Вы находитесь на странице: 1из 142

Binary relations

en.wikipedia.org
Chapter 1

Accessibility relation

In modal logic, an accessibility relation R is a binary relation such that R W W where W is a set of possible
worlds. The accessibility relation determines for each world w W which worlds w are accessible from w . If a
possible world w is accessible from w we usually write wRw (or sometimes Rww ).

1.1 Description of terms

A statement in logic refers to a sentence (with a subject, predicate, and verb) that can be true or false. So, The
room is cold' is a statement because it contains a subject, predicate and verb, and it can be true that 'the room is cold'
or false that 'the room is cold.'
Generally, commands, beliefs and sentences about probabilities aren't judged as true or false. Inhale and exhale is
therefore not a statement in logic because it is a command and cannot be true or false, although a person can obey or
refuse that command. I believe I can y or I can't y isn't taken as a statement of truth or falsity, because beliefs
don't say anything about the truth or falsity of the parts of the entire and or or statement and therefore the entire
and or or statement.
A possible world is a possible situation. In every case, a possible world is contrasted with an actual situation. Earth
one minute from now is a possible world. Earth as it is actually is is also a possible world. Hence the oddity
of and controversy in contrasting a possible world with an actual world (Earth is necessarily possible). In logic,
worlds are described as a non-empty set, where the set could consist of anything, depending on what the statement
says.
Modal Logic is a description of the reasoning in making statements about possibility or necessity. It is possible that
it rains tomorrow is a statement in modal logic, because it is a statement about possibility. It is necessary that it
rains tomorrow also counts as a statement in modal logic, because it is a statement about necessity. There are at
least six logical axioms or principles that show what people mean whenever they make statements about necessity or
possibility (described below).[1]
As described in greater detail below:
Necessarily p means that p is true at every 'possible world' w such that R(w , w).
Possibly p means that p is true at some possible world w such that R(w , w) .
Truth-Value is whether a statement is true or false. Whether or not a statement is true, in turn, depends on the
meanings of words, laws of logic, or experience (observation, hearing, etc.).
Formal Semantics refers to the meaning of statements written in symbols. The sentence (p q) (p q) ,
for example, is a statement about 'necessity' in 'formal semantics.' It has a meaning that can be represented by the
symbol R .
The 'accessibility relation' is a relationship between two 'possible worlds.' More preciselyplease clarify denition , the
'accessibility relation' is the idea that modal statements, like 'its possible that it rains tomorrow,' may not take the
same truth-value in all 'possible worlds.' On earth, the statement could be true or false. By contrast, in a planet where
water is non-existent, this statement will always be false.

2
1.2. BASIC REVIEW OF (PROPOSITIONAL) MODAL LOGIC 3

Due to the diculty in judging if a modal statement is true in every 'possible world,' logicians have derived certain
axioms or principles that show on what basis any statement is true in any 'possible world.' These axioms describing
the relationship between 'possible worlds is the 'accessibility relation' in detail.
Put another way, these modal axioms describe in detail the 'accessibility relation,' R between two 'worlds.' That
relation, R symbolizes that from any given 'possible world' some other 'possible worlds may be accessible, and others
may not be.
The accessibility relation has important uses in both the formal/theoretical aspects of modal logic (theories about
'modal logic'). It also has applications to a variety of disciplines including epistemology (theories about how people
know something is true or false), metaphysics (theories about reality), value theory (theories about morality and
ethics), and computer science (theories about programmatic manipulation of data).

1.2 Basic Review of (Propositional) Modal Logic


The reasoning behind the 'accessibility relation' uses the basics of 'propositional modal logic' (see modal logic for a
detailed discussion). 'Propositional modal logic' is traditional propositional logic with the addition of two key unary
operators:
symbolizes the phrase 'It is necessary that...'
symbolizes the phrase 'It is possible that...'
These operators can be attached to a single sentence to form a new compound sentence.
For example, can be attached to a sentence such as 'I walk outside.' The new sentence would look like: 'I walk
outside.' The entire new sentence would therefore read: 'It is necessary that I walk outside.'
But the symbol A can be used to stand for any sentence instead of writing out entire sentences. So any sentence such
as 'I walk outside' or 'I walk outside and I look around' are symbolized by A .
Thus for any sentence A (simple or compound), the compound sentences A and A can be formed. Sentences
such as 'It is necessary that I walk outside' or 'It is possible that I walk outside' would therefore look like A and
A , respectively.
However, the symbols p , q can also be used to stand for any statement of our language. For example, p can stand
for 'I walk outside' or 'I walk outside and I look around.' The sentence 'It is necessary that I walk outside' would look
like: q . The sentence 'It is possible that I walk outside' would look like: q .
Six Basic Axioms of Modal Logic:
There are at least six basic axioms or principles of almost all modal logics or steps in thinking/reasoning. The rst
two hold in all regular modal logics, and the last holds in all normal modal logics.
1st Modal Axiom:

p p (Duality)

The double arrow stands symbolizes 'if and only if,' 'necessary and sucient' conditions. A 'necessary' condition is
something that must be the case for something else. Being literate, for instance, is a 'necessary' condition for reading
about the 'accessibility relation.' A 'sucient condition' a condition that is good enough for something else. Being
literate, for instance, is a 'sucient' condition for learning about the accessibility relation.' In other words, it is enough
to be literate in order to learn about the 'accessibility relation,' but may not be 'necessary' because the relation could be
learned in dierent ways (such as through speech). Aside from 'necessary and sucient,' the double arrow represents
equivalence between the meaning of two statements, the statement to the left and the statement to the right of the
double arrow.
The half square symbols before the diamond and p symbol in the sentence ' p p ' stand for 'it is not the
case, or 'not.'
The p symbol stands for any statement such as 'I walk outside.' Therefore it could also stand for 'The apple is Red.'
Example 1:
The rst principle says that any statement involving 'necessity' on the left side of the double arrow is equivalent to the
statement about the negation of 'possibility' on the right.
4 CHAPTER 1. ACCESSIBILITY RELATION

So using the symbols and their meaning, the rst modal axiom, p p could stand for: 'Its necessary that I
walk outside if and only if its not possible that it is not the case that I walk outside.'
And when I say that 'Its necessary that I walk outside,' this is the same as saying that 'Its not possible that it is not
the case that I walk outside.' Furthermore, when I say that 'Its not possible that it is not the case that I walk outside,'
this is the same as saying that 'Its necessary that I walk outside.'
Example 2:
p stands for 'The apple is red.'
So using the symbols and their meaning described above, the rst modal axiom, p p could stand for: 'Its
necessary that the apple is red if and only if its not possible that it is not the case that the apple is red.'
And when I say that 'Its necessary that the apple is red,' this is the same as saying that 'Its not possible that it is not
the case that the apple is red.' Furthermore, when I say that 'Its not possible that it is not the case that the apple is
red,' this is the same as saying that 'Its necessary that the apple is red.'
Second Modal Axiom:

p p (Duality)

Example 1:
The second principle says that any statement involving 'possibility' on the left side of the double arrow is the same as
the statement about the negation of 'necessity' on the right.
p stands for 'Spring has not arrived.'
Using the symbols and their meaning described above, the second modal axiom, p p could stand for: 'Its
possible that Spring has not arrived if and only if it is not the case that it is necessary that it is not the case that Spring
has not arrived.'
Essentially, the second axiom says that any statement about possibility called 'X' is the same as a negation or denial
of a dierent statement about necessity 'Y.' The statement about necessity shows the denial of the same original
statement 'X.'
The other axioms can be read and interpreted in the same way, by substituting letters p for any statement and following
the reasoning. Brackets in a symbolized sentence mean that anything inside the brackets counts as a whole sentence.
Any symbol before the brackets therefore applies to the sentence as a whole, not just the letters or an individual
sentence.
An arrow stands for then where the left statement before the arrow is the if of the entire sentence.
Other Modal Axioms:
* (p q) (p q)
* (p q) (p q)
* (p q) (p q) (Kripke property)
Most of the other axioms concerning the modal operators are controversial and not widely agreed upon. Here are the
most commonly used and discussed of these:

(T) p p

(4) p p

(5) p p

(B) p p

Here, "(T)","(4)","(5)", and "(B)" represent the traditional names of these axioms (or principles).
According to the traditional 'possible worlds semantics of modal logic, the compound sentences that are formed out
of the modal operators are to be interpreted in terms of quantication over possible worlds, subject to the relation
1.3. THE FOUR TYPES OF THE 'ACCESSIBILITY RELATION' IN FORMAL SEMANTICS 5

of accessibility. A sentence like (p q) (p q) is to be interpreted as true or false in all or some 'possible


worlds.' In turn, the grounds on which the sentence is true (symmetry, transitive property, etc.) in all 'possible worlds
is the 'accessibility relation.'
The relation of accessibility can now be dened as an (uninterpreted) relation R(w1 , w2 ) that holds between 'possible
worlds w1 and w2 only when w2 is accessible from w1 .
Additionally, to make things more specic, any formula, axiom like (p q) (p q) can be translated into a
formula of rst-order logic through standard translation. That rst-order logic formula or sentence makes the meaning
of the boxes and diamonds in modal logic explicit.

