Set (Mathematics)
Set (Mathematics)
en.wikipedia.org
Chapter 1
For an introduction to the subject, see Boolean algebra. For an alternative presentation, see Boolean algebras canon-
ically dened.
In abstract algebra, a Boolean algebra or Boolean lattice is a complemented distributive lattice. This type of
algebraic structure captures essential properties of both set operations and logic operations. A Boolean algebra can
be seen as a generalization of a power set algebra or a eld of sets, or its elements can be viewed as generalized truth
values. It is also a special case of a De Morgan algebra and a Kleene algebra (with involution).
Every Boolean algebra gives rise to a Boolean ring, and vice versa, with ring multiplication corresponding to conjunction
or meet , and ring addition to exclusive disjunction or symmetric dierence (not disjunction ). However, the theory
of Boolean rings has an inherent asymmetry between the two operators, while the axioms and theorems of Boolean
algebra express the symmetry of the theory described by the duality principle.[1]
{x,y,z}
2
1.1. HISTORY 3
1.1 History
The term Boolean algebra honors George Boole (18151864), a self-educated English mathematician. He intro-
duced the algebraic system initially in a small pamphlet, The Mathematical Analysis of Logic, published in 1847 in
response to an ongoing public controversy between Augustus De Morgan and William Hamilton, and later as a more
substantial book, The Laws of Thought, published in 1854. Booles formulation diers from that described above
in some important respects. For example, conjunction and disjunction in Boole were not a dual pair of operations.
Boolean algebra emerged in the 1860s, in papers written by William Jevons and Charles Sanders Peirce. The rst sys-
tematic presentation of Boolean algebra and distributive lattices is owed to the 1890 Vorlesungen of Ernst Schrder.
The rst extensive treatment of Boolean algebra in English is A. N. Whitehead's 1898 Universal Algebra. Boolean
algebra as an axiomatic algebraic structure in the modern axiomatic sense begins with a 1904 paper by Edward V.
Huntington. Boolean algebra came of age as serious mathematics with the work of Marshall Stone in the 1930s,
and with Garrett Birkho's 1940 Lattice Theory. In the 1960s, Paul Cohen, Dana Scott, and others found deep
new results in mathematical logic and axiomatic set theory using oshoots of Boolean algebra, namely forcing and
Boolean-valued models.
1.2 Denition
A Boolean algebra is a six-tuple consisting of a set A, equipped with two binary operations (called meet or
and), (called join or or), a unary operation (called complement or not) and two elements 0 and 1
(called bottom and top, or least and greatest element, also denoted by the symbols and , respectively),
such that for all elements a, b and c of A, the following axioms hold:[2]
Note, however, that the absorption law can be excluded from the set of axioms as it can be derived from the other
axioms (see Proven properties).
A Boolean algebra with only one element is called a trivial Boolean algebra or a degenerate Boolean algebra.
(Some authors require 0 and 1 to be distinct elements in order to exclude this case.)
It follows from the last three pairs of axioms above (identity, distributivity and complements), or from the absorption
axiom, that
a = b a if and only if a b = b.
The relation dened by a b if these equivalent conditions hold, is a partial order with least element 0 and greatest
element 1. The meet a b and the join a b of two elements coincide with their inmum and supremum, respectively,
with respect to .
The rst four pairs of axioms constitute a denition of a bounded lattice.
It follows from the rst ve pairs of axioms that any complement is unique.
The set of axioms is self-dual in the sense that if one exchanges with and 0 with 1 in an axiom, the result is
again an axiom. Therefore, by applying this operation to a Boolean algebra (or Boolean lattice), one obtains another
Boolean algebra with the same elements; it is called its dual.[3]
1.3 Examples
The simplest non-trivial Boolean algebra, the two-element Boolean algebra, has only two elements, 0 and 1,
and is dened by the rules:
It has applications in logic, interpreting 0 as false, 1 as true, as and, as or, and as not.
Expressions involving variables and the Boolean operations represent statement forms, and two
such expressions can be shown to be equal using the above axioms if and only if the corresponding
statement forms are logically equivalent.
4 CHAPTER 1. BOOLEAN ALGEBRA (STRUCTURE)
The two-element Boolean algebra is also used for circuit design in electrical engineering; here 0
and 1 represent the two dierent states of one bit in a digital circuit, typically high and low voltage.
Circuits are described by expressions containing variables, and two such expressions are equal
for all values of the variables if and only if the corresponding circuits have the same input-output
behavior. Furthermore, every possible input-output behavior can be modeled by a suitable Boolean
expression.
The two-element Boolean algebra is also important in the general theory of Boolean algebras,
because an equation involving several variables is generally true in all Boolean algebras if and only
if it is true in the two-element Boolean algebra (which can be checked by a trivial brute force
algorithm for small numbers of variables). This can for example be used to show that the following
laws (Consensus theorems) are generally valid in all Boolean algebras:
(a b) (a c) (b c) (a b) (a c)
(a b) (a c) (b c) (a b) (a c)
The power set (set of all subsets) of any given nonempty set S forms a Boolean algebra, an algebra of sets, with
the two operations := (union) and := (intersection). The smallest element 0 is the empty set and the
largest element 1 is the set S itself.
After the two-element Boolean algebra, the simplest Boolean algebra is that dened by the power
set of two atoms:
The set of all subsets of S that are either nite or conite is a Boolean algebra, an algebra of sets.
Starting with the propositional calculus with sentence symbols, form the Lindenbaum algebra (that is, the set
of sentences in the propositional calculus modulo tautology). This construction yields a Boolean algebra. It is
in fact the free Boolean algebra on generators. A truth assignment in propositional calculus is then a Boolean
algebra homomorphism from this algebra to the two-element Boolean algebra.
Given any linearly ordered set L with a least element, the interval algebra is the smallest algebra of subsets of
L containing all of the half-open intervals [a, b) such that a is in L and b is either in L or equal to . Interval
algebras are useful in the study of Lindenbaum-Tarski algebras; every countable Boolean algebra is isomorphic
to an interval algebra.
For any natural number n, the set of all positive divisors of n, dening ab if a divides b, forms a distributive
lattice. This lattice is a Boolean algebra if and only if n is square-free. The bottom and the top element of this
Boolean algebra is the natural number 1 and n, respectively. The complement of a is given by n/a. The meet
and the join of a and b is given by the greatest common divisor (gcd) and the least common multiple (lcm) of
a and b, respectively. The ring addition a+b is given by lcm(a,b)/gcd(a,b). The picture shows an example for
n = 30. As a counter-example, considering the non-square-free n=60, the greatest common divisor of 30 and
its complement 2 would be 2, while it should be the bottom element 1.
Other examples of Boolean algebras arise from topological spaces: if X is a topological space, then the col-
lection of all subsets of X which are both open and closed forms a Boolean algebra with the operations :=
(union) and := (intersection).
30
6 10 15
2 3 5
It then follows that f(a) = f(a) for all a in A. The class of all Boolean algebras, together with this notion of
morphism, forms a full subcategory of the category of lattices.
ring is the 1 of the Boolean algebra. This ring has the property that a a = a for all a in A; rings with this property
are called Boolean rings.
Conversely, if a Boolean ring A is given, we can turn it into a Boolean algebra by dening x y := x + y + (x y) and
x y := x y. [4][5] Since these two constructions are inverses of each other, we can say that every Boolean ring arises
from a Boolean algebra, and vice versa. Furthermore, a map f : A B is a homomorphism of Boolean algebras
if and only if it is a homomorphism of Boolean rings. The categories of Boolean rings and Boolean algebras are
equivalent.[6]
Hsiang (1985) gave a rule-based algorithm to check whether two arbitrary expressions denote the same value in every
Boolean ring. More generally, Boudet, Jouannaud, and Schmidt-Schau (1989) gave an algorithm to solve equations
between arbitrary Boolean-ring expressions. Employing the similarity of Boolean rings and Boolean algebras, both
algorithms have applications in automated theorem proving.
An ideal of the Boolean algebra A is a subset I such that for all x, y in I we have x y in I and for all a in A we have
a x in I. This notion of ideal coincides with the notion of ring ideal in the Boolean ring A. An ideal I of A is called
prime if I A and if a b in I always implies a in I or b in I. Furthermore, for every a A we have that a -a =
0 I and then a I or -a I for every a A, if I is prime. An ideal I of A is called maximal if I A and if the
only ideal properly containing I is A itself. For an ideal I, if a I and -a I, then I {a} or I {-a} is properly
contained in another ideal J. Hence, that an I is not maximal and therefore the notions of prime ideal and maximal
ideal are equivalent in Boolean algebras. Moreover, these notions coincide with ring theoretic ones of prime ideal
and maximal ideal in the Boolean ring A.
The dual of an ideal is a lter. A lter of the Boolean algebra A is a subset p such that for all x, y in p we have x y
in p and for all a in A we have a x in p. The dual of a maximal (or prime) ideal in a Boolean algebra is ultralter.
Ultralters can alternatively be described as 2-valued morphisms from A to the two-element Boolean algebra. The
statement every lter in a Boolean algebra can be extended to an ultralter is called the Ultralter Theorem and can
not be proved in ZF, if ZF is consistent. Within ZF, it is strictly weaker than the axiom of choice. The Ultralter
Theorem has many equivalent formulations: every Boolean algebra has an ultralter, every ideal in a Boolean algebra
can be extended to a prime ideal, etc.
1.7 Representations
It can be shown that every nite Boolean algebra is isomorphic to the Boolean algebra of all subsets of a nite set.
Therefore, the number of elements of every nite Boolean algebra is a power of two.
Stones celebrated representation theorem for Boolean algebras states that every Boolean algebra A is isomorphic to
the Boolean algebra of all clopen sets in some (compact totally disconnected Hausdor) topological space.
1.8 Axiomatics
The rst axiomatization of Boolean lattices/algebras in general was given by Alfred North Whitehead in 1898.[7][8]
It included the above axioms and additionally x1=1 and x0=0. In 1904, the American mathematician Edward
V. Huntington (18741952) gave probably the most parsimonious axiomatization based on , , , even proving the
associativity laws (see box).[9] He also proved that these axioms are independent of each other.[10] In 1933, Huntington
set out the following elegant axiomatization for Boolean algebra. It requires just one binary operation + and a unary
functional symbol n, to be read as 'complement', which satisfy the following laws:
1. Commutativity: x + y = y + x.
2. Associativity: (x + y) + z = x + (y + z).
1.9. GENERALIZATIONS 7
Herbert Robbins immediately asked: If the Huntington equation is replaced with its dual, to wit:
do (1), (2), and (4) form a basis for Boolean algebra? Calling (1), (2), and (4) a Robbins algebra, the question then
becomes: Is every Robbins algebra a Boolean algebra? This question (which came to be known as the Robbins
conjecture) remained open for decades, and became a favorite question of Alfred Tarski and his students. In 1996,
William McCune at Argonne National Laboratory, building on earlier work by Larry Wos, Steve Winker, and Bob
Vero, answered Robbinss question in the armative: Every Robbins algebra is a Boolean algebra. Crucial to
McCunes proof was the automated reasoning program EQP he designed. For a simplication of McCunes proof,
see Dahn (1998).
1.9 Generalizations
Removing the requirement of existence of a unit from the axioms of Boolean algebra yields generalized Boolean
algebras. Formally, a distributive lattice B is a generalized Boolean lattice, if it has a smallest element 0 and for any
elements a and b in B such that a b, there exists an element x such that a x = 0 and a x = b. Dening a b
as the unique x such that (a b) x = a and (a b) x = 0, we say that the structure (B,,,,0) is a generalized
Boolean algebra, while (B,,0) is a generalized Boolean semilattice. Generalized Boolean lattices are exactly the ideals
of Boolean lattices.
A structure that satises all axioms for Boolean algebras except the two distributivity axioms is called an orthocomplemented
lattice. Orthocomplemented lattices arise naturally in quantum logic as lattices of closed subspaces for separable
Hilbert spaces.
