1143C. doi: 10.1073/pnas.50.6.1143 . JSTOR   71858. PMC   221287 . PMID   16578557.
  14. Cohen, Paul J. (January 15, 1964). "The Independence of the Continuum Hypothesis, II". Proceedings of the National Academy of Sciences of the United States of America. 51 (1): 105–110. Bibcode:1964PNAS...51..105C. doi: 10.1073/pnas.51.1.105 . JSTOR   72252. PMC   300611 . PMID   16591132.
  15. Penrose, R (2005), The Road to Reality: A Complete guide to the Laws of the Universe, Vintage Books, ISBN   0-09-944068-7
  16. Georg Cantor (1887), "Mitteilungen zur Lehre vom Transfiniten", Zeitschrift für Philosophie und philosophische Kritik , 91: 81–125
    Reprinted in: Georg Cantor (1932), Adolf Fraenkel (Lebenslauf); Ernst Zermelo (eds.), Gesammelte Abhandlungen mathematischen und philosophischen Inhalts, Berlin: Springer, pp. 378–439 Here: p.413 bottom
  17. Applied Abstract Algebra, K.H. Kim, F.W. Roush, Ellis Horwood Series, 1983, ISBN   0-85312-612-7 (student edition), ISBN   0-85312-563-5 (library edition)