1.3 The Four Types of the 'Accessibility Relation' in Formal Semantics


'Formal semantics studies the meaning of statements written in symbols. The sentence (p q) (p q) ,
for example, is a statement about 'necessity' in 'formal semantics.' It has a meaning that can be represented by the
symbol R , where R takes the form of the 'necessity relation' described below.
So, the 'accessibility relation,' R can take on at least four forms, that is, the 'accessibility relation' can be described
in at least four ways.
Each type is either about 'possibility' or 'necessity' where 'possibility' and 'necessity' is dened as:

(TS) Necessarily p means that p is true at every 'possible world' w such that R(w , w) .
Possibly p means that p is true at some possible world w such that R(w , w) .

The four types of R will be a variation of these two general types. They will specify on what conditions a statement
is true either in every possible world, or some possible. The four specic types of R are:
Reexive, or *Axiom (T) above:
If R is reexive, every world is accessible to itself. Reexivity guarantees that any world at which A is true is a
world from which there is an accessible world at which A is true, and thus A is possible at worlds where its true,
which isn't necessarily the case in worlds that aren't accessible to themselves. Without the reexivity condition, A
can be necessary at a world where its false, if that world isn't accessible to itself; thus axiom Tthat A at a world
implies A is true at that worldfollows from reexivity.
Transitive, or *Axiom (4) above:
If R is transitive, any world accessible to any world w accessible to world w is also accessible to w . Transitively,
A is true at a world w only when A is true at every world w accessible to w , including every world w
accessible to any w , and every world accessible to any w , etc., so when A is true at w , its also true at every
w and every w , etc., which means A is also true at w , which is axiom 4.
Euclidean or *Axiom (5) above:
If R is euclidean, any two worlds accessible to a given world are accessible to each other. A is true at a world w
if and only if, for every world w accessible to w , there is a world w accessible to w at which A is true. If A
is true at a world w accessible to w , then if that world is accessible to every other world accessible to w , it will
be true that for every world accessible to w there is an accessible world ( w ) at which A is true, so A is true at
all worlds accessible to w . The euclidean property thus entails that A implies A , which is axiom 5.
Symmetric or *Axiom (B) above:
If R is symmetric, then if world w is accessible to world w , w is accessible to w . If A is true at w , then at
every w accessible to w , there is a world ( w ) accessible to w at which A is true, so A is possible at all w ,
and thus its necessary at w that A is possible, which is axiom B.

1.4 Philosophical Applications


One of the applications of 'possible worlds semantics and the 'accessibility relation' is to physics. Instead of just
talking generically about 'necessity (or logical necessity),' the relation in physics deals with 'nomological necessity.'
The fundamental translational schema (TS) described earlier can be exemplied as follows for physics:
6 CHAPTER 1. ACCESSIBILITY RELATION

(TSN) P is nomologically necessary means that P is true at all possible worlds that are nomologically accessible
from the actual world. In other words, P is true at all possible worlds that obey the physical laws of the actual
world.

The interesting thing to observe is that instead of having to ask, now, Does nomological necessity satisfy the axiom
(5)?", that is, Is something that is nomologically possible nomologically necessarily possible?", we can ask instead:
Is the nomological accessibility relation euclidean?" And dierent theories of the nature of physical laws will result
in dierent answers to this question. (Notice however that if the objection raised earlier is true, each dierent theory
of the nature of physical laws would be 'possible' and 'necessary,' since the euclidean concept depends on the idea
about 'possibility' and 'necessity'). The theory of Lewis, for example, is asymmetric. His counterpart theory also
requires an intransitive relation of accessibility because it is based on the notion of similarity and similarity is generally
intransitive. For example, a pile of straw with one less handful of straw may be similar to the whole pile but a pile
with two (or more) less handfuls may not be. So x can be necessarily P without x being necessarily necessarily P
. On the other hand, Saul Kripke has an account of de re modality which is based on (metaphysical) identity across
worlds and is therefore transitive.
Another interpretation of the 'accessibility relation' with a physical meaning was given in Gerla 1987 where the claim
is possible P in the world w is interpreted as it is possible to transform w into a world in which P is true. So,
the properties of the modal operators depend on the algebraic properties of the set of admissible transformations.
There are other applications of the 'accessibility relation' in philosophy. In epistemology, one can, instead of talking
about nomological accessibility, talk about epistemic accessibility. A world w is epistemically accessible from w
for an individual I in w if and only if I does not know something which would rule out the hypothesis that w = w
. We can ask whether the relation is transitive. If I knows nothing that rules out the possibility that w = w and
knows nothing that rules the possibility that w = w , it does not follow that I knows nothing which rules out the
hypothesis that w = w . To return to our earlier example, one may not be able to distinguish a pile of sand from the
same pile with one less handful and one may not be able to distinguish the pile with one less handful from the same
pile with two less handfuls of sand, but one may still be able to distinguish the original pile from the pile with two
less handfuls of sand.
Yet another example of the use of the 'accessibility relation' is in deontic logic. If we think of obligatoriness as truth
in all morally perfect worlds, and permissibility as truth in some morally perfect world, then we will have to restrict
out universe to include only morally perfect worlds. But, in that case, we will have left out the actual world. A better
alternative would be to include all the metaphysically possible worlds but restrict the 'accessibility relation' to morally
perfect worlds. Transitivity and the euclidean property will hold, but reexivity and symmetry will not.

1.5 Computer Science Applications


In modeling a computation, a 'possible world' can be a possible computer state. Given the current computer state, you
might dene the accessible possible worlds to be all future possible computer states, or to be all possible immediate
next computer states (assuming a discrete computer). Either choice denes a particular 'accessibility relation' giving
rise to a particular modal logic suited specically for theorems about the computation.

1.6 See also


Modal logic

Possible worlds

Propositional attitude

Modal depth

1.7 References
[1] For a detailed explanation on modal logic, see here.
1.7. REFERENCES 7

Gerla, G.; Transformational semantics for rst order logic, Logique et Analyse, No. 117118, pp. 6979,
1987.
Fitelson, Brandon; Notes on Accessibility and Modality, 2003.

Brown, Curtis; Propositional Modal Logic: A Few First Steps, 2002.


Kripke, Saul; Naming and Necessity, Oxford, 1980.

Lewis, David K.; Counterpart Theory and Quantied Modal Logic (subscription required) , The Journal of Philosophy,
Vol. LXV, No. 5 (1968-03-07), pp. 113126, 1968

List of Logic Systems List of most of the more popular modal logics.
Chapter 2

Ancestral relation

In mathematical logic, the ancestral relation (often shortened to ancestral) of a binary relation R is its transitive
closure, however dened in a dierent way, see below.
Ancestral relations make their rst appearance in Frege's Begrisschrift. Frege later employed them in his Grundge-
setze as part of his denition of the nite cardinals. Hence the ancestral was a key part of his search for a logicist
foundation of arithmetic.