1.11 Notes
[1] Givant and Paul Halmos, 2009, p. 20
[3] Goodstein, R. L. (2012), Chapter 2: The self-dual system of axioms, Boolean Algebra, Courier Dover Publications, pp.
21, ISBN 9780486154978.
[7] Padmanabhan, p. 73
1.12 References
Brown, Stephen; Vranesic, Zvonko (2002), Fundamentals of Digital Logic with VHDL Design (2nd ed.),
McGrawHill, ISBN 978-0-07-249938-4. See Section 2.5.
8 CHAPTER 1. BOOLEAN ALGEBRA (STRUCTURE)
A. Boudet; J.P. Jouannaud; M. Schmidt-Schau (1989). Unication in Boolean Rings and Abelian Groups
(PDF). Journal of Symbolic Computation. 8: 449477. doi:10.1016/s0747-7171(89)80054-9.
Cohn, Paul M. (2003), Basic Algebra: Groups, Rings, and Fields, Springer, pp. 51, 7081, ISBN 9781852335878
Cori, Rene; Lascar, Daniel (2000), Mathematical Logic: A Course with Exercises, Oxford University Press,
ISBN 978-0-19-850048-3. See Chapter 2.
Dahn, B. I. (1998), Robbins Algebras are Boolean: A Revision of McCunes Computer-Generated Solution
of the Robbins Problem, Journal of Algebra, 208 (2): 526532, doi:10.1006/jabr.1998.7467.
B.A. Davey; H.A. Priestley (1990). Introduction to Lattices and Order. Cambridge Mathematical Textbooks.
Cambridge University Press.
Givant, Steven; Halmos, Paul (2009), Introduction to Boolean Algebras, Undergraduate Texts in Mathematics,
Springer, ISBN 978-0-387-40293-2.
Halmos, Paul (1963), Lectures on Boolean Algebras, Van Nostrand, ISBN 978-0-387-90094-0.
Halmos, Paul; Givant, Steven (1998), Logic as Algebra, Dolciani Mathematical Expositions, 21, Mathematical
Association of America, ISBN 978-0-88385-327-6.
Hsiang, Jieh (1985). Refutational Theorem Proving Using Term Rewriting Systems (PDF). AI. 25: 255300.
doi:10.1016/0004-3702(85)90074-8.
Edward V. Huntington (1904). Sets of Independent Postulates for the Algebra of Logic. Transactions of the
American Mathematical Society. 5: 288309. JSTOR 1986459. doi:10.1090/s0002-9947-1904-1500675-4.
Huntington, E. V. (1933), New sets of independent postulates for the algebra of logic (PDF), Transactions
of the American Mathematical Society, American Mathematical Society, 35 (1): 274304, JSTOR 1989325,
doi:10.2307/1989325.
Huntington, E. V. (1933), Boolean algebra: A correction, Transactions of the American Mathematical Society,
35 (2): 557558, JSTOR 1989783, doi:10.2307/1989783.
Mendelson, Elliott (1970), Boolean Algebra and Switching Circuits, Schaums Outline Series in Mathematics,
McGrawHill, ISBN 978-0-07-041460-0.
Monk, J. Donald; Bonnet, R., eds. (1989), Handbook of Boolean Algebras, North-Holland, ISBN 978-0-
444-87291-3. In 3 volumes. (Vol.1:ISBN 978-0-444-70261-6, Vol.2:ISBN 978-0-444-87152-7, Vol.3:ISBN
978-0-444-87153-4)
Padmanabhan, Ranganathan; Rudeanu, Sergiu (2008), Axioms for lattices and boolean algebras, World Scien-
tic, ISBN 978-981-283-454-6.
Sikorski, Roman (1966), Boolean Algebras, Ergebnisse der Mathematik und ihrer Grenzgebiete, Springer Ver-
lag.
Stoll, R. R. (1963), Set Theory and Logic, W. H. Freeman, ISBN 978-0-486-63829-4. Reprinted by Dover
Publications, 1979.
Marshall H. Stone (1936). The Theory of Representations for Boolean Algebra. Transactions of the Ameri-
can Mathematical Society. 40: 37111. doi:10.1090/s0002-9947-1936-1501865-8.
A.N. Whitehead (1898). A Treatise on Universal Algebra. Cambridge University Press. ISBN 1-4297-0032-7.
Burris, Stanley N.; Sankappanavar, H. P., 1981. A Course in Universal Algebra. Springer-Verlag. ISBN
3-540-90578-2.
Weisstein, Eric W. Boolean Algebra. MathWorld.
Chapter 2
In the mathematical area of order theory, every partially ordered set P gives rise to a dual (or opposite) partially
ordered set which is often denoted by P op or P d . This dual order P op is dened to be the set with the inverse order,
i.e. x y holds in P op if and only if y x holds in P. It is easy to see that this construction, which can be depicted by
ipping the Hasse diagram for P upside down, will indeed yield a partially ordered set. In a broader sense, two posets
are also said to be duals if they are dually isomorphic, i.e. if one poset is order isomorphic to the dual of the other.
The importance of this simple denition stems from the fact that every denition and theorem of order theory can
readily be transferred to the dual order. Formally, this is captured by the Duality Principle for ordered sets:
If a given statement is valid for all partially ordered sets, then its dual statement, obtained by inverting
the direction of all order relations and by dualizing all order theoretic denitions involved, is also valid
for all partially ordered sets.
If a statement or denition is equivalent to its dual then it is said to be self-dual. Note that the consideration of dual
orders is so fundamental that it often occurs implicitly when writing for the dual order of without giving any prior
denition of this new symbol.
2.1 Examples
Naturally, there are a great number of examples for concepts that are dual:
Since partial orders are antisymmetric, the only ones that are self-dual are the equivalence relations.
10
2.2. SEE ALSO 11
a=1 e'=1
b c' d'
c d b'
e=0 a'=0
A bounded distributive lattice, and its dual
Transpose graph
2.3 References
[1] The quantiers are essential: for individual elements x, y, z, e.g. the rst equation may be violated, but the second may
hold; see the N5 lattice for an example.
Davey, B.A.; Priestley, H. A. (2002), Introduction to Lattices and Order (2nd ed.), Cambridge University Press,
ISBN 978-0-521-78451-1
Chapter 3
Filter (mathematics)
1, 2, 3, 4
1, 2, 3 1, 2, 4 1, 3, 4 2, 3, 4
1, 2 1, 3 1, 4 2, 3 2, 4 3, 4
1 2 3 4
The powerset lattice of the set {1,2,3,4}, with the upper set {1,4} colored yellow. It is a principal lter, but not an ultralter, as
it can be extended to the larger nontrivial lter {1}, by including also the light green elements. Since {1} cannot be extended any
further, it is an ultralter.
In mathematics, a lter is a special subset of a partially ordered set. For example, the power set of some set, partially
ordered by set inclusion, is a lter. Filters appear in order and lattice theory, but can also be found in topology whence
they originate. The dual notion of a lter is an ideal.
Filters were introduced by Henri Cartan in 1937[1][2] and subsequently used by Bourbaki in their book Topologie
Gnrale as an alternative to the similar notion of a net developed in 1922 by E. H. Moore and H. L. Smith.
12
3.1. MOTIVATION 13
3.1 Motivation
Intuitively, a lter in a partially ordered set (poset), X, is a subset of X that includes as members those elements that
are large enough to satisfy some criterion. For example, if x is an element of the poset, then the set of elements that
are above x is a lter, called the principal lter at x. (Notice that if x and y are incomparable elements of the poset,
then neither of the principal lters at x and y is contained in the other one, and conversely.)
Similarly, a lter on a set contains those subsets that are suciently large to contain something. For example, if the
set is the real line and x is one of its points, then the family of sets that include x in their interior is a lter, called the
lter of neighbourhoods of x. (Notice that the thing in this case is slightly larger than x, but it still doesn't contain
any other specic point of the line.)
The above interpretations do not really, without elaboration, explain the condition 2. of the general denition of lter
(see below). For, why should two large enough things contain a common large enough thing? (Note, however,
that they do explain conditions 1 and 3: Clearly the empty set is not large enough, and clearly the collection of
large enough things should be upward closed.)
Alternatively, a lter can be viewed as a locating scheme": Suppose we try to locate something (a point or a subset)
in the space X. Call a lter the collection of subsets of X that might contain what we are looking for. Then this
lter should possess the following natural structure: 1. Empty set cannot contain anything so it will not belong to
our lter. 2. If two subsets, E and F, both might contain what we are looking for, then so might their intersection.
Thus our lter should be closed with respect to nite intersection. 3. If a set E might contain what we are looking
for, so might any superset of it. Thus our lter is upward closed.
An ultralter can be viewed as a perfect locating scheme where each subset E of the space X can be used in
deciding whether what we are looking for might lie in E.
From this interpretation, compactness (see the mathematical characterization below) can be viewed as the property
that no location scheme can end up with nothing, or, to put it another way, we will always nd something.
The mathematical notion of lter provides a precise language to treat these situations in a rigorous and general way,
which is useful in analysis, general topology and logic.
A subset F of a partially ordered set (P,) is a lter if the following conditions hold:
1. F is nonempty.
2. For every x, y in F, there is some element z in F such that z x and z y. (F is a lter base (see below), or
downward directed)
A lter is proper if it is not equal to the whole set P. This condition is sometimes added to the denition of a lter.
While the above denition is the most general way to dene a lter for arbitrary posets, it was originally dened for
lattices only. In this case, the above denition can be characterized by the following equivalent statement: A subset
F of a lattice (P,) is a lter, if and only if it is an upper set that is closed under nite intersection (inma or meet),
i.e., for all x, y in F, we nd that x y is also in F.
The smallest lter that contains a given element p is a principal lter and p is a principal element in this situation.
The principal lter for p is just given by the set {x P | p x} and is denoted by prexing p with an upward arrow:
p.
The dual notion of a lter, i.e. the concept obtained by reversing all and exchanging with , is ideal. Because of
this duality, the discussion of lters usually boils down to the discussion of ideals. Hence, most additional information
on this topic (including the denition of maximal lters and prime lters) is to be found in the article on ideals.
There is a separate article on ultralters.
14 CHAPTER 3. FILTER (MATHEMATICS)
1. S is in F, and if A and B are in F, then so is their intersection. (F is closed under nite intersection)
2. If A is in F and A is a subset of B, then B is in F, for all subsets B of S. (F is upward closed)
[3]
If the empty set is not in F, we say F is a proper lter.
The rst two properties imply that a lter on a set has the nite intersection property. Note that with this denition,
a lter on a set is indeed a lter. The only nonproper lter on S is P(S).
A lter base (or lter basis) is a subset B of P(S) with the following properties:
1. B is non-empty and the intersection of any two members of B contains a member of B (B is downward directed).
2. The empty set is not a member of B (B is a proper lter base).
Given a lter base B, the lter generated or spanned by B is dened as the minimum lter containing B. It is the family
of all the subsets of S which contain a member of B. Every lter is also a lter base, so the process of passing from
lter base to lter may be viewed as a sort of completion.
If B and C are two lter bases on S, one says C is ner than B (or that C is a renement of B) if for each B0 B,
there is a C 0 C such that C 0 B0 . If also B is ner than C, one says that they are equivalent lter bases.
If B and C are lter bases, then C is ner than B if and only if the lter spanned by C contains the lter spanned
by B. Therefore, B and C are equivalent lter bases if and only if they generate the same lter.
For lter bases A, B, and C, if A is ner than B and B is ner than C then A is ner than C. Thus the renement
relation is a preorder on the set of lter bases, and the passage from lter base to lter is an instance of passing
from a preordering to the associated partial ordering.
For any subset T of P(S) there is a smallest (possibly nonproper) lter F containing T, called the lter generated or
spanned by T. It is constructed by taking all nite intersections of T, which then form a lter base for F. This lter is
proper if and only if any nite intersection of elements of T is non-empty, and in that case we say that T is a lter
subbase.
3.3.1 Examples
Let S be a nonempty set and C be a nonempty subset of S. Then {C} is a lter base. The lter it generates
(i.e., the collection of all subsets containing C) is called the principal lter generated by C.