2.1 Denition
The numbered propositions below are taken from his Begrisschrift and recast in contemporary notation.
A property P is called R-hereditary if, whenever x is P and xRy holds, then y is also P:

(P x xRy) P y
Frege dened b to be an R-ancestor of a, written aR* b, if b has every R-hereditary property that all objects x such
that aRx have:

76 : aR b F [x(aRx F x) xy(F x xRy F y) F b]


The ancestral is a transitive relation:

98 : (aR b bR c) aR c
Let the notation I(R) denote that R is functional (Frege calls such relations many-one):

115 : I(R) xyz[(xRy xRz) y = z]


If R is functional, then the ancestral of R is what nowadays is called connected:

133 : (I(R) aR b aR c) (bR c b = c cR b)

2.2 Relationship to transitive closure


The Ancestral relation R is equal to the transitive closure R+ of R . Indeed, R is transitive (see 98 above), R
contains R (indeed, if aRb then, of course, b has every R-hereditary property that all objects x such that aRx have,
because b is one of them), and nally, R is contained in R+ (indeed, assume aR b ; take the property F x to be
aR+ x ; then the two premises, x(aRx F x) and xy(F x xRy F y) , are obviously satised; therefore,
F b , which means aR+ b , by our choice of F ). See also Booloss book below, page 8.

8
2.3. DISCUSSION 9

2.3 Discussion
Principia Mathematica made repeated use of the ancestral, as does Quines (1951) Mathematical Logic.
However, it is worth noting that the ancestral relation cannot be dened in rst-order logic. It is controversial whether
second-order logic is really logic at all. Quine famously claimed that it was not, despite his reliance upon it for
his 1951 book (which largely retells Principia in abbreviated form, for which second-order logic is required to t its
theorems).

2.4 See also


Begrisschrift
Gottlob Frege

Transitive closure

2.5 References
George Boolos, 1998. Logic, Logic, and Logic. Harvard Univ. Press.
Ivor Grattan-Guinness, 2000. In Search of Mathematical Roots. Princeton Univ. Press.

Willard Van Orman Quine, 1951 (1940). Mathematical Logic. Harvard Univ. Press. ISBN 0-674-55451-5.

2.6 External links


Stanford Encyclopedia of Philosophy: "Freges Logic, Theorem, and Foundations for Arithmetic" -- by Edward
N. Zalta. Section 4.2.
Chapter 3

Antisymmetric relation

In mathematics, a binary relation R on a set X is anti-symmetric if there is no pair of distinct elements of X each of
which is related by R to the other. More formally, R is anti-symmetric precisely if for all a and b in X

if R(a,b) and R(b,a), then a = b,

or, equivalently,

if R(a,b) with a b, then R(b,a) must not hold.

As a simple example, the divisibility order on the natural numbers is an anti-symmetric relation. In this context,
anti-symmetry means that the only way each of two numbers can be divisible by the other is if the two are, in fact,
the same number; equivalently, if n and m are distinct and n is a factor of m, then m cannot be a factor of n.
In mathematical notation, this is:

a, b X, (aRb bRa) a = b

or, equivalently,

a, b X, (aRb a = b bRa).

The usual order relation on the real numbers is anti-symmetric: if for two real numbers x and y both inequalities
x y and y x hold then x and y must be equal. Similarly, the subset order on the subsets of any given set is
anti-symmetric: given two sets A and B, if every element in A also is in B and every element in B is also in A, then A
and B must contain all the same elements and therefore be equal:

ABB AA=B

Partial and total orders are anti-symmetric by denition. A relation can be both symmetric and anti-symmetric (e.g.,
the equality relation), and there are relations which are neither symmetric nor anti-symmetric (e.g., the preys on
relation on biological species).
Anti-symmetry is dierent from asymmetry, which requires both anti-symmetry and irreexivity.

3.1 Examples
The relation "x is even, y is odd between a pair (x, y) of integers is anti-symmetric:

10
3.2. SEE ALSO 11

Every asymmetric relation is also an anti-symmetric relation.

3.2 See also


Symmetric relation
Asymmetric relation

Symmetry in mathematics

3.3 References
Weisstein, Eric W. Antisymmetric Relation. MathWorld.

Lipschutz, Seymour; Marc Lars Lipson (1997). Theory and Problems of Discrete Mathematics. McGraw-Hill.
p. 33. ISBN 0-07-038045-7.
Chapter 4

Asymmetric relation

In mathematics, an asymmetric relation is a binary relation on a set X where:

For all a and b in X, if a is related to b, then b is not related to a.[1]

In mathematical notation, this is:

a, b X(aRb (bRa))

An example is the "less than" relation < between real numbers: if x < y, then necessarily y is not less than x.The
less than or equal relation , on the other hand, is not asymmetric, because reversing e.g. x x produces x x
and both are true. In general, any relation in which x R x holds for some x (that is, which is not irreexive) is also not
asymmetric.
Asymmetry is not the same thing as not symmetric": the less-than-or-equal relation is an example of a relation that
is neither symmetric nor asymmetric. The empty relation is the only relation that is (vacuously) both symmetric and
asymmetric.

4.1 Properties

A relation is asymmetric if and only if it is both antisymmetric and irreexive.[2]

Restrictions and inverses of asymmetric relations are also asymmetric. For example, the restriction of < from
the reals to the integers is still asymmetric, and the inverse > of < is also asymmetric.

A transitive relation is asymmetric if and only if it is irreexive:[3] if a R b and b R a, transitivity gives a R a,


contradicting irreexivity.

An asymmetric relation need not be total. For example, strict subset or is asymmetric, and neither of the
sets {1,2} and {3,4} is a strict subset of the other. In general, every strict partial order is asymmetric, and
conversely, every transitive asymmetric relation is a strict partial order. Not all asymmetric relations are strict
partial orders. An example of an asymmetric intransitive relation is the rock-paper-scissors relation: if X beats
Y, then Y does not beat X; but if X beats Y and Y beats Z, then X does not beat Z.

4.2 See also

Tarskis axiomatization of the reals part of this is the requirement that < over the real numbers be asymmetric.

12
4.3. REFERENCES 13

4.3 References
[1] Gries, David; Schneider, Fred B. (1993), A Logical Approach to Discrete Math, Springer-Verlag, p. 273.

[2] Nievergelt, Yves (2002), Foundations of Logic and Mathematics: Applications to Computer Science and Cryptography,
Springer-Verlag, p. 158.

[3] Flaka, V.; Jeek, J.; Kepka, T.; Kortelainen, J. (2007). Transitive Closures of Binary Relations I (PDF). Prague: School
of Mathematics - Physics Charles University. p. 1. Lemma 1.1 (iv). Note that this source refers to asymmetric relations
as strictly antisymmetric.
Chapter 5

Better-quasi-ordering

In order theory a better-quasi-ordering or bqo is a quasi-ordering that does not admit a certain type of bad array.
Every bqo is well-quasi-ordered.

5.1 Motivation

Though wqo is an appealing notion, many important innitary operations do not preserve wqoness. An example
due to Richard Rado illustrates this.[1] In a 1965 paper Crispin Nash-Williams formulated the stronger notion of
bqo in order to prove that the class of trees of height is wqo under the topological minor relation.[2] Since then,
many quasi-orders have been proven to be wqo by proving them to be bqo. For instance, Richard Laver established
Frass's conjecture by proving that the class of scattered linear order types is bqo.[3] More recently, Carlos Martinez-
Ranero has proven that, under the Proper Forcing Axiom, the class of Aronszajn lines is bqo under the embeddability
relation.[4]

5.2 Denition

It is common in bqo theory to write x for the sequence x with the rst term omitted. Write []< for the set of
nite, strictly increasing sequences with terms in , and dene a relation on []< as follows: s t if and only if
there is u such that s is a strict initial segment of u and t = u . Note that the relation is not transitive.

A block is an innite subset B of []< that contains an initial segment of every innite subset of B . For a quasi-
order Q a Q -pattern is a function from a block B into Q . A Q -pattern f : B Q is said to be bad if f (s) Q f (t)
for every pair s, t B such that s t ; otherwise f is good. A quasi-order Q is better-quasi-ordered (bqo) if there is
no bad Q -pattern.
In order to make this denition easier to work with, Nash-Williams denes a barrier to be a block whose elements
are pairwise incomparable under the inclusion relation . A Q -array is a Q -pattern whose domain is a barrier. By
observing that every block contains a barrier, one sees that Q is bqo if and only if there is no bad Q -array.