A lter is said to be a free lter if the intersection of all of its members is empty. A principal lter is not
free. Since the intersection of any nite number of members of a lter is also a member, no lter on a nite
set is free, and indeed is the principal lter generated by the common intersection of all of its members. A
nonprincipal lter on an innite set is not necessarily free.
The Frchet lter on an innite set S is the set of all subsets of S that have nite complement. A lter on S is
free if and only if it contains the Frchet lter.
Every uniform structure on a set X is a lter on XX.
A lter in a poset can be created using the Rasiowa-Sikorski lemma, often used in forcing.
The set {{N, N + 1, N + 2, . . . } : N {1, 2, 3, . . . }} is called a lter base of tails of the sequence of
natural numbers (1, 2, 3, . . . ) . A lter base of tails can be made of any net (x )A using the construction
{{x : A, 0 } : 0 A} where the lter that this lter base generates is called the nets eventuality
lter. Therefore, all nets generate a lter base (and therefore a lter). Since all sequences are nets, this holds
for sequences as well.
3.3. FILTER ON A SET 15
1 ifA F
m(A) = 0 ifS \ A F
undened otherwise
is nitely additive a "measure" if that term is construed rather loosely. Therefore the statement
{ x S : (x) } F
can be considered somewhat analogous to the statement that holds almost everywhere. That interpretation of
membership in a lter is used (for motivation, although it is not needed for actual proofs) in the theory of ultraproducts
in model theory, a branch of mathematical logic.
Neighbourhood bases
Take Nx to be the neighbourhood lter at point x for X. This means that Nx is the set of all topological
neighbourhoods of the point x. It can be veried that Nx is a lter. A neighbourhood system is another name
for a neighbourhood lter.
To say that N is a neighbourhood base at x for X means that each subset V 0 of X is a neighbourhood of x if
and only if there exists N 0 N such that N 0 V 0 . Note that every neighbourhood base at x is a lter base that
generates the neighbourhood lter at x.
To say that a lter base B converges to x, denoted B x, means that for every neighbourhood U of x, there is
a B0 B such that B0 U. In this case, x is called a limit of B and B is called a convergent lter base.
Every neighbourhood base N of x converges to x.
If N is a neighbourhood base at x and C is a lter base on X, then C x if and only if C is ner than N.
If Y X, a point p X is called a limit point of Y in X if and only if each neighborhood U of p in X
intersects Y. This happens if and only if there is a lter base of subsets of Y that converges to p in X.
For Y X, the following are equivalent:
16 CHAPTER 3. FILTER (MATHEMATICS)
(i) There exists a lter base F whose elements are all contained in Y such that F x.
(ii) There exists a lter F such that Y is an element of F and F x.
(iii) The point x lies in the closure of Y.
Indeed:
(i) implies (ii): if F is a lter base satisfying the properties of (i), then the lter associated to F satises the properties
of (ii).
(ii) implies (iii): if U is any open neighborhood of x then by the denition of convergence U contains an element of
F; since also Y is an element of F, U and Y have nonempty intersection.
(iii) implies (i): Dene F = {U Y | U Nx } . Then F is a lter base satisfying the properties of (i).
Clustering
A lter base B on X is said to cluster at x (or have x as a cluster point) if and only if each element of B has
nonempty intersection with each neighbourhood of x.
If a lter base B clusters at x and is ner than a lter base C, then C clusters at x too.
Every limit of a lter base is also a cluster point of the base.
A lter base B that has x as a cluster point may not converge to x. But there is a ner lter base that does.
For example the lter base of nite intersections of sets of the subbase B Nx .
For a lter base B, the set {cl(B0 ) : B0 B} is the set of all cluster points of B (note: cl(B0 ) is the closure
of B0 ). Assume that X is a complete lattice.
The limit inferior of B is the inmum of the set of all cluster points of B.
The limit superior of B is the supremum of the set of all cluster points of B.
B is a convergent lter base if and only if its limit inferior and limit superior agree; in this case, the
value on which they agree is the limit of the lter base.
X is a Hausdor space if and only if every lter base on X has at most one limit.
X is compact if and only if every lter base on X clusters or has a cluster point.
X is compact if and only if every lter base on X is a subset of a convergent lter base.
Let X , Y be topological spaces. Let A be a lter base on X and f : X Y be a function. The image of A under
f is dened as the set {f [a] : a A} . The image is denoted f [A] and forms a lter base on Y .
Cauchy lters
To say that a lter base B on X is Cauchy means that for each real number >0, there is a B0 B such that the
metric diameter of B0 is less than .
Take (xn) to be a sequence in metric space X. (xn) is a Cauchy sequence if and only if the lter base {{xN,xN+1,...}
: N {1,2,3,...} } is Cauchy.
More generally, given a uniform space X, a lter F on X is called Cauchy lter if for every entourage U there is an A
F with (x,y) U for all x,y A. In a metric space this agrees with the previous denition. X is said to be complete
if every Cauchy lter converges. Conversely, on a uniform space every convergent lter is a Cauchy lter. Moreover,
every cluster point of a Cauchy lter is a limit point.
A compact uniform space is complete: on a compact space each lter has a cluster point, and if the lter is Cauchy,
such a cluster point is a limit point. Further, a uniformity is compact if and only if it is complete and totally bounded.
Most generally, a Cauchy space is a set equipped with a class of lters declared to be Cauchy. These are required to
have the following properties:
3. if F and G are Cauchy lters and each member of F intersects each member of G, then F G is Cauchy.
The Cauchy lters on a uniform space have these properties, so every uniform space (hence every metric space)
denes a Cauchy space.
Filtration (mathematics)
Net (mathematics)
Generic lter
3.5 Notes
[1] H. Cartan, Thorie des ltres, CR Acad. Paris, 205, (1937) 595598.
3.6 References
Nicolas Bourbaki, General Topology (Topologie Gnrale), ISBN 0-387-19374-X (Ch. 1-4): Provides a good
reference for lters in general topology (Chapter I) and for Cauchy lters in uniform spaces (Chapter II)
Stephen Willard, General Topology, (1970) Addison-Wesley Publishing Company, Reading Massachusetts.
(Provides an introductory review of lters in topology.)
18 CHAPTER 3. FILTER (MATHEMATICS)
David MacIver, Filters in Analysis and Topology (2004) (Provides an introductory review of lters in topology
and in metric spaces.)
Burris, Stanley N., and H.P. Sankappanavar, H. P., 1981. A Course in Universal Algebra. Springer-Verlag.
ISBN 3-540-90578-2.
Chapter 4
Greatest element
In mathematics, especially in order theory, the greatest element of a subset S of a partially ordered set (poset) is an
element of S that is greater than every other element of S. The term least element is dened dually, that is, it is an
element of S that is smaller than every other element of S.
Formally, given a partially ordered set (P, ), an element g of a subset S of P is the greatest element of S if
Hence, the greatest element of S is an upper bound of S that is contained within this subset. It is necessarily unique.
By using instead of in the above denition, one denes the least element of S.
Like upper bounds, greatest elements may fail to exist. Even if a set has some upper bounds, it need not have a
greatest element, as shown by the example of the negative real numbers. This example also demonstrates that the
existence of a least upper bound (the number 0 in this case) does not imply the existence of a greatest element either.
Similar conclusions hold for least elements. A nite chain always has a greatest and a least element.
A greatest element of a partially ordered subset must not be confused with maximal elements of the set, which are
elements that are not smaller than any other elements. A set can have several maximal elements without having a
greatest element. However, if it has a greatest element, it can't have any other maximal element.
In a totally ordered set both terms coincide; it is also called maximum; in the case of function values it is also called
the absolute maximum, to avoid confusion with a local maximum.[1] The dual terms are minimum and absolute
minimum. Together they are called the absolute extrema.
The least and greatest element of the whole partially ordered set plays a special role and is also called bottom and top,
or zero (0) and unit (1), or and , respectively. If both exists, the poset is called a bounded poset. The notation
of 0 and 1 is used preferably when the poset is even a complemented lattice, and when no confusion is likely, i.e.
when one is not talking about partial orders of numbers that already contain elements 0 and 1 dierent from bottom
and top. The existence of least and greatest elements is a special completeness property of a partial order.
Further introductory information is found in the article on order theory.
4.1 Examples
The subset of integers has no upper bound in the set of real numbers.
Let the relation "" on {a, b, c, d} be given by a c, a d, b c, b d. The set {a, b} has upper bounds c
and d, but no least upper bound, and no greatest element.
In the rational numbers, the set of numbers with their square less than 2 has upper bounds but no greatest
element and no least upper bound.
In , the set of numbers less than 1 has a least upper bound, viz. 1, but no greatest element.
In , the set of numbers less than or equal to 1 has a greatest element, viz. 1, which is also its least upper
bound.
19
20 CHAPTER 4. GREATEST ELEMENT
In with the product order, the set of pairs (x, y) with 0 < x < 1 has no upper bound.
In with the lexicographical order, this set has upper bounds, e.g. (1, 0). It has no least upper bound.
Maximal element
4.3 References
[1] The notion of locality requires the functions domain to be at least a topological space.
Davey, B. A.; Priestley, H. A. (2002). Introduction to Lattices and Order (2nd ed.). Cambridge University
Press. ISBN 978-0-521-78451-1.
Chapter 5
In mathematical order theory, an ideal is a special subset of a partially ordered set (poset). Although this term
historically was derived from the notion of a ring ideal of abstract algebra, it has subsequently been generalized to a
dierent notion. Ideals are of great importance for many constructions in order and lattice theory.
1. I is non-empty,
3. for every x, y in I, there is some element z in I, such that x z and y z. (I is a directed set).
While this is the most general way to dene an ideal for arbitrary posets, it was originally dened for lattices only. In
this case, the following equivalent denition can be given: a subset I of a lattice (P,) is an ideal if and only if it is a
lower set that is closed under nite joins (suprema), i.e., it is nonempty and for all x, y in I, the element x y of P is
also in I.[3]
The dual notion of an ideal, i.e., the concept obtained by reversing all and exchanging with , is a lter.
Some authors use the term ideal to mean a lower set, i.e., they include only condition 2 above.[4][5][6] With this weaker
denition, an ideal of a lattice seen as a poset is not closed under joins, so it is not necessarily an ideal of the lattice.[6]
Wikipedia uses only ideal/lter (of order theory)" and lower/upper set to avoid confusion.
Frink ideals, pseudoideals and Doyle pseudoideals are dierent generalizations of the notion of a lattice ideal.
An ideal or lter is said to be proper if it is not equal to the whole set P.[3]
The smallest ideal that contains a given element p is a principal ideal and p is said to be a principal element of the
ideal in this situation. The principal ideal p for a principal p is thus given by p = {x in P | x p}.
21
22 CHAPTER 5. IDEAL (ORDER THEORY)
It is easily checked that this is indeed equivalent to stating that P\I is a lter (which is then also prime, in the dual
sense).
For a complete lattice the further notion of a completely prime ideal is meaningful. It is dened to be a proper ideal
I with the additional property that, whenever the meet (inmum) of some arbitrary set A is in I, some element of A
is also in I. So this is just a specic prime ideal that extends the above conditions to innite meets.
The existence of prime ideals is in general not obvious, and often a satisfactory amount of prime ideals cannot be
derived within ZF (ZermeloFraenkel set theory without the axiom of choice). This issue is discussed in various
prime ideal theorems, which are necessary for many applications that require prime ideals.
Proof. Assume the ideal M is maximal with respect to disjointness from the lter F. Suppose
for a contradiction that M is not prime, i.e. there exists a pair of elements a and b such that
a b in M but neither a nor b are in M. Consider the case that for all m in M, m a is not in
F. One can construct an ideal N by taking the downward closure of the set of all binary joins
of this form, i.e. N = { x | x m a for some m in M}. It is readily checked that N is indeed
an ideal disjoint from F which is strictly greater than M. But this contradicts the maximality
of M and thus the assumption that M is not prime.
For the other case, assume that there is some m in M with m a in F. Now if any element
n in M is such that n b is in F, one nds that (m n) b and (m n) a are both in F.
But then their meet is in F and, by distributivity, (m n) (a b) is in F too. On the other
hand, this nite join of elements of M is clearly in M, such that the assumed existence of n
contradicts the disjointness of the two sets. Hence all elements n of M have a join with b that
is not in F. Consequently one can apply the above construction with b in place of a to obtain
an ideal that is strictly greater than M while being disjoint from F. This nishes the proof.
However, in general it is not clear whether there exists any ideal M that is maximal in this sense. Yet, if we assume the
Axiom of Choice in our set theory, then the existence of M for every disjoint lterideal-pair can be shown. In the
special case that the considered order is a Boolean algebra, this theorem is called the Boolean prime ideal theorem.
It is strictly weaker than the Axiom of Choice and it turns out that nothing more is needed for many order theoretic
applications of ideals.
5.4 Applications
The construction of ideals and lters is an important tool in many applications of order theory.
In Stones representation theorem for Boolean algebras, the maximal ideals (or, equivalently via the negation
map, ultralters) are used to obtain the set of points of a topological space, whose clopen sets are isomorphic
to the original Boolean algebra.
5.5. HISTORY 23
Order theory knows many completion procedures, to turn posets into posets with additional completeness
properties. For example, the ideal completion of a given partial order P is the set of all ideals of P ordered by
subset inclusion. This construction yields the free dcpo generated by P. An ideal is principal if and only if it is
compact in the ideal completion, so the original poset can be recovered as the sub-poset consisting of compact
elements. Furthermore, every algebraic dcpo can be reconstructed as the ideal completion of its set of compact
elements.
5.5 History
Ideals were introduced rst by Marshall H. Stone, who derived their name from the ring ideals of abstract algebra.
He adopted this terminology because, using the isomorphism of the categories of Boolean algebras and of Boolean
rings, the two notions do indeed coincide.
5.6 Literature
Ideals and lters are among the most basic concepts of order theory. See the introductory books given for order
theory and lattice theory, and the literature on the Boolean prime ideal theorem.
5.8 Notes
[1] Taylor (1999), p. 141: A directed lower subset of a poset X is called an ideal
[2] Gierz, G.; Hofmann, K. H.; Keimel, K.; Lawson, J. D.; Mislove, M. W.; Scott, D. S. (2003). Continuous Lattices and
Domains. Encyclopedia of Mathematics and its Applications. 93. Cambridge University Press. p. 3. ISBN 0521803381.
[6] Davey & Priestley, Introduction to Lattices and Order (Second Edition), 2002, p. 20 and 44
5.9 References
Burris, Stanley N.; Sankappanavar, Hanamantagouda P. (1981). A Course in Universal Algebra. Springer-
Verlag. ISBN 3-540-90578-2.
Lawson, M.V. (1998). Inverse semigroups: the theory of partial symmetries. World Scientic. ISBN 978-981-
02-3316-7.
Stanley, R.P. (2002). Enumerative combinatorics. Cambridge studies in advanced mathematics. 1. Cambridge
University Press. ISBN 978-0-521-66351-9.
Taylor, Paul (1999), Practical foundations of mathematics, Cambridge Studies in Advanced Mathematics, 59,
Cambridge University Press, Cambridge, ISBN 0-521-63107-6, MR 1694820
Chapter 6
Ideal Product redirects here. For products named Ideal, see Ideal (disambiguation).
In ring theory, a branch of abstract algebra, an ideal is a special subset of a ring. Ideals generalize certain subsets
of the integers, such as the even numbers or the multiples of 3. Addition and subtraction of even numbers preserves
evenness, and multiplying an even number by any other integer results in another even number; these closure and
absorption properties are the dening properties of an ideal. An ideal can be used to construct a quotient ring similarly
to the way that, in group theory, a normal subgroup can be used to construct a quotient group.
Among the integers, the ideals correspond one-for-one with the non-negative integers: in this ring, every ideal is a
principal ideal consisting of the multiples of a single non-negative number. However, in other rings, the ideals may
be distinct from the ring elements, and certain properties of integers, when generalized to rings, attach more naturally
to the ideals than to the elements of the ring. For instance, the prime ideals of a ring are analogous to prime numbers,
and the Chinese remainder theorem can be generalized to ideals. There is a version of unique prime factorization for
the ideals of a Dedekind domain (a type of ring important in number theory).
The concept of an order ideal in order theory is derived from the notion of ideal in ring theory. A fractional ideal is
a generalization of an ideal, and the usual ideals are sometimes called integral ideals for clarity.
6.1 History
Ideals were rst proposed by Richard Dedekind in 1876 in the third edition of his book Vorlesungen ber Zahlentheorie
(English: Lectures on Number Theory). They were a generalization of the concept of ideal numbers developed by
Ernst Kummer.[1][2] Later the concept was expanded by David Hilbert and especially Emmy Noether.
6.2 Denitions
For an arbitrary ring (R, +, ) , let (R, +) be its additive group. A subset I is called a two-sided ideal (or simply an
ideal) of R if it is an additive subgroup of R that absorbs multiplication by elements of R. Formally we mean that
I is an ideal if it satises the following conditions:
2. x I, r R : x r, r x I
24
6.3. PROPERTIES 25
2. x I, r R : x r I.
The left ideals in R are exactly the right ideals in the opposite ring Ro and vice versa. A two-sided ideal is a left ideal
that is also a right ideal, and is often called an ideal except to emphasize that there might exist single-sided ideals.
When R is a commutative ring, the denitions of left, right, and two-sided ideal coincide, and the term ideal is used
alone.
6.3 Properties
{0} and R are ideals in every ring R. If R is a division ring or a eld, then these are its only ideals. The ideal R is
called the unit ideal. Any ideal I is a proper ideal if it is a proper subset of R, that is, if I does not equal R.[5]
Just as normal subgroups of groups are kernels of group homomorphisms, ideals have interpretations as kernels. For
a nonempty subset A of R:
If p is in I, then pR is a right ideal and Rp is a left ideal of R. These are called, respectively, the principal right and left
ideals generated by p. To remember which is which, note that right ideals are stable under right-multiplication (IR
I) and left ideals are stable under left-multiplication (RI I).
The connection between cosets and ideals can be seen by switching the operation from multiplication to addition.
6.4 Motivation
Intuitively, the denition can be motivated as follows: Suppose we have a subset of elements Z of a ring R and that
we would like to obtain a ring with the same structure as R, except that the elements of Z should be zero (they are in
some sense negligible).
But if z1 = 0 and z2 = 0 in our new ring, then surely z1 + z2 should be zero too, and rz1 as well as z1 r should be
zero for any element r (zero or not).
The denition of an ideal is such that the ideal I generated (see below) by Z is exactly the set of elements that are
forced to become zero if Z becomes zero, and the quotient ring R/I is the desired ring where Z is zero, and only
elements that are forced by Z to be zero are zero. The requirement that R and R/I should have the same structure
(except that I becomes zero) is formalized by the condition that the projection from R to R/I is a (surjective) ring
homomorphism.
26 CHAPTER 6. IDEAL (RING THEORY)
6.5 Examples
In a ring R, the set R itself forms an ideal of R. Also, the subset containing only the additive identity 0R forms
an ideal. These two ideals are usually referred to as the trivial ideals of R.
The even integers form an ideal in the ring Z of all integers; it is usually denoted by 2Z . This is because the
sum of any even integers is even, and the product of any integer with an even integer is also even. Similarly,
the set of all integers divisible by a xed integer n is an ideal denoted nZ .
The set of all polynomials with real coecients which are divisible by the polynomial x2 + 1 is an ideal in the
ring of all polynomials.
The set of all n-by-n matrices whose last row is zero forms a right ideal in the ring of all n-by-n matrices. It is
not a left ideal. The set of all n-by-n matrices whose last column is zero forms a left ideal but not a right ideal.
The ring C(R) of all continuous functions f from R to R under pointwise multiplication contains the ideal of
all continuous functions f such that f(1) = 0. Another ideal in C(R) is given by those functions which vanish
for large enough arguments, i.e. those continuous functions f for which there exists a number L > 0 such that
f(x) = 0 whenever |x| > L.
Compact operators form an ideal in the ring of bounded operators.
{r1 x1 + + rn xn | n N, ri R, xi X}.
Each element described would have to be in every left ideal containing X, so this left ideal is in fact the left ideal
generated by X. The right ideal and ideal generated by X can also be expressed in the same way:
{x1 r1 + + xn rn | n N, ri R, xi X}
{r1 x1 s1 + + rn xn sn | n N, ri R, si R, xi X}.
The former is the right ideal generated by X, and the latter is the ideal generated by X.
By convention, 0 is viewed as the sum of zero such terms, agreeing with the fact that the ideal of R generated by is
{0} by the previous denition.
If a left ideal I of R has a nite subset F such that I is the left ideal generated by F, then the left ideal I is said to be
nitely generated. Similar terms are also applied to right ideals and two-sided ideals generated by nite subsets.
In the special case where the set X is just a singleton {a} for some a in R, then the above denitions turn into the
following:
Ra = {ra | r R}
aR = {ar | r R}
RaR = {r1 as1 + + rn asn | n N, ri R, si R}.
These ideals are known as the left/right/two-sided principal ideals generated by a. It is also very common to denote
the two-sided ideal generated by a as (a).
6.7. TYPES OF IDEALS 27
If R does not have a unit, then the internal descriptions above must be modied slightly. In addition to the nite sums
of products of things in X with things in R, we must allow the addition of n-fold sums of the form x+x+...+x, and
n-fold sums of the form (x)+(x)+...+(x) for every x in X and every n in the natural numbers. When R has a unit,
this extra requirement becomes superuous.
6.6.1 Example
In the ring Z of integers, every ideal can be generated by a single number (so Z is a principal ideal domain),
and the only two generators of pR are p and p. The concepts of ideal and number are therefore almost
identical in Z . If aR = bR in an arbitrary domain, then au = b for some unit u. Conversely, for any unit u,
aR = auu1 R = auR. So, in a commutative principal ideal domain, the generators of the ideal aR are just the
elements au where u is an arbitrary unit. This explains the case of Z since 1 and 1 are the only units of Z .
are certain subsets of R associated to I. The most well known examples are the rings Z/nZ, created from the ring Z
of integers and its ideals.
To simplify the description all rings are assumed to be commutative. The non-commutative case is discussed
in detail in the respective articles.
Ideals are important because they appear as kernels of ring homomorphisms and allow one to dene factor rings.
Dierent types of ideals are studied because they can be used to construct dierent types of factor rings.
Maximal ideal: A proper ideal I is called a maximal ideal if there exists no other proper ideal J with I a
proper subset of J. The factor ring of a maximal ideal is a simple ring in general and is a eld for commutative
rings.[6]
Minimal ideal: A nonzero ideal is called minimal if it contains no other nonzero ideal.
Prime ideal: A proper ideal I is called a prime ideal if for any a and b in R, if ab is in I, then at least one
of a and b is in I. The factor ring of a prime ideal is a prime ring in general and is an integral domain for
commutative rings.
Radical ideal or semiprime ideal: A proper ideal I is called radical or semiprime if for any a in R, if an is
in I for some n, then a is in I. The factor ring of a radical ideal is a semiprime ring for general rings, and is a
reduced ring for commutative rings.
Primary ideal: An ideal I is called a primary ideal if for all a and b in R, if ab is in I, then at least one of
a and bn is in I for some natural number n. Every prime ideal is primary, but not conversely. A semiprime
primary ideal is prime.
Primitive ideal: A left primitive ideal is the annihilator of a simple left module. A right primitive ideal is
dened similarly. Actually (despite the name) the left and right primitive ideals are always two-sided ideals.
Primitive ideals are prime. A factor rings constructed with a right (left) primitive ideals is a right (left) primitive
ring. For commutative rings the primitive ideals are maximal, and so commutative primitive rings are all elds.