5.3 Simpsons alternative denition

Simpson introduced an alternative denition of bqo in terms of Borel maps [] Q , where [] , the set of innite
subsets of , is given the usual (product) topology.[5]
Let Q be a quasi-order and endow Q with the discrete topology. A Q -array is a Borel function [A] Q for
some innite subset A of . A Q -array f is bad if f (X) Q f ( X) for every X [A] ; f is good otherwise.
The quasi-order Q is bqo if there is no bad Q -array in this sense.

14
5.4. MAJOR THEOREMS 15

5.4 Major theorems


Many major results in bqo theory are consequences of the Minimal Bad Array Lemma, which appears in Simpsons
paper[5] as follows. See also Lavers paper,[6] where the Minimal Bad Array Lemma was rst stated as a result. The
technique was present in Nash-Williams original 1965 paper.
Suppose (Q, Q ) is a quasi-order. A partial ranking of Q is a well-founded partial ordering of Q such that
q r q Q r . For bad Q -arrays (in the sense of Simpson) f : [A] Q and g : [B] Q , dene:

g f if B A and g(X) f (X) every for X [B]

g < f if B A and g(X) < f (X) every for X [B]


We say a bad Q -array g is minimal bad (with respect to the partial ranking ) if there is no bad Q -array f such
that f < g . Note that the denitions of and < depend on a partial ranking of Q . Note also that the relation
< is not the strict part of the relation .
Theorem (Minimal Bad Array Lemma). Let Q be a quasi-order equipped with a partial ranking and suppose f is a
bad Q -array. Then there is a minimal bad Q -array g such that g f .

5.5 See also


Well-quasi-ordering
Well-order

5.6 References
[1] Rado, Richard (1954). Partial well-ordering of sets of vectors. Mathematika. 1 (2): 8995. MR 0066441. doi:10.1112/S0025579300000565.

[2] Nash-Williams, C. St. J. A. (1965). On well-quasi-ordering innite trees. Mathematical Proceedings of the Cambridge
Philosophical Society. 61 (3): 697720. Bibcode:1965PCPS...61..697N. ISSN 0305-0041. MR 0175814. doi:10.1017/S0305004100039062.

[3] Laver, Richard (1971). On Fraisses Order Type Conjecture. The Annals of Mathematics. 93 (1): 89111. JSTOR
1970754. doi:10.2307/1970754.

[4] Martinez-Ranero, Carlos (2011). Well-quasi-ordering Aronszajn lines. Fundamenta Mathematicae. 213 (3): 197211.
ISSN 0016-2736. MR 2822417. doi:10.4064/fm213-3-1.

[5] Simpson, Stephen G. (1985). BQO Theory and Frass's Conjecture. In Manseld, Richard; Weitkamp, Galen. Recursive
Aspects of Descriptive Set Theory. The Clarendon Press, Oxford University Press. pp. 12438. ISBN 978-0-19-503602-2.
MR 786122.

[6] Laver, Richard (1978). Better-quasi-orderings and a class of trees. In Rota, Gian-Carlo. Studies in foundations and
combinatorics. Academic Press. pp. 3148. ISBN 978-0-12-599101-8. MR 0520553.
Chapter 6

Binary relation

Relation (mathematics)" redirects here. For a more general notion of relation, see nitary relation. For a more
combinatorial viewpoint, see theory of relations. For other uses, see Relation (disambiguation).

In mathematics, a binary relation on a set A is a collection of ordered pairs of elements of A. In other words, it is a
subset of the Cartesian product A2 = A A. More generally, a binary relation between two sets A and B is a subset
of A B. The terms correspondence, dyadic relation and 2-place relation are synonyms for binary relation.
An example is the "divides" relation between the set of prime numbers P and the set of integers Z, in which every
prime p is associated with every integer z that is a multiple of p (but with no integer that is not a multiple of p). In
this relation, for instance, the prime 2 is associated with numbers that include 4, 0, 6, 10, but not 1 or 9; and the
prime 3 is associated with numbers that include 0, 6, and 9, but not 4 or 13.
Binary relations are used in many branches of mathematics to model concepts like "is greater than", "is equal to", and
divides in arithmetic, "is congruent to" in geometry, is adjacent to in graph theory, is orthogonal to in linear
algebra and many more. The concept of function is dened as a special kind of binary relation. Binary relations are
also heavily used in computer science.
A binary relation is the special case n = 2 of an n-ary relation R A1 An, that is, a set of n-tuples where the
jth component of each n-tuple is taken from the jth domain Aj of the relation. An example for a ternary relation on
ZZZ is " ... lies between ... and ..., containing e.g. the triples (5,2,8), (5,8,2), and (4,9,7).
In some systems of axiomatic set theory, relations are extended to classes, which are generalizations of sets. This
extension is needed for, among other things, modeling the concepts of is an element of or is a subset of in set
theory, without running into logical inconsistencies such as Russells paradox.

6.1 Formal denition

A binary relation R between arbitrary sets (or classes) X (the set of departure) and Y (the set of destination or
codomain) is specied by its graph G, which is a subset of the Cartesian product X Y. The binary relation R itself
is usually identied with its graph G, but some authors dene it as an ordered triple (X, Y, G), which is otherwise
referred to as a correspondence.[1]
The statement (x, y) G is read "x is R-related to y", and is denoted by xRy or R(x, y). The latter notation corresponds
to viewing R as the characteristic function of the subset G of X Y, i.e. R(x, y) equals to 1 (true), if (x, y) G, and
0 (false) otherwise.
The order of the elements in each pair of G is important: if a b, then aRb and bRa can be true or false, independently
of each other. Resuming the above example, the prime 3 divides the integer 9, but 9 doesn't divide 3.
The domain of R is the set of all x such that xRy for at least one y. The range of R is the set of all y such that xRy
for at least one x. The eld of R is the union of its domain and its range.[2][3][4]

16
6.2. SPECIAL TYPES OF BINARY RELATIONS 17

6.1.1 Is a relation more than its graph?


According to the denition above, two relations with identical graphs but dierent domains or dierent codomains
are considered dierent. For example, if G = {(1, 2), (1, 3), (2, 7)} , then (Z, Z, G) , (R, N, G) , and (N, R, G) are
three distinct relations, where Z is the set of integers, R is the set of real numbers and N is the set of natural numbers.
Especially in set theory, binary relations are often dened as sets of ordered pairs, identifying binary relations with
their graphs. The domain of a binary relation R is then dened as the set of all x such that there exists at least one
y such that (x, y) R , the range of R is dened as the set of all y such that there exists at least one x such that
(x, y) R , and the eld of R is the union of its domain and its range.[2][3][4]
A special case of this dierence in points of view applies to the notion of function. Many authors insist on distin-
guishing between a functions codomain and its range. Thus, a single rule, like mapping every real number x to
x2 , can lead to distinct functions f : R R and f : R R+ , depending on whether the images under that
rule are understood to be reals or, more restrictively, non-negative reals. But others view functions as simply sets of
ordered pairs with unique rst components. This dierence in perspectives does raise some nontrivial issues. As an
example, the former camp considers surjectivityor being ontoas a property of functions, while the latter sees it
as a relationship that functions may bear to sets.
Either approach is adequate for most uses, provided that one attends to the necessary changes in language, notation,
and the denitions of concepts like restrictions, composition, inverse relation, and so on. The choice between the two
denitions usually matters only in very formal contexts, like category theory.

6.1.2 Example
Example: Suppose there are four objects {ball, car, doll, gun} and four persons {John, Mary, Ian, Venus}. Suppose
that John owns the ball, Mary owns the doll, and Venus owns the car. Nobody owns the gun and Ian owns nothing.
Then the binary relation is owned by is given as

R = ({ball, car, doll, gun}, {John, Mary, Ian, Venus}, {(ball, John), (doll, Mary), (car, Venus)}).