Irreducible ideal: An ideal is said to be irreducible if it cannot be written as an intersection of ideals which
properly contain it.
Comaximal ideals: Two ideals i, j are said to be comaximal if x + y = 1 for some x i and y j .
28 CHAPTER 6. IDEAL (RING THEORY)
Regular ideal: This term has multiple uses. See the article for a list.
Nil ideal: An ideal is a nil ideal if each of its elements is nilpotent.
Two other important terms using ideal are not always ideals of their ring. See their respective articles for details:
Fractional ideal: This is usually dened when R is a commutative domain with quotient eld K. Despite their
names, fractional ideals are R submodules of K with a special property. If the fractional ideal is contained
entirely in R, then it is truly an ideal of R.
Invertible ideal: Usually an invertible ideal A is dened as a fractional ideal for which there is another fractional
ideal B such that AB=BA=R. Some authors may also apply invertible ideal to ordinary ring ideals A and B
with AB=BA=R in rings other than domains.
a + b := {a + b | a a and b b}
and
i.e. the product of two ideals a and b is dened to be the ideal ab generated by all products of the form ab with a in
a and b in b . The product ab is contained in the intersection of a and b .
Note that a + b is also the intersection of all ideals containing both a and b .
The sum and the intersection of ideals is again an ideal; with these two operations as join and meet, the set of all
ideals of a given ring forms a complete modular lattice. Also, the union of two ideals is a subset of the sum of those
two ideals, because for any element a inside an ideal, we can write it as a+0, or 0+a, therefore, it is contained in the
sum as well. However, the union of two ideals is not necessarily an ideal.
6.10. EXAMPLES OF IDEAL OPERATIONS 29
since (n) (m) is the set of integers which are divisible by both n and m .
Let R = C[x, y, z, w] and let I = (z, w), J = (x + z, y + w), K = (x + z, w) . Then,
IJ = (z(x + z), z(y + w), w(x + z), w(y + w)) = (z 2 + xz, zy + wz, wx + wz, wy + w2 )
I J = IJ while I K = (w, xz + z 2 ) = IK
In the rst computation, we see the general pattern for taking the sum of two nitely generated ideals, it is the ideal
generated by the union of their generators. In the last three we observe that products and intersections agree whenever
the two ideals intersect in the zero ideal. These computations can be checked using macaulay2.[7][8][9]
Modular arithmetic
Ideal theory
Ideal quotient
Ideal norm
Artinian ideal
Noncommutative ring
Regular ideal
Idealizer
30 CHAPTER 6. IDEAL (RING THEORY)
6.13 References
[1] Harold M. Edwards (1977). Fermats last theorem. A genetic introduction to algebraic number theory. p. 76.
[4] In fact, since R is assumed to be unital, it suces that x + y is in I, since the second condition implies that y is in I.
[6] Because simple commutative rings are elds. See Lam (2001). A First Course in Noncommutative Rings. p. 39.
Lang, Serge (2005). Undergraduate Algebra (Third ed.). Springer-Verlag. ISBN 978-0-387-22025-3
Michiel Hazewinkel, Nadiya Gubareni, Nadezhda Mikhalovna Gubareni, Vladimir V. Kirichenko. Algebras,
rings and modules. Volume 1. 2004. Springer, 2004. ISBN 1-4020-2690-0
Chapter 7
In the mathematical eld of set theory, an ideal is a collection of sets that are considered to be small or negligible.
Every subset of an element of the ideal must also be in the ideal (this codies the idea that an ideal is a notion of
smallness), and the union of any two elements of the ideal must also be in the ideal.
More formally, given a set X, an ideal I on X is a nonempty subset of the powerset of X, such that:
1. I
2.if A I and B A , then B I , and
3.if A, B I , then A B I .
Some authors add a third condition that X itself is not in I; ideals with this extra property are called proper ideals.
Ideals in the set-theoretic sense are exactly ideals in the order-theoretic sense, where the relevant order is set inclusion.
Also, they are exactly ideals in the ring-theoretic sense on the Boolean ring formed by the powerset of the underlying
set.
7.1 Terminology
An element of an ideal I is said to be I-null or I-negligible, or simply null or negligible if the ideal I is understood from
context. If I is an ideal on X, then a subset of X is said to be I-positive (or just positive) if it is not an element of I.
The collection of all I-positive subsets of X is denoted I + .
31
32 CHAPTER 7. IDEAL (SET THEORY)
The ideal of asymptotically zero-density sets on the natural numbers, denoted Z0 , is the collection of all sets
A of natural numbers such that the fraction of natural numbers less than n that belong to A, tends to zero as n
tends to innity. (That is, the asymptotic density of A is zero.)
A I J {x X|{y|x, y A}
/ J} I
That is, a set is negligible in the product ideal if only a negligible collection of x-coordinates correspond to a non-
negligible slice of A in the y-direction. (Perhaps clearer: A set is positive in the product ideal if positively many
x-coordinates correspond to positive slices.)
An ideal I on a set X induces an equivalence relation on P(X), the powerset of X, considering A and B to be equivalent
(for A, B subsets of X) if and only if the symmetric dierence of A and B is an element of I. The quotient of P(X) by
this equivalence relation is a Boolean algebra, denoted P(X) / I (read P of X mod I").
To every ideal there is a corresponding lter, called its dual lter. If I is an ideal on X, then the dual lter of I is the
collection of all sets X \ A, where A is an element of I. (Here X \ A denotes the relative complement of A in X; that
is, the collection of all elements of X that are not in A.)
7.6 References
Farah, Ilijas (November 2000). Analytic quotients: Theory of liftings for quotients over analytic ideals on the
integers. Memoirs of the AMS. American Mathematical Society.
Chapter 8
Maximal element
In mathematics, especially in order theory, a maximal element of a subset S of some partially ordered set (poset) is
an element of S that is not smaller than any other element in S. A minimal element of a subset S of some partially
ordered set is dened dually as an element of S that is not greater than any other element in S.
The notions of maximal and minimal elements are weaker than those of greatest element and least element which are
also known, respectively, as maximum and minimum. The maximum of a subset S of a partially ordered set is an
element of S which is greater than or equal to any other element of S, and the minimum of S is again dened dually.
While a partially ordered set can have at most one each maximum and minimum it may have multiple maximal and
minimal elements.[1][2] For totally ordered sets, the notions of maximal element and maximum coincide, and the
notions of minimal element and minimum coincide.
As an example, in the collection
ordered by containment, the element {d, o} is minimal, the element {g, o, a, d} is maximal, the element {d, o, g} is
neither, and the element {o, a, f} is both minimal and maximal. By contrast, neither a maximum nor a minimum
exists for S.
Zorns lemma states that every partially ordered set for which every totally ordered subset has an upper bound contains
at least one maximal element. This lemma is equivalent to the well-ordering theorem and the axiom of choice[3]
and implies major results in other mathematical areas like the HahnBanach theorem and Tychonos theorem, the
existence of a Hamel basis for every vector space, and the existence of an algebraic closure for every eld.
8.1 Denition
Let (P, ) be a partially ordered set and S P . Then m S is a maximal element of S if
for all s S , m s implies m = s.
The denition for minimal elements is obtained by using instead of .
Example 1: Let S = [1,) , for all mS we have s=m+1S but m<s (that is, ms but not m=s).
Example 2: Let S = {s: 1s2 2} and recall that 2 .
In general is only a partial order on S. If m is a maximal element and sS, it remains the possibility that neither sm
nor ms. This leaves open the possibility that there are many maximal elements.
33
34 CHAPTER 8. MAXIMAL ELEMENT
Example 3: In the fence a1 < b1 > a2 < b2 > a3 < b3 > ..., all the ai are minimal, and all the bi are
maximal, see picture.
Example 4: Let A be a set with at least two elements and let S={{a}: aA} be the subset of the power
set P(A) consisting of singletons, partially ordered by . This is the discrete posetno two elements are
comparableand thus every element {a}S is maximal (and minimal) and for any distinct a,a neither
{a} {a} nor {a} {a}.
{x,y,z}
Power set of {x,y,z}, partially ordered by . Its greatest element, {x,y,z}, is its only maximal element.
8.5 Examples
In Pareto eciency, a Pareto optimum is a maximal element with respect to the partial order of Pareto
improvement, and the set of maximal elements is called the Pareto frontier.
In decision theory, an admissible decision rule is a maximal element with respect to the partial order of
dominating decision rule.
In modern portfolio theory, the set of maximal elements with respect to the product order on risk and return
is called the ecient frontier.
In set theory, a set is nite if and only if every non-empty family of subsets has a minimal element when ordered
by the inclusion relation.
In abstract algebra, the concept of a maximal common divisor is needed to generalize greatest common divisors
to number systems in which the common divisors of a set of elements may have more than one maximal element.
In computational geometry, the maxima of a point set are maximal with respect to the partial order of coordi-
natewise domination.
In consumer theory the consumption space is some set X , usually the positive orthant of some vector space so that
each x X represents a quantity of consumption specied for each existing commodity in the economy. Preferences
of a consumer are usually represented by a total preorder so that x, y X and x y reads: x is at most as preferred
as y . When x y and y x it is interpreted that the consumer is indierent between x and y but is no reason to
conclude that x = y , preference relations are never assumed to be antisymmetric. In this context, for any B X ,
we call x B a maximal element if
y B implies y x
and it is interpreted as a consumption bundle that is not dominated by any other bundle in the sense that x y , that
is x y and not y x .
It should be remarked that the formal denition looks very much like that of a greatest element for an ordered set.
However, when is only a preorder, an element x with the property above behaves very much like a maximal element
in an ordering. For instance, a maximal element x B is not unique for y x does not preclude the possibility that
x y (while y x and x y do not imply x = y but simply indierence x y ). The notion of greatest element
for a preference preorder would be that of most preferred choice. That is, some x B with
y B implies y x.
An obvious application is to the denition of demand correspondence. Let P be the class of functionals on X . An
element p P is called a price functional or price system and maps every consumption bundle x X into its
market value p(x) R+ . The budget correspondence is a correspondence : P R+ X mapping any price
system and any level of income into a subset
The demand correspondence maps any price p and any level of income m into the set of -maximal elements of
(p, m) .
{ }
D(p, m) = x X | x is a maximal element of (p, m) .
It is called demand correspondence because the theory predicts that for p and m given, the rational choice of a
consumer x will be some element x D(p, m) .
8.7 References
[1] Richmond, Bettina; Richmond, Thomas (2009), A Discrete Transition to Advanced Mathematics, American Mathematical
Society, p. 181, ISBN 978-0-8218-4789-3.
[2] Scott, William Raymond (1987), Group Theory (2nd ed.), Dover, p. 22, ISBN 978-0-486-65377-8
[3] Jech, Thomas (2008) [originally published in 1973]. The Axiom of Choice. Dover Publications. ISBN 0-486-46624-8.
Chapter 9
Maximal ideal
In mathematics, more specically in ring theory, a maximal ideal is an ideal that is maximal (with respect to set
inclusion) amongst all proper ideals.[1][2] In other words, I is a maximal ideal of a ring R if there are no other ideals
contained between I and R.
Maximal ideals are important because the quotient rings of maximal ideals are simple rings, and in the special case
of unital commutative rings they are also elds.
In noncommutative ring theory, a maximal right ideal is dened analogously as being a maximal element in the
poset of proper right ideals, and similarly, a maximal left ideal is dened to be a maximal element of the poset of
proper left ideals. Since a one sided maximal ideal A is not necessarily two-sided, the quotient R/A is not necessarily
a ring, but it is a simple module over R. If R has a unique maximal right ideal, then R is known as a local ring, and
the maximal right ideal is also the unique maximal left and unique maximal two-sided ideal of the ring, and is in fact
the Jacobson radical J(R).
It is possible for a ring to have a unique maximal ideal and yet lack unique maximal one sided ideals: for example,
in the ring of 2 by 2 square matrices over a eld, the zero ideal is a maximal ideal, but there are many maximal right
ideals.
9.1 Denition
There are other equivalent ways of expressing the denition of maximal one-sided and maximal two-sided ideals.
Given a ring R and a proper ideal I of R (that is I R), I is a maximal ideal of R if any of the following equivalent
conditions hold:
There is an analogous list for one-sided ideals, for which only the right-hand versions will be given. For a right ideal
A of a ring R, the following conditions are equivalent to A being a maximal right ideal of R:
Maximal right/left/two-sided ideals are the dual notion to that of minimal ideals.
37
38 CHAPTER 9. MAXIMAL IDEAL
9.2 Examples
The maximal ideals of the polynomial ring C[x] are principal ideals generated by x c for some c C .
In the ring Z of integers, the maximal ideals are the principal ideals generated by a prime number.
More generally, all nonzero prime ideals are maximal in a principal ideal domain.
The maximal ideals of the polynomial ring K[x1 ,...,xn] over an algebraically closed eld K are the ideals of the
form (x1 a1 ,...,xn an). This result is known as the weak nullstellensatz.
9.3 Properties
An important ideal of the ring called the Jacobson radical can be dened using maximal right (or maximal left)
ideals.
If R is a unital commutative ring with an ideal m, then k = R/m is a eld if and only if m is a maximal ideal.
In that case, R/m is known as the residue eld. This fact can fail in non-unital rings. For example, 4Z is a
maximal ideal in 2Z , but 2Z/4Z is not a eld.
If L is a maximal left ideal, then R/L is a simple left R module. Conversely in rings with unity, any simple left
R module arises this way. Incidentally this shows that a collection of representatives of simple left R modules
is actually a set since it can be put into correspondence with part of the set of maximal left ideals of R.
Krulls theorem (1929): Every nonzero unital ring has a maximal ideal. The result is also true if ideal is
replaced with right ideal or left ideal. More generally, it is true that every nonzero nitely generated module
has a maximal submodule. Suppose I is an ideal which is not R (respectively, A is a right ideal which is not
R). Then R/I is a ring with unity, (respectively, R/A is a nitely generated module), and so the above theorems
can be applied to the quotient to conclude that there is a maximal ideal (respectively maximal right ideal) of R
containing I (respectively, A).
Krulls theorem can fail for rings without unity. A radical ring, i.e. a ring in which the Jacobson radical is the
entire ring, has no simple modules and hence has no maximal right or left ideals. See regular ideals for possible
ways to circumvent this problem.
In a commutative ring with unity, every maximal ideal is a prime ideal. The converse is not always true: for
example, in any noneld integral domain the zero ideal is a prime ideal which is not maximal. Commutative
rings in which prime ideals are maximal are known as zero-dimensional rings, where the dimension used is the
Krull dimension.
9.4 Generalization
For an R module A, a maximal submodule M of A is a submodule MA for which for any other submodule N, if
MNA then N=M or N=A. Equivalently, M is a maximal submodule if and only if the quotient module A/M is a
simple module. Clearly the maximal right ideals of a ring R are exactly the maximal submodules of the module RR.
Unlike rings with unity however, a module does not necessarily have maximal submodules. However, as noted above,
nitely generated nonzero modules have maximal submodules, and also projective modules have maximal submodules.
As with rings, one can dene the radical of a module using maximal submodules.
Furthermore, maximal ideals can be generalized by dening a maximal sub-bimodule M of a bimodule B to be a
proper sub-bimodule of M which is contained by no other proper sub-bimodule of M. So, the maximal ideals of R
are exactly the maximal sub-bimodules of the bimodule RRR.
9.5. REFERENCES 39
9.5 References
[1] Dummit, David S.; Foote, Richard M. (2004). Abstract Algebra (3rd ed.). John Wiley & Sons. ISBN 0-471-43334-9.
[2] Lang, Serge (2002). Algebra. Graduate Texts in Mathematics. Springer. ISBN 0-387-95385-X.
Anderson, Frank W.; Fuller, Kent R. (1992), Rings and categories of modules, Graduate Texts in Mathematics,
13 (2 ed.), New York: Springer-Verlag, pp. x+376, ISBN 0-387-97845-3, MR 1245487
Lam, T. Y. (2001), A rst course in noncommutative rings, Graduate Texts in Mathematics, 131 (2 ed.), New
York: Springer-Verlag, pp. xx+385, ISBN 0-387-95183-0, MR 1838439
Chapter 10
Order theory
Order theory is a branch of mathematics which investigates the intuitive notion of order using binary relations. It
provides a formal framework for describing statements such as this is less than that or this precedes that. This
article introduces the eld and provides basic denitions. A list of order-theoretic terms can be found in the order
theory glossary.
Orders are everywhere in mathematics and related elds like computer science. The rst order often discussed in
primary school is the standard order on the natural numbers e.g. 2 is less than 3, 10 is greater than 5, or Does
Tom have fewer cookies than Sally?". This intuitive concept can be extended to orders on other sets of numbers,
such as the integers and the reals. The idea of being greater than or less than another number is one of the basic
intuitions of number systems (compare with numeral systems) in general (although one usually is also interested in
the actual dierence of two numbers, which is not given by the order). Other familiar examples of orderings are the
alphabetical order of words in a dictionary and the genealogical property of lineal descent within a group of people.
The notion of order is very general, extending beyond contexts that have an immediate, intuitive feel of sequence
or relative quantity. In other contexts orders may capture notions of containment or specialization. Abstractly, this
type of order amounts to the subset relation, e.g., "Pediatricians are physicians, and "Circles are merely special-case
ellipses.
Some orders, like less-than on the natural numbers and alphabetical order on words, have a special property: each
element can be compared to any other element, i.e. it is smaller (earlier) than, larger (later) than, or identical to.
However, many other orders do not. Consider for example the subset order on a collection of sets: though the set of
birds and the set of dogs are both subsets of the set of animals, neither the birds nor the dogs constitutes a subset of
the other. Those orders like the subset-of relation for which there exist incomparable elements are called partial
orders; orders for which every pair of elements is comparable are total orders.
Order theory captures the intuition of orders that arises from such examples in a general setting. This is achieved
by specifying properties that a relation must have to be a mathematical order. This more abstract approach makes
much sense, because one can derive numerous theorems in the general setting, without focusing on the details of any
particular order. These insights can then be readily transferred to many less abstract applications.
Driven by the wide practical usage of orders, numerous special kinds of ordered sets have been dened, some of
which have grown into mathematical elds of their own. In addition, order theory does not restrict itself to the
various classes of ordering relations, but also considers appropriate functions between them. A simple example of an
order theoretic property for functions comes from analysis where monotone functions are frequently found.
40
10.2. BASIC DEFINITIONS 41
a a (reexivity)
if a b and b a then a = b (antisymmetry)
if a b and b c then a c (transitivity).
A set with a partial order on it is called a partially ordered set, poset, or just an ordered set if the intended meaning
is clear. By checking these properties, one immediately sees that the well-known orders on natural numbers, integers,
rational numbers and reals are all orders in the above sense. However, they have the additional property of being
total, i.e., for all a and b in P, we have that:
a b or b a (totality).
These orders can also be called linear orders or chains. While many classical orders are linear, the subset order on
sets provides an example where this is not the case. Another example is given by the divisibility (or "is-a-factor-of")
relation "|". For two natural numbers n and m, we write n|m if n divides m without remainder. One easily sees that
this yields a partial order. The identity relation = on any set is also a partial order in which every two distinct elements
are incomparable. It is also the only relation that is both a partial order and an equivalence relation. Many advanced
properties of posets are interesting mainly for non-linear orders.
The notation 0 is frequently found for the least element, even when no numbers are concerned. However, in orders
on sets of numbers, this notation might be inappropriate or ambiguous, since the number 0 is not always least. An
example is given by the above divisibility order |, where 1 is the least element since it divides all other numbers. In
contrast, 0 is the number that is divided by all other numbers. Hence it is the greatest element of the order. Other
frequent terms for the least and greatest elements is bottom and top or zero and unit.
Least and greatest elements may fail to exist, as the example of the real numbers shows. But if they exist, they are
always unique. In contrast, consider the divisibility relation | on the set {2,3,4,5,6}. Although this set has neither top
42 CHAPTER 10. ORDER THEORY
60
12
30 20
15 10 6 4
3 2
5
1
Hasse diagram of the set of all divisors of 60, partially ordered by divisibility
nor bottom, the elements 2, 3, and 5 have no elements below them, while 4, 5 and 6 have none above. Such elements
are called minimal and maximal, respectively. Formally, an element m is minimal if:
Exchanging with yields the denition of maximality. As the example shows, there can be many maximal elements
and some elements may be both maximal and minimal (e.g. 5 above). However, if there is a least element, then it
is the only minimal element of the order. Again, in innite posets maximal elements do not always exist - the set
of all nite subsets of a given innite set, ordered by subset inclusion, provides one of many counterexamples. An
important tool to ensure the existence of maximal elements under certain conditions is Zorns Lemma.
Subsets of partially ordered sets inherit the order. We already applied this by considering the subset {2,3,4,5,6} of
the natural numbers with the induced divisibility ordering. Now there are also elements of a poset that are special
with respect to some subset of the order. This leads to the denition of upper bounds. Given a subset S of some
poset P, an upper bound of S is an element b of P that is above all elements of S. Formally, this means that
s b, for all s in S.
Lower bounds again are dened by inverting the order. For example, 5 is a lower bound of the natural numbers
as a subset of the integers. Given a set of sets, an upper bound for these sets under the subset ordering is given by
their union. In fact, this upper bound is quite special: it is the smallest set that contains all of the sets. Hence, we
have found the least upper bound of a set of sets. This concept is also called supremum or join, and for a set S
one writes sup(S) or S forits least upper bound. Conversely, the greatest lower bound is known as inmum or
meet and denoted inf(S) or S . These concepts play an important role in many applications of order theory. For
two elements x and y, one also writes x y and x y for sup({x,y}) and inf({x,y}), respectively.
For example, 1 is the inmum of the positive integers as a subset of integers.
10.3. FUNCTIONS BETWEEN ORDERS 43
For another example, consider again the relation | on natural numbers. The least upper bound of two numbers is the
smallest number that is divided by both of them, i.e. the least common multiple of the numbers. Greatest lower
bounds in turn are given by the greatest common divisor.
10.2.4 Duality
In the previous denitions, we often noted that a concept can be dened by just inverting the ordering in a former
denition. This is the case for least and greatest, for minimal and maximal, for upper bound and lower
bound, and so on. This is a general situation in order theory: A given order can be inverted by just exchanging
its direction, pictorially ipping the Hasse diagram top-down. This yields the so-called dual, inverse, or opposite
order.
Every order theoretic denition has its dual: it is the notion one obtains by applying the denition to the inverse order.
Since all concepts are symmetric, this operation preserves the theorems of partial orders. For a given mathematical
result, one can just invert the order and replace all denitions by their duals and one obtains another valid theorem.
This is important and useful, since one obtains two theorems for the price of one. Some more details and examples
can be found in the article on duality in order theory.
There are many ways to construct orders out of given orders. The dual order is one example. Another important
construction is the cartesian product of two partially ordered sets, taken together with the product order on pairs of
elements. The ordering is dened by (a, x) (b, y) if (and only if) a b and x y. (Notice carefully that there
are three distinct meanings for the relation symbol in this denition.) The disjoint union of two posets is another
typical example of order construction, where the order is just the (disjoint) union of the original orders.
Every partial order gives rise to a so-called strict order <, by dening a < b if a b and not b a. This transformation
can be inverted by setting a b if a < b or a = b. The two concepts are equivalent although in some circumstances
one can be more convenient to work with than the other.
to special elements and constructions. For example, when talking about posets with least element, it may seem
reasonable to consider only monotonic functions that preserve this element, i.e. which map least elements to least
elements. If binary inma exist, then a reasonable property might be to require that f(x y) = f(x) f(y), for all x
and y. All of these properties, and indeed many more, may be compiled under the label of limit-preserving functions.