Thus the rst element of R is the set of objects, the second is the set of persons, and the last element is a set of ordered
pairs of the form (object, owner).
The pair (ball, John), denoted by RJ means that the ball is owned by John.
Two dierent relations could have the same graph. For example: the relation

({ball, car, doll, gun}, {John, Mary, Venus}, {(ball, John), (doll, Mary), (car, Venus)})

is dierent from the previous one as everyone is an owner. But the graphs of the two relations are the same.
Nevertheless, R is usually identied or even dened as G(R) and an ordered pair (x, y) G(R)" is usually denoted as
"(x, y) R".[5]

6.2 Special types of binary relations


Some important types of binary relations R between two sets X and Y are listed below. To emphasize that X and Y
can be dierent sets, some authors call such binary relations heterogeneous.[6][7]
Uniqueness properties:

injective (also called left-unique[8] ): for all x and z in X and y in Y it holds that if xRy and zRy then x = z. For
example, the green relation in the diagram is injective, but the red relation is not, as it relates e.g. both x = 5
and z = +5 to y = 25.
functional (also called univalent[9] or right-unique[8] or right-denite[10] ): for all x in X, and y and z in Y
it holds that if xRy and xRz then y = z; such a binary relation is called a partial function. Both relations in
the picture are functional. An example for a non-functional relation can be obtained by rotating the red graph
clockwise by 90 degrees, i.e. by considering the relation x=y2 which relates e.g. x=25 to both y=5 and z=+5.
18 CHAPTER 6. BINARY RELATION

Example relations between real numbers. Red: y=x2 . Green: y=2x+20.

one-to-one (also written 1-to-1): injective and functional. The green relation is one-to-one, but the red is not.

Totality properties (only denable if the sets of departure X resp. destination Y are specied; not to be confused with
a total relation):

left-total:[8] for all x in X there exists a y in Y such that xRy. For example, R is left-total when it is a function
or a multivalued function. Note that this property, although sometimes also referred to as total, is dierent
from the denition of total in the next section. Both relations in the picture are left-total. The relation x=y2 ,
obtained from the above rotation, is not left-total, as it doesn't relate, e.g., x = 14 to any real number y.
surjective (also called right-total[8] or onto): for all y in Y there exists an x in X such that xRy. The green
relation is surjective, but the red relation is not, as it doesn't relate any real number x to e.g. y = 14.
6.3. RELATIONS OVER A SET 19

Uniqueness and totality properties:

A function: a relation that is functional and left-total. Both the green and the red relation are functions.

An injective function: a relation that is injective, functional, and left-total.

A surjective function or surjection: a relation that is functional, left-total, and right-total.

A bijection: a surjective one-to-one or surjective injective function is said to be bijective, also known as
one-to-one correspondence.[11] The green relation is bijective, but the red is not.

6.2.1 Difunctional
Less commonly encountered is the notion of difunctional (or regular) relation, dened as a relation R such that
R=RR1 R.[12]
To understand this notion better, it helps to consider a relation as mapping every element xX to a set xR = { yY
| xRy }.[12] This set is sometimes called the successor neighborhood of x in R; one can dene the predecessor
neighborhood analogously.[13] Synonymous terms for these notions are afterset and respectively foreset.[6]
A difunctional relation can then be equivalently characterized as a relation R such that wherever x1 R and x2 R have a
non-empty intersection, then these two sets coincide; formally x1 R x2 R implies x1 R = x2 R.[12]
As examples, any function or any functional (right-unique) relation is difunctional; the converse doesn't hold. If one
considers a relation R from set to itself (X = Y), then if R is both transitive and symmetric (i.e. a partial equivalence
relation), then it is also difunctional.[14] The converse of this latter statement also doesn't hold.
A characterization of difunctional relations, which also explains their name, is to consider two functions f: A C
and g: B C and then dene the following set which generalizes the kernel of a single function as joint kernel: ker(f,
g) = { (a, b) A B | f(a) = g(b) }. Every difunctional relation R A B arises as the joint kernel of two functions
f: A C and g: B C for some set C.[15]
In automata theory, the term rectangular relation has also been used to denote a difunctional relation. This ter-
minology is justied by the fact that when represented as a boolean matrix, the columns and rows of a difunctional
relation can be arranged in such a way as to present rectangular blocks of true on the (asymmetric) main diagonal.[16]
Other authors however use the term rectangular to denote any heterogeneous relation whatsoever.[7]

6.3 Relations over a set


If X = Y then we simply say that the binary relation is over X, or that it is an endorelation over X.[17] In computer
science, such a relation is also called a homogeneous (binary) relation.[7][17][18] Some types of endorelations are
widely studied in graph theory, where they are known as simple directed graphs permitting loops.
The set of all binary relations Rel(X) on a set X is the set 2X X which is a Boolean algebra augmented with the
involution of mapping of a relation to its inverse relation. For the theoretical explanation see Relation algebra.
Some important properties that a binary relation R over a set X may have are:

reexive: for all x in X it holds that xRx. For example, greater than or equal to () is a reexive relation but
greater than (>) is not.

irreexive (or strict): for all x in X it holds that not xRx. For example, > is an irreexive relation, but is not.

coreexive relation: for all x and y in X it holds that if xRy then x = y.[19] An example of a coreexive relation
is the relation on integers in which each odd number is related to itself and there are no other relations. The
equality relation is the only example of a both reexive and coreexive relation, and any coreexive relation is
a subset of the identity relation.

The previous 3 alternatives are far from being exhaustive; e.g. the red relation y=x2 from the
above picture is neither irreexive, nor coreexive, nor reexive, since it contains the pair
(0,0), and (2,4), but not (2,2), respectively.
20 CHAPTER 6. BINARY RELATION

symmetric: for all x and y in X it holds that if xRy then yRx. Is a blood relative of is a symmetric relation,
because x is a blood relative of y if and only if y is a blood relative of x.

antisymmetric: for all x and y in X, if xRy and yRx then x = y. For example, is anti-symmetric; so is >, but
vacuously (the condition in the denition is always false).[20]

asymmetric: for all x and y in X, if xRy then not yRx. A relation is asymmetric if and only if it is both
anti-symmetric and irreexive.[21] For example, > is asymmetric, but is not.

transitive: for all x, y and z in X it holds that if xRy and yRz then xRz. For example, is ancestor of is transitive,
while is parent of is not. A transitive relation is irreexive if and only if it is asymmetric.[22]

total: for all x and y in X it holds that xRy or yRx (or both). This denition for total is dierent from left total
in the previous section. For example, is a total relation.

trichotomous: for all x and y in X exactly one of xRy, yRx or x = y holds. For example, > is a trichotomous
relation, while the relation divides on natural numbers is not.[23]

Right Euclidean: for all x, y and z in X it holds that if xRy and xRz, then yRz.

Left Euclidean: for all x, y and z in X it holds that if yRx and zRx, then yRz.

Euclidean: A Euclidean relation is both left and right Euclidean. Equality is a Euclidean relation because if
x=y and x=z, then y=z.

serial: for all x in X, there exists y in X such that xRy. "Is greater than" is a serial relation on the integers. But
it is not a serial relation on the positive integers, because there is no y in the positive integers such that 1>y.[24]
However, "is less than" is a serial relation on the positive integers, the rational numbers and the real numbers.
Every reexive relation is serial: for a given x, choose y=x. A serial relation can be equivalently characterized
as every element having a non-empty successor neighborhood (see the previous section for the denition of this
notion). Similarly an inverse serial relation is a relation in which every element has non-empty predecessor
neighborhood.[13]

set-like (or local): for every x in X, the class of all y such that yRx is a set. (This makes sense only if relations
on proper classes are allowed.) The usual ordering < on the class of ordinal numbers is set-like, while its inverse
> is not.

A relation that is reexive, symmetric, and transitive is called an equivalence relation. A relation that is symmetric,
transitive, and serial is also reexive. A relation that is only symmetric and transitive (without necessarily being
reexive) is called a partial equivalence relation.
A relation that is reexive, antisymmetric, and transitive is called a partial order. A partial order that is total is called
a total order, simple order, linear order, or a chain.[25] A linear order where every nonempty subset has a least element
is called a well-order.

6.4 Operations on binary relations


If R, S are binary relations over X and Y, then each of the following is a binary relation over X and Y:

Union: R S X Y, dened as R S = { (x, y) | (x, y) R or (x, y) S }. For example, is the union of >
and =.

Intersection: R S X Y, dened as R S = { (x, y) | (x, y) R and (x, y) S }.

If R is a binary relation over X and Y, and S is a binary relation over Y and Z, then the following is a binary relation
over X and Z: (see main article composition of relations)
6.4. OPERATIONS ON BINARY RELATIONS 21

Composition: S R, also denoted R ; S (or R S), dened as S R = { (x, z) | there exists y Y, such that (x, y)
R and (y, z) S }. The order of R and S in the notation S R, used here agrees with the standard notational
order for composition of functions. For example, the composition is mother of is parent of yields is
maternal grandparent of, while the composition is parent of is mother of yields is grandmother of.

A relation R on sets X and Y is said to be contained in a relation S on X and Y if R is a subset of S, that is, if x R y
always implies x S y. In this case, if R and S disagree, R is also said to be smaller than S. For example, > is contained
in .
If R is a binary relation over X and Y, then the following is a binary relation over Y and X:

Inverse or converse: R 1 , dened as R 1 = { (y, x) | (x, y) R }. A binary relation over a set is equal to its
inverse if and only if it is symmetric. See also duality (order theory). For example, is less than (<) is the
inverse of is greater than (>).

If R is a binary relation over X, then each of the following is a binary relation over X:

Reexive closure: R = , dened as R = = { (x, x) | x X } R or the smallest reexive relation over X containing
R. This can be proven to be equal to the intersection of all reexive relations containing R.
Reexive reduction: R , dened as R
= R \ { (x, x) | x X } or the largest irreexive relation over X
contained in R.
Transitive closure: R + , dened as the smallest transitive relation over X containing R. This can be seen to be
equal to the intersection of all transitive relations containing R.
Reexive transitive closure: R *, dened as R * = (R + ) = , the smallest preorder containing R.
Reexive transitive symmetric closure: R , dened as the smallest equivalence relation over X containing
R.

6.4.1 Complement
If R is a binary relation over X and Y, then the following too:

The complement S is dened as x S y if not x R y. For example, on real numbers, is the complement of >.

The complement of the inverse is the inverse of the complement.


If X = Y, the complement has the following properties:

If a relation is symmetric, the complement is too.


The complement of a reexive relation is irreexive and vice versa.
The complement of a strict weak order is a total preorder and vice versa.

The complement of the inverse has these same properties.

6.4.2 Restriction
The restriction of a binary relation on a set X to a subset S is the set of all pairs (x, y) in the relation for which x and
y are in S.
If a relation is reexive, irreexive, symmetric, antisymmetric, asymmetric, transitive, total, trichotomous, a partial
order, total order, strict weak order, total preorder (weak order), or an equivalence relation, its restrictions are too.
However, the transitive closure of a restriction is a subset of the restriction of the transitive closure, i.e., in general
not equal. For example, restricting the relation "x is parent of y" to females yields the relation "x is mother of
the woman y"; its transitive closure doesn't relate a woman with her paternal grandmother. On the other hand, the
22 CHAPTER 6. BINARY RELATION

transitive closure of is parent of is is ancestor of"; its restriction to females does relate a woman with her paternal
grandmother.
Also, the various concepts of completeness (not to be confused with being total) do not carry over to restrictions.
For example, on the set of real numbers a property of the relation "" is that every non-empty subset S of R with an
upper bound in R has a least upper bound (also called supremum) in R. However, for a set of rational numbers this
supremum is not necessarily rational, so the same property does not hold on the restriction of the relation "" to the
set of rational numbers.
The left-restriction (right-restriction, respectively) of a binary relation between X and Y to a subset S of its domain
(codomain) is the set of all pairs (x, y) in the relation for which x (y) is an element of S.

6.4.3 Algebras, categories, and rewriting systems


Various operations on binary endorelations can be treated as giving rise to an algebraic structure, known as relation
algebra. It should not be confused with relational algebra which deals in nitary relations (and in practice also nite
and many-sorted).
For heterogenous binary relations, a category of relations arises.[7]
Despite their simplicity, binary relations are at the core of an abstract computation model known as an abstract
rewriting system.

6.5 Sets versus classes


Certain mathematical relations, such as equal to, member of, and subset of, cannot be understood to be binary
relations as dened above, because their domains and codomains cannot be taken to be sets in the usual systems of
axiomatic set theory. For example, if we try to model the general concept of equality as a binary relation =, we
must take the domain and codomain to be the class of all sets, which is not a set in the usual set theory.
In most mathematical contexts, references to the relations of equality, membership and subset are harmless because
they can be understood implicitly to be restricted to some set in the context. The usual work-around to this problem
is to select a large enough set A, that contains all the objects of interest, and work with the restriction =A instead of
=. Similarly, the subset of relation needs to be restricted to have domain and codomain P(A) (the power set of
a specic set A): the resulting set relation can be denoted A. Also, the member of relation needs to be restricted
to have domain A and codomain P(A) to obtain a binary relation A that is a set. Bertrand Russell has shown that
assuming to be dened on all sets leads to a contradiction in naive set theory.
Another solution to this problem is to use a set theory with proper classes, such as NBG or MorseKelley set theory,
and allow the domain and codomain (and so the graph) to be proper classes: in such a theory, equality, membership,
and subset are binary relations without special comment. (A minor modication needs to be made to the concept of
the ordered triple (X, Y, G), as normally a proper class cannot be a member of an ordered tuple; or of course one
can identify the function with its graph in this context.)[26] With this denition one can for instance dene a function
relation between every set and its power set.

6.6 The number of binary relations


2
The number of distinct binary relations on an n-element set is 2n (sequence A002416 in the OEIS):
Notes:

The number of irreexive relations is the same as that of reexive relations.


The number of strict partial orders (irreexive transitive relations) is the same as that of partial orders.
The number of strict weak orders is the same as that of total preorders.
The total orders are the partial orders that are also total preorders. The number of preorders that are neither
a partial order nor a total preorder is, therefore, the number of preorders, minus the number of partial orders,
minus the number of total preorders, plus the number of total orders: 0, 0, 0, 3, and 85, respectively.
6.7. EXAMPLES OF COMMON BINARY RELATIONS 23

the number of equivalence relations is the number of partitions, which is the Bell number.

The binary relations can be grouped into pairs (relation, complement), except that for n = 0 the relation is its own
complement. The non-symmetric ones can be grouped into quadruples (relation, complement, inverse, inverse com-
plement).

6.7 Examples of common binary relations


order relations, including strict orders:

greater than
greater than or equal to
less than
less than or equal to
divides (evenly)
is a subset of

equivalence relations:

equality
is parallel to (for ane spaces)
is in bijection with
isomorphy

dependency relation, a nite, symmetric, reexive relation.

independency relation, a symmetric, irreexive relation which is the complement of some dependency relation.

6.8 See also


Conuence (term rewriting)

Hasse diagram

Incidence structure

Logic of relatives

Order theory

Triadic relation

6.9 Notes
[1] Encyclopedic dictionary of Mathematics. MIT. 2000. pp. 13301331. ISBN 0-262-59020-4.

[2] Suppes, Patrick (1972) [originally published by D. van Nostrand Company in 1960]. Axiomatic Set Theory. Dover. ISBN
0-486-61630-4.

[3] Smullyan, Raymond M.; Fitting, Melvin (2010) [revised and corrected republication of the work originally published in
1996 by Oxford University Press, New York]. Set Theory and the Continuum Problem. Dover. ISBN 978-0-486-47484-7.

[4] Levy, Azriel (2002) [republication of the work published by Springer-Verlag, Berlin, Heidelberg and New York in 1979].
Basic Set Theory. Dover. ISBN 0-486-42079-5.