Finally, one can invert the view, switching from functions of orders to orders of functions. Indeed, the functions
between two posets P and Q can be ordered via the pointwise order. For two functions f and g, we have f g if f(x)
g(x) for all elements x of P. This occurs for example in domain theory, where function spaces play an important
role.
Bounded posets, i.e. posets with a least and greatest element (which are just the supremum and inmum of
the empty subset),
Lattices, in which every non-empty nite set has a supremum and inmum,
Complete lattices, where every set has a supremum and inmum, and
Directed complete partial orders (dcpos), that guarantee the existence of suprema of all directed subsets and
that are studied in domain theory.
Partial orders with complements, or poc sets,[1] are posets S having a unique bottom element 0S, along with
an order-reversing involution, such that a a a = 0 .
However, one can go even further: if all nite non-empty inma exist, then can be viewed as a total binary operation
in the sense of universal algebra. Hence, in a lattice, two operations and are available, and one can dene new
properties by giving identities, such as
This condition is called distributivity and gives rise to distributive lattices. There are some other important distribu-
tivity laws which are discussed in the article on distributivity in order theory. Some additional order structures that
are often specied via algebraic operations and dening identities are
which both introduce a new operation ~ called negation. Both structures play a role in mathematical logic and
especially Boolean algebras have major applications in computer science. Finally, various structures in mathematics
10.5. SUBSETS OF ORDERED SETS 45
combine orders with even more algebraic operations, as in the case of quantales, that allow for the denition of an
addition operation.
Many other important properties of posets exist. For example, a poset is locally nite if every closed interval [a,
b] in it is nite. Locally nite posets give rise to incidence algebras which in turn can be used to dene the Euler
characteristic of nite bounded posets.
10.6.2 Topology
In topology orders play a very prominent role. In fact, the set of open sets provides a classical example of a complete
lattice, more precisely a complete Heyting algebra (or "frame" or "locale"). Filters and nets are notions closely related
to order theory and the closure operator of sets can be used to dene topology. Beyond these relations, topology can
be looked at solely in terms of the open set lattices, which leads to the study of pointless topology. Furthermore, a
natural preorder of elements of the underlying set of a topology is given by the so-called specialization order, that is
actually a partial order if the topology is T0 .
Conversely, in order theory, one often makes use of topological results. There are various ways to dene subsets of
an order which can be considered as open sets of a topology. Especially, it is interesting to consider topologies on a
poset (X, ) that in turn induce as their specialization order. The nest such topology is the Alexandrov topology,
given by taking all upper sets as opens. Conversely, the coarsest topology that induces the specialization order is the
upper topology, having the complements of principal ideals (i.e. sets of the form {y in X | y x} for some x) as a
subbase. Additionally, a topology with specialization order may be order consistent, meaning that their open sets
are inaccessible by directed suprema (with respect to ). The nest order consistent topology is the Scott topology,
which is coarser than the Alexandrov topology. A third important topology in this spirit is the Lawson topology.
There are close connections between these topologies and the concepts of order theory. For example, a function
preserves directed suprema i it is continuous with respect to the Scott topology (for this reason this order theoretic
property is also called Scott-continuity).
46 CHAPTER 10. ORDER THEORY
The visualization of orders with Hasse diagrams has a straightforward generalization: instead of displaying lesser
elements below greater ones, the direction of the order can also be depicted by giving directions to the edges of a
graph. In this way, each order is seen to be equivalent to a directed acyclic graph, where the nodes are the elements
of the poset and there is a directed path from a to b if and only if a b. Dropping the requirement of being acyclic,
one can also obtain all preorders.
When equipped with all transitive edges, these graphs in turn are just special categories, where elements are objects
and each set of morphisms between two elements is at most singleton. Functions between orders become functors
between categories. Interestingly, many ideas of order theory are just concepts of category theory in small. For
example, an inmum is just a categorical product. More generally, one can capture inma and suprema under the
abstract notion of a categorical limit (or colimit, respectively). Another place where categorical ideas occur is the
concept of a (monotone) Galois connection, which is just the same as a pair of adjoint functors.
But category theory also has its impact on order theory on a larger scale. Classes of posets with appropriate functions
as discussed above form interesting categories. Often one can also state constructions of orders, like the product
order, in terms of categories. Further insights result when categories of orders are found categorically equivalent to
other categories, for example of topological spaces. This line of research leads to various representation theorems,
often collected under the label of Stone duality.
10.7 History
As explained before, orders are ubiquitous in mathematics. However, earliest explicit mentionings of partial orders are
probably to be found not before the 19th century. In this context the works of George Boole are of great importance.
Moreover, works of Charles Sanders Peirce, Richard Dedekind, and Ernst Schrder also consider concepts of order
theory. Certainly, there are others to be named in this context and surely there exists more detailed material on the
history of order theory.
The term poset as an abbreviation for partially ordered set was coined by Garrett Birkho in the second edition of
his inuential book Lattice Theory.[2][3]
Hierarchy
Incidence algebra
Causal Sets
10.9 Notes
[1] Roller, Martin A. (1998), Poc sets, median algebras and group actions. An extended study of Dunwoodys construction and
Sageevs theorem (PDF), Southampton Preprint Archive
10.10 References
Birkho, Garrett (1940). Lattice Theory. 25 (3rd Revised ed.). American Mathematical Society. ISBN
978-0-8218-1025-5.
10.11. EXTERNAL LINKS 47
Burris, S. N.; Sankappanavar, H. P. (1981). A Course in Universal Algebra. Springer. ISBN 978-0-387-90578-
5.
Davey, B. A.; Priestley, H. A. (2002). Introduction to Lattices and Order (2nd ed.). Cambridge University
Press. ISBN 0-521-78451-4.
Gierz, G.; Hofmann, K. H.; Keimel, K.; Mislove, M.; Scott, D. S. (2003). Continuous Lattices and Domains.
Encyclopedia of Mathematics and its Applications. 93. Cambridge University Press. ISBN 978-0-521-80338-
0.
Ultralter
This article is about the mathematical concept. For the physical device, see ultraltration.
In the mathematical eld of set theory, an ultralter on a given partially ordered set (poset) P is a maximal lter
1, 2, 3, 4
1, 2, 3 1, 2, 4 1, 3, 4 2, 3, 4
1, 2 1, 3 1, 4 2, 3 2, 4 3, 4
1 2 3 4
The powerset lattice of the set {1,2,3,4}, with the upper set {1,4} colored yellow. It is a principal lter, but not an ultralter, as
it can be extended to the larger nontrivial lter {1}, by including also the light green elements. Since {1} cannot be extended any
further, it is an ultralter.
on P, that is, a lter on P that cannot be enlarged. Filters and ultralters are special subsets of P. If P happens to be
a Boolean algebra, each ultralter is also a prime lter, and vice versa.[1]:186
If X is an arbitrary set, its power set (X), ordered by set inclusion, is always a Boolean algebra, and (ultra)lters on
(X) are usually called "(ultra)lters on X".[note 1] Ultralters have many applications in set theory, model theory, and
topology.[1]:186 An ultralter on a set X may be considered as a nitely additive measure. In this view, every subset
of X is either considered "almost everything" (has measure 1) or almost nothing (has measure 0).
48
11.1. ULTRAFILTERS ON PARTIAL ORDERS 49
1. F is an ultralter on P,
2. F is a prime lter on P,
3. for each a in P, either a is in F or (a) is in F.[1]:186
Given a homomorphism of a Boolean algebra onto {true, false}, the inverse image of true is an ultralter,
and the inverse image of false is a maximal ideal.
Given a maximal ideal of a Boolean algebra, its complement is an ultralter, and there is a unique homomor-
phism onto {true, false} taking the maximal ideal to false.
Given an ultralter of a Boolean algebra, its complement is a maximal ideal, and there is a unique homomor-
phism onto {true, false} taking the ultralter to true.
A characterization is given by the following theorem. A lter U on (X) is an ultralter if any of the following
conditions are true:
Another way of looking at ultralters on a power set (X) is to dene a function m on (X) by setting m(A) = 1 if
A is an element of U and m(A) = 0 otherwise. Such a function is called a 2-valued morphism. Then m is nitely
additive, and hence a content on (X), and every property of elements of X is either true almost everywhere or false
almost everywhere. However, m is usually not countably additive, and hence does not dene a measure in the usual
sense.
For a lter F that is not an ultralter, one would say m(A) = 1 if A F and m(A) = 0 if X \ A F, leaving m undened
elsewhere.
11.4 Completeness
The completeness of an ultralter U on a powerset is the smallest cardinal such that there are elements of U
whose intersection is not in U. The denition implies that the completeness of any powerset ultralter is at least 0 .
An ultralter whose completeness is greater than 0 that is, the intersection of any countable collection of elements
of U is still in Uis called countably complete or -complete.
The completeness of a countably complete nonprincipal ultralter on a powerset is always a measurable cardinal.
11.6 Applications
Ultralters on powersets are useful in topology, especially in relation to compact Hausdor spaces, and in model
theory in the construction of ultraproducts and ultrapowers. Every ultralter on a compact Hausdor space converges
to exactly one point. Likewise, ultralters on Boolean algebras play a central role in Stones representation theorem.
11.7. ORDERING ON ULTRAFILTERS 51
The set G of all ultralters of a poset P can be topologized in a natural way, that is in fact closely related to the above-
mentioned representation theorem. For any element a of P, let Da = {U G | a U}. This is most useful when P
is again a Boolean algebra, since in this situation the set of all Da is a base for a compact Hausdor topology on G.
Especially, when considering the ultralters on a powerset (S), the resulting topological space is the Stoneech
compactication of a discrete space of cardinality |S|.
The ultraproduct construction in model theory uses ultralters to produce elementary extensions of structures. For
example, in constructing hyperreal numbers as an ultraproduct of the real numbers, the domain of discourse is ex-
tended from real numbers to sequences of real numbers. This sequence space is regarded as a superset of the reals by
identifying each real with the corresponding constant sequence. To extend the familiar functions and relations (e.g.,
+ and <) from the reals to the hyperreals, the natural idea is to dene them pointwise. But this would lose important
logical properties of the reals; for example, pointwise < is not a total ordering. So instead the functions and relations
are dened "pointwise modulo U", where U is an ultralter on the index set of the sequences; by o' theorem, this
preserves all properties of the reals that can be stated in rst-order logic. If U is nonprincipal, then the extension
thereby obtained is nontrivial.
In geometric group theory, non-principal ultralters are used to dene the asymptotic cone of a group. This con-
struction yields a rigorous way to consider looking at the group from innity, that is the large scale geometry of the
group. Asymptotic cones are particular examples of ultralimits of metric spaces.
Gdels ontological proof of Gods existence uses as an axiom that the set of all positive properties is an ultralter.
In social choice theory, non-principal ultralters are used to dene a rule (called a social welfare function) for ag-
gregating the preferences of innitely many individuals. Contrary to Arrows impossibility theorem for nitely many
individuals, such a rule satises the conditions (properties) that Arrow proposes (e.g., Kirman and Sondermann,
1972[3] ). Mihara (1997,[4] 1999[5] ) shows, however, such rules are practically of limited interest to social scientists,
since they are non-algorithmic or non-computable.
C V f 1 [C] U
11.8 Ultralters on ()
There are several special properties that an ultralter on () may possess, which prove useful in various areas of
set theory and topology.
A non-principal ultralter U is called a P-point (or weakly selective) if for every partition { Cn | n< } of
such that n<: Cn U, there exists some A U such that A Cn is a nite set for each n.
A non-principal ultralter U is called Ramsey (or selective) if for every partition { Cn | n< } of such that
n<: Cn U, there exists some A U such that A Cn is a singleton set for each n.
It is a trivial observation that all Ramsey ultralters are P-points. Walter Rudin proved that the continuum hypothesis
implies the existence of Ramsey ultralters.[6] In fact, many hypotheses imply the existence of Ramsey ultralters,
including Martins axiom. Saharon Shelah later showed that it is consistent that there are no P-point ultralters.[7]
Therefore, the existence of these types of ultralters is independent of ZFC.