[5] Megill, Norman (5 August 1993). df-br (Metamath Proof Explorer)". Retrieved 18 November 2016.
24 CHAPTER 6. BINARY RELATION

[6] Christodoulos A. Floudas; Panos M. Pardalos (2008). Encyclopedia of Optimization (2nd ed.). Springer Science & Business
Media. pp. 299300. ISBN 978-0-387-74758-3.

[7] Michael Winter (2007). Goguen Categories: A Categorical Approach to L-fuzzy Relations. Springer. pp. xxi. ISBN
978-1-4020-6164-6.

[8] Kilp, Knauer and Mikhalev: p. 3. The same four denitions appear in the following:

Peter J. Pahl; Rudolf Damrath (2001). Mathematical Foundations of Computational Engineering: A Handbook.
Springer Science & Business Media. p. 506. ISBN 978-3-540-67995-0.
Eike Best (1996). Semantics of Sequential and Parallel Programs. Prentice Hall. pp. 1921. ISBN 978-0-13-
460643-9.
Robert-Christoph Riemann (1999). Modelling of Concurrent Systems: Structural and Semantical Methods in the High
Level Petri Net Calculus. Herbert Utz Verlag. pp. 2122. ISBN 978-3-89675-629-9.

[9] Gunther Schmidt, 2010. Relational Mathematics. Cambridge University Press, ISBN 978-0-521-76268-7, Chapt. 5

[10] Ms, Stephan (2007), Reasoning on Spatial Semantic Integrity Constraints, Spatial Information Theory: 8th International
Conference, COSIT 2007, Melbourne, Australia, September 1923, 2007, Proceedings, Lecture Notes in Computer Science,
4736, Springer, pp. 285302, doi:10.1007/978-3-540-74788-8_18

[11] Note that the use of correspondence here is narrower than as general synonym for binary relation.

[12] Chris Brink; Wolfram Kahl; Gunther Schmidt (1997). Relational Methods in Computer Science. Springer Science &
Business Media. p. 200. ISBN 978-3-211-82971-4.

[13] Yao, Y. (2004). Semantics of Fuzzy Sets in Rough Set Theory. Transactions on Rough Sets II. Lecture Notes in Computer
Science. 3135. p. 309. ISBN 978-3-540-23990-1. doi:10.1007/978-3-540-27778-1_15.

[14] William Craig (2006). Semigroups Underlying First-order Logic. American Mathematical Soc. p. 72. ISBN 978-0-8218-
6588-0.

[15] Gumm, H. P.; Zarrad, M. (2014). Coalgebraic Simulations and Congruences. Coalgebraic Methods in Computer Science.
Lecture Notes in Computer Science. 8446. p. 118. ISBN 978-3-662-44123-7. doi:10.1007/978-3-662-44124-4_7.

[16] Julius Richard Bchi (1989). Finite Automata, Their Algebras and Grammars: Towards a Theory of Formal Expressions.
Springer Science & Business Media. pp. 3537. ISBN 978-1-4613-8853-1.

[17] M. E. Mller (2012). Relational Knowledge Discovery. Cambridge University Press. p. 22. ISBN 978-0-521-19021-3.

[18] Peter J. Pahl; Rudolf Damrath (2001). Mathematical Foundations of Computational Engineering: A Handbook. Springer
Science & Business Media. p. 496. ISBN 978-3-540-67995-0.

[19] Fonseca de Oliveira, J. N., & Pereira Cunha Rodrigues, C. D. J. (2004). Transposing Relations: From Maybe Functions
to Hash Tables. In Mathematics of Program Construction (p. 337).

[20] Smith, Douglas; Eggen, Maurice; St. Andre, Richard (2006), A Transition to Advanced Mathematics (6th ed.), Brooks/Cole,
p. 160, ISBN 0-534-39900-2

[21] Nievergelt, Yves (2002), Foundations of Logic and Mathematics: Applications to Computer Science and Cryptography,
Springer-Verlag, p. 158.

[22] Flaka, V.; Jeek, J.; Kepka, T.; Kortelainen, J. (2007). Transitive Closures of Binary Relations I (PDF). Prague: School
of Mathematics Physics Charles University. p. 1. Lemma 1.1 (iv). This source refers to asymmetric relations as strictly
antisymmetric.

[23] Since neither 5 divides 3, nor 3 divides 5, nor 3=5.

[24] Yao, Y.Y.; Wong, S.K.M. (1995). Generalization of rough sets using relationships between attribute values (PDF).
Proceedings of the 2nd Annual Joint Conference on Information Sciences: 3033..

[25] Joseph G. Rosenstein, Linear orderings, Academic Press, 1982, ISBN 0-12-597680-1, p. 4

[26] Tarski, Alfred; Givant, Steven (1987). A formalization of set theory without variables. American Mathematical Society. p.
3. ISBN 0-8218-1041-3.
6.10. REFERENCES 25

6.10 References
M. Kilp, U. Knauer, A.V. Mikhalev, Monoids, Acts and Categories: with Applications to Wreath Products and
Graphs, De Gruyter Expositions in Mathematics vol. 29, Walter de Gruyter, 2000, ISBN 3-11-015248-7.
Gunther Schmidt, 2010. Relational Mathematics. Cambridge University Press, ISBN 978-0-521-76268-7.

6.11 External links


Hazewinkel, Michiel, ed. (2001) [1994], Binary relation, Encyclopedia of Mathematics, Springer Science+Business
Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Chapter 7

Comparability

See also: Comparison (mathematics)

In mathematics, any two elements x and y of a set P that is partially ordered by a binary relation are comparable
when either x y or y x. If it is not the case that x and y are comparable, then they are called incomparable.
A totally ordered set is exactly a partially ordered set in which every pair of elements is comparable.
It follows immediately from the denitions of comparability and incomparability that both relations are symmetric,
that is x is comparable to y if and only if y is comparable to x, and likewise for incomparability.

7.1 Notation
<
Comparability is denoted by the symbol = , and incomparability by the symbol .[1] Thus, for any pair of elements x
>
and y of a partially ordered set, exactly one of

<
x=y and
>

is true.

7.2 Comparability graphs

Main article: Comparability graph

The comparability graph of a partially ordered set P has as vertices the elements of P and has as edges precisely those
<
pairs {x, y} of elements for which x=y .[2]
>

7.3 Classication

When classifying mathematical objects (e.g., topological spaces), two criteria are said to be comparable when the
objects that obey one criterion constitute a subset of the objects that obey the other, which is to say when they are
comparable under the partial order . For example, the T1 and T2 criteria are comparable, while the T1 and sobriety
criteria are not.

26
7.4. SEE ALSO 27

7.4 See also


Strict weak ordering, a partial ordering in which incomparability is a transitive relation

7.5 References
PlanetMath: partial order. Retrieved 6 April 2010.

[1] Trotter, William T. (1992), Combinatorics and Partially Ordered Sets:Dimension Theory, Johns Hopkins Univ. Press, p. 3

[2] Gilmore, P. C.; Homan, A. J. (1964), A characterization of comparability graphs and of interval graphs, Canadian
Journal of Mathematics, 16: 539548, doi:10.4153/CJM-1964-055-5.
Chapter 8

Congruence relation

For the term as used in elementary geometry, see congruence (geometry).

In abstract algebra, a congruence relation (or simply congruence) is an equivalence relation on an algebraic structure
(such as a group, ring, or vector space) that is compatible with the structure.[1] Every congruence relation has a corre-
sponding quotient structure, whose elements are the equivalence classes (or congruence classes) for the relation.[2]

8.1 Basic example


This section is about the (mod n) notation. For the binary mod operation, see modulo operation.

The prototypical example of a congruence relation is congruence modulo n on the set of integers. For a given positive
integer n , two integers a and b are called congruent modulo n , written

ab (mod n)

if a b is divisible by n (or equivalently if a and b have the same remainder when divided by n ).
for example, 37 and 57 are congruent modulo 10 ,

37 57 (mod 10)

since 37 57 = 20 is a multiple of 10, or equivalently since both 37 and 57 have a remainder of 7 when divided
by 10 .
Congruence modulo n (for a xed n ) is compatible with both addition and multiplication on the integers. That is,
if

a1 a2 (mod n) and b1 b2 (mod n)

then

a1 + b1 a2 + b2 (mod n) and a1 b1 a2 b2 (mod n)

The corresponding addition and multiplication of equivalence classes is known as modular arithmetic. From the point
of view of abstract algebra, congruence modulo n is a congruence relation on the ring of integers, and arithmetic
modulo n occurs on the corresponding quotient ring.

28
8.2. DEFINITION 29

8.2 Denition
The denition of a congruence depends on the type of algebraic structure under consideration. Particular denitions of
congruence can be made for groups, rings, vector spaces, modules, semigroups, lattices, and so forth. The common
theme is that a congruence is an equivalence relation on an algebraic object that is compatible with the algebraic
structure, in the sense that the operations are well-dened on the equivalence classes.
For example, a group is an algebraic object consisting of a set together with a single binary operation, satisfying
certain axioms. If G is a group with operation , a congruence relation on G is an equivalence relation on the
elements of G satisfying

g1 g2 and h1 h2 = g1 h1 g2 h2

for all g1 , g2 , h1 , h2 G . For a congruence on a group, the equivalence class containing the identity element
is always a normal subgroup, and the other equivalence classes are the cosets of this subgroup. Together, these
equivalence classes are the elements of a quotient group.
When an algebraic structure includes more than one operation, congruence relations are required to be compatible
with each operation. For example, a ring possesses both addition and multiplication, and a congruence relation on a
ring must satisfy

r1 + s1 r2 + s2 and r1 s1 r2 s2

whenever r1 r2 and s1 s2 . For a congruence on a ring, the equivalence class containing 0 is always a two-sided
ideal, and the two operations on the set of equivalence classes dene the corresponding quotient ring.
The general notion of a congruence relation can be given a formal denition in the context of universal algebra, a
eld which studies ideas common to all algebraic structures. In this setting, a congruence relation is an equivalence
relation on an algebraic structure that satises

(a1 ,a2 , . . . ,an ) (a1 ,a2 , . . . ,an )

for every n-ary operation , and all elements a1 , . . . ,an ,a1 , . . . ,an for each i.

8.3 Relation with homomorphisms


If f : A B is a homomorphism between two algebraic structures (such as homomorphism of groups, or a linear
map between vector spaces), then the relation R dened by

a1 R a2 if and only if f (a1 ) = f (a2 )

is a congruence relation. By the rst isomorphism theorem, the image of A under f is a substructure of B isomorphic
to the quotient of A by this congruence.

8.4 Congruences of groups, and normal subgroups and ideals


In the particular case of groups, congruence relations can be described in elementary terms as follows: If G is a group
(with identity element e and operation *) and ~ is a binary relation on G, then ~ is a congruence whenever:

1. Given any element a of G, a ~ a (reexivity);


2. Given any elements a and b of G, if a ~ b, then b ~ a (symmetry);
3. Given any elements a, b, and c of G, if a ~ b and b ~ c, then a ~ c (transitivity);
30 CHAPTER 8. CONGRUENCE RELATION

4. Given any elements a, a' , b, and b' of G, if a ~ a' and b ~ b' , then a * b ~ a' * b' ;

5. Given any elements a and a' of G, if a ~ a' , then a1 ~ a' 1 (this can actually be proven from the other four,
so is strictly redundant).

Conditions 1, 2, and 3 say that ~ is an equivalence relation.


A congruence ~ is determined entirely by the set {a G : a ~ e} of those elements of G that are congruent to the
identity element, and this set is a normal subgroup. Specically, a ~ b if and only if b1 * a ~ e. So instead of talking
about congruences on groups, people usually speak in terms of normal subgroups of them; in fact, every congruence
corresponds uniquely to some normal subgroup of G.

8.4.1 Ideals of rings and the general case


A similar trick allows one to speak of kernels in ring theory as ideals instead of congruence relations, and in module
theory as submodules instead of congruence relations.
A more general situation where this trick is possible is with Omega-groups (in the general sense allowing operators
with multiple arity). But this cannot be done with, for example, monoids, so the study of congruence relations plays
a more central role in monoid theory.

8.5 Universal algebra


The idea is generalized in universal algebra: A congruence relation on an algebra A is a subset of the direct product
A A that is both an equivalence relation on A and a subalgebra of A A.
The kernel of a homomorphism is always a congruence. Indeed, every congruence arises as a kernel. For a given
congruence ~ on A, the set A/~ of equivalence classes can be given the structure of an algebra in a natural fashion,
the quotient algebra. The function that maps every element of A to its equivalence class is a homomorphism, and the
kernel of this homomorphism is ~.
The lattice Con(A) of all congruence relations on an algebra A is algebraic.

8.6 See also


Table of congruences

Linear congruence theorem


Congruence lattice problem

8.7 Notes
[1] Hungerford, Thomas W.. Algebra. Springer-Verlag, 1974, p. 27

[2] Hungerford, 1974, p. 26

8.8 References
Horn and Johnson, Matrix Analysis, Cambridge University Press, 1985. ISBN 0-521-38632-2. (Section 4.5
discusses congruency of matrices.)
Chapter 9

Covering relation

For other uses of Cover in mathematics, see Cover (mathematics).


In mathematics, especially order theory, the covering relation of a partially ordered set is the binary relation which

{x,y,z}

{x,y} {x,z} {y,z}

{x} {y} {z}

The Hasse diagram of the power set of three elements, partially ordered by inclusion.

holds between comparable elements that are immediate neighbours. The covering relation is commonly used to
graphically express the partial order by means of the Hasse diagram.

9.1 Denition
Let X be a set with a partial order . As usual, let < be the relation on X such that x < y if and only if x y and
x = y .

31
32 CHAPTER 9. COVERING RELATION

Let x and y be elements of X .


Then y covers x , written x y , if x < y and there is no element z such that x < z < y . Equivalently, y covers x
if the interval [x, y] is the two-element set {x, y} .
When x y , it is said that y is a cover of x . Some authors also use the term cover to denote any such pair (x, y) in
the covering relation.

9.2 Examples
In a nite linearly ordered set {1, 2, ..., n}, i + 1 covers i for all i between 1 and n 1 (and there are no other
covering relations).
In the Boolean algebra of the power set of a set S, a subset B of S covers a subset A of S if and only if B is
obtained from A by adding one element not in A.
In Youngs lattice, formed by the partitions of all nonnegative integers, a partition covers a partition if and
only if the Young diagram of is obtained from the Young diagram of by adding an extra cell.
The Hasse diagram depicting the covering relation of a Tamari lattice is the skeleton of an associahedron.

The covering relation of any nite distributive lattice forms a median graph.
On the real numbers with the usual total order , the cover set is empty: no number covers another.

9.3 Properties
If a partially ordered set is nite, its covering relation is the transitive reduction of the partial order relation.
Such partially ordered sets are therefore completely described by their Hasse diagrams. On the other hand, in
a dense order, such as the rational numbers with the standard order, no element covers another.

9.4 References
Knuth, Donald E. (2006), The Art of Computer Programming, Volume 4, Fascicle 4, Addison-Wesley, ISBN
0-321-33570-8.

Stanley, Richard P. (1997), Enumerative Combinatorics, 1 (2nd ed.), Cambridge University Press, ISBN 0-
521-55309-1.

Davey, B.A. (2002), Introduction to Lattices and Order, 1 (2nd ed.), Cambridge University Press, ISBN 978-
0-521-78451-1.
Chapter 10

Dense order

In mathematics, a partial order or total order < on a set X is said to be dense if, for all x and y in X for which x < y,
there is a z in X such that x < z < y.

10.1 Example
The rational numbers with the ordinary ordering are a densely ordered set in this sense, as are the real numbers. On
the other hand, the ordinary ordering on the integers is not dense.

10.2 Uniqueness
Georg Cantor proved that every two densely totally ordered countable sets without lower or upper bounds are order-
isomorphic.[1] In particular, there exists an order-isomorphism between the rational numbers and other densely or-
dered countable sets including the dyadic rationals and the algebraic numbers. The proof of this result uses the