52 CHAPTER 11. ULTRAFILTER
P-points are called as such because they are topological P-points in the usual topology of the space \ of non-
principal ultralters. The name Ramsey comes from Ramseys theorem. To see why, one can prove that an ultralter
is Ramsey if and only if for every 2-coloring of []2 there exists an element of the ultralter that has a homogeneous
color.
An ultralter on () is Ramsey if and only if it is minimal in the RudinKeisler ordering of non-principal powerset
ultralters.
11.10 Notes
[1] If X happens to be partially ordered, too, particular care is needed to understand from the context whether an (ultra)lter
on (X) or an (ultra)lter just on X is meant; both kinds of (ultra)lters are quite dierent. Some authors use "(ultra)lter
of a partial ordered set vs. "on an arbitrary set"; i.e. they write "(ultra)lter on X" to abbreviate "(ultra)lter of (X)".
[3] To see the if direction: If {s1 ,...,sn} U, then {s1 } U, or ..., or {sn} U by induction on n, using Nr.2 of the above
characterization theorem. That is, some {si} is the principal element of U.
[4] U is non-principal i it contains no nite set, i.e. (by Nr.3 of the above characterization theorem) i it contains every
conite set, i.e. every member of the Frchet lter.
11.11 References
[1] B.A. Davey and H.A. Priestley (1990). Introduction to Lattices and Order. Cambridge Mathematical Textbooks. Cambridge
University Press.
[2] Stanley N. Burris and H.P. Sankappanavar (2012). A Course in Universal Algebra (PDF). ISBN 978-0-9880552-0-9.
[3] Kirman, A.; Sondermann, D. (1972). Arrows theorem, many agents, and invisible dictators. Journal of Economic
Theory. 5: 267. doi:10.1016/0022-0531(72)90106-8.
[4] Mihara, H. R. (1997). Arrows Theorem and Turing computability (PDF). Economic Theory. 10 (2): 257276. doi:10.1007/s001990050157R
in K. V. Velupillai , S. Zambelli, and S. Kinsella, ed., Computable Economics, International Library of Critical Writings
in Economics, Edward Elgar, 2011.
[5] Mihara, H. R. (1999). Arrows theorem, countably many agents, and more visible invisible dictators. Journal of Mathe-
matical Economics. 32: 267277. doi:10.1016/S0304-4068(98)00061-5.
[6] Rudin, Walter (1956), Homogeneity problems in the theory of ech compactications, Duke Mathematical Journal, 23
(3): 409419, doi:10.1215/S0012-7094-56-02337-7
[7] Wimmers, Edward (March 1982), The Shelah P-point independence theorem, Israel Journal of Mathematics, Hebrew
University Magnes Press, 43 (1): 2848, doi:10.1007/BF02761683
Algebraist, YurikBot, Wasseralm, SmackBot, Andyvn22, Nbarth, J. Finkelstein, JAnDbot, David Eppstein, Wlodzimierz, Erhasalz, Lo-
kiClock, Leafyplant, PaulTanenbaum, Xurxo.duran, Die Mensch-Maschine, SetaLyas, Harshal7788, Addbot, Luckas-bot, AnomieBOT,
Xqbot, Makraim, Erik9bot, Jkock, Alihzubi, Alph Bot, EmausBot, ZroBot, Helpful Pixie Bot, RMCD bot, Jochen Burghardt, Loraof,
Bender the Bot, Octonious and Anonymous: 23
Maximal ideal Source: https://en.wikipedia.org/wiki/Maximal_ideal?oldid=755083938 Contributors: Toby~enwiki, Michael Hardy,
TakuyaMurata, Charles Matthews, Chentianran~enwiki, MathMartin, SimonMayer, Giftlite, Markus Krtzsch, Lupin, Ssd, Vivacis-
samamente, Smimram, El C, EmilJ, Culix, HannsEwald, Chobot, Saintali, Hairy Dude, Grubber, DomenicDenicola, Reyk, SmackBot,
Mipchunk, Viebel, JoshuaZ, Jim.belk, Rschwieb, Wafulz, Xantharius, Kprateek88, Sullivan.t.j, Error792, Flyer22 Reborn, Marsupilamov,
Addbot, Fraggle81, TaBOT-zerem, KamikazeBot, Point-set topologist, Erik9bot, Kiefer.Wolfowitz, DixonDBot, Slawekb, ZroBot, Max-
imalIdeal, BG19bot, Connor.sempek and Anonymous: 28
Order theory Source: https://en.wikipedia.org/wiki/Order_theory?oldid=796416278 Contributors: Bryan Derksen, Toby Bartels, Michael
Hardy, Dineshjk, Ehn, Charles Matthews, Dcoetzee, Jitse Niesen, Wik, Natevw, VeryVerily, Populus, Topbanana, Robbot, Henrygb,
ElBenevolente, Tea2min, Giftlite, Markus Krtzsch, Elias, DefLog~enwiki, Yarnover, APH, SimonLyall, Xrchz, Abar, Pjacobi, Paul Au-
gust, Tompw, Nickj, Themusicgod1, Lysdexia, Arthena, Joriki, Linas, Je3000, Josh Parris, Salix alba, Mathbot, Hairy Dude, Dmharvey,
Trovatore, Yahya Abdal-Aziz, Ott2, Arthur Rubin, Modify, Netrapt, That Guy, From That Show!, SmackBot, KnowledgeOfSelf, Mhss,
Bluebot, RDBrown, Cybercobra, Kntrabssi, Dreadstar, JohnI, 16@r, Loadmaster, Dicklyon, Landonproctor, Levineps, Dreftymac, Ma-
jora4, CRGreathouse, CBM, Sam Staton, Skittleys, Marek69, Ankit mcgill, MER-C, VoABot II, Cic, David Eppstein, Gwern, Maurice
Carbonaro, Inquam, Daniel5Ko, Jorfer, JohnBlackburne, Trondarild, Philip Trueman, GcSwRhIc, The Tetrast, Magmi, PaulTanenbaum,
Geometry guy, Wiae, Tomaxer, StevenJohnston, AS, Anchor Link Bot, Randomblue, ClueBot, Justin W Smith, Hans Adler, Wikidsp, Ad-
dbot, Barak Sh, Badou517, Legobot, AnomieBOT, Sawyeriii, Buenasdiaz, Smallman12q, Tuetschek, FrescoBot, Mark Renier, Orhang-
hazi, Chenopodiaceous, Gamewizard71, Genezistan, John of Reading, Dadaist6174, ClueBot NG, Syamino, MerlIwBot, Helpful Pixie
Bot, Knwlgc, VolunBute, Brad7777, Toploftical, Nicuchalan1, Rupert loup, K401sTL3, JaconaFrere, Srlgator, Pheello87, KasparBot and
Anonymous: 64
Ultralter Source: https://en.wikipedia.org/wiki/Ultrafilter?oldid=776513664 Contributors: AxelBoldt, Zundark, Michael Hardy, Chinju,
Stevan White, Charles Matthews, Timwi, Prumpf, MathMartin, Giftlite, Gene Ward Smith, Markus Krtzsch, Jason Quinn, Salasks, Paul
August, EmilJ, TenOfAllTrades, Oleg Alexandrov, Graham87, Rjwilmsi, R.e.b., Vlad Patryshev, FlaBot, Karelj, YurikBot, Benja, Trova-
tore, Crasshopper, Arthur Rubin, Psolrzan, Eskimbot, Mhss, Henning Makholm, Physis, JRSpriggs, CRGreathouse, Gregbard, Nick
Number, JAnDbot, .anacondabot, Magioladitis, EdwardLockhart, Sullivan.t.j, David Eppstein, Danimey, Quux0r, Gogobera, VolkovBot,
Cbigorgne, M gol, Paradoctor, Anchor Link Bot, Nsk92, Mpd1989, Hans Adler, Hugo Herbelin, Legobot, Luckas-bot, Yobot, Ht686rg90,
Kilom691, AnomieBOT, Xqbot, Howard McCay, FrescoBot, Theorist2, Citation bot 1, Tkuvho, RjwilmsiBot, EmausBot, Dewritech,
Wgunther, Bbbbbbbbba, Nosuchforever, CitationCleanerBot, Dexbot, Jochen Burghardt, Mark viking, Grabigail, Baking Soda, -oo- and
Anonymous: 28
11.13.2 Images
File:Commons-logo.svg Source: https://upload.wikimedia.org/wikipedia/en/4/4a/Commons-logo.svg License: PD Contributors: ? Orig-
inal artist: ?
File:Duale_Verbaende.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/e5/Duale_Verbaende.svg License: CC0 Con-
tributors: Own work Original artist: Mini-oh
File:Filter_vs_ultrafilter.svg Source: https://upload.wikimedia.org/wikipedia/commons/6/68/Filter_vs_ultrafilter.svg License: CC BY-
SA 4.0 Contributors: Own work, based on File:Upset.svg Original artist: Jochen Burghardt
File:Folder_Hexagonal_Icon.svg Source: https://upload.wikimedia.org/wikipedia/en/4/48/Folder_Hexagonal_Icon.svg License: Cc-
by-sa-3.0 Contributors: ? Original artist: ?
File:Hasse_diagram_of_powerset_of_3.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/ea/Hasse_diagram_of_powerset_
of_3.svg License: CC-BY-SA-3.0 Contributors: self-made using graphviz's dot. Original artist: KSmrq
File:Lattice_of_the_divisibility_of_60.svg Source: https://upload.wikimedia.org/wikipedia/commons/5/51/Lattice_of_the_divisibility_
of_60.svg License: CC-BY-SA-3.0 Contributors: No machine-readable source provided. Own work assumed (based on copyright claims).
Original artist: No machine-readable author provided. Ed g2s assumed (based on copyright claims).
File:Merge-split-transwiki_default.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/e0/Merge-split-transwiki_default.
svg License: Public domain Contributors: Self-made, based on <a href='//commons.wikimedia.org/wiki/File:Merge-split-transwiki_
default.gif' class='image'><img alt='Merge-split-transwiki default.gif' src='https://upload.wikimedia.org/wikipedia/commons/8/86/Merge-split-transwiki_
default.gif' width='30' height='37' data-le-width='30' data-le-height='37' /></a> by Father Goose. Arrows derived from <a href='//commons.
wikimedia.org/wiki/File:Merge-arrows.svg' class='image'><img alt='Merge-arrows.svg' src='https://upload.wikimedia.org/wikipedia/commons/
thumb/5/52/Merge-arrows.svg/50px-Merge-arrows.svg.png' width='50' height='20' srcset='https://upload.wikimedia.org/wikipedia/commons/
thumb/5/52/Merge-arrows.svg/75px-Merge-arrows.svg.png 1.5x, https://upload.wikimedia.org/wikipedia/commons/thumb/5/52/Merge-arrows.
svg/100px-Merge-arrows.svg.png 2x' data-le-width='50' data-le-height='20' /></a>. Original artist: Davidgothberg
File:Nuvola_apps_edu_mathematics_blue-p.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/3e/Nuvola_apps_edu_
mathematics_blue-p.svg License: GPL Contributors: Derivative work from Image:Nuvola apps edu mathematics.png and Image:Nuvola
apps edu mathematics-p.svg Original artist: David Vignoni (original icon); Flamurai (SVG convertion); bayo (color)
File:People_icon.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/37/People_icon.svg License: CC0 Contributors: Open-
Clipart Original artist: OpenClipart
File:Portal-puzzle.svg Source: https://upload.wikimedia.org/wikipedia/en/f/fd/Portal-puzzle.svg License: Public domain Contributors:
? Original artist: ?
File:T_30.svg Source: https://upload.wikimedia.org/wikipedia/commons/1/10/T_30.svg License: CC0 Contributors: Own work Original
artist: Mini-oh
File:Text_document_with_red_question_mark.svg Source: https://upload.wikimedia.org/wikipedia/commons/a/a4/Text_document_
with_red_question_mark.svg License: Public domain Contributors: Created by bdesham with Inkscape; based upon Text-x-generic.svg
from the Tango project. Original artist: Benjamin D. Esham (bdesham)
11.13. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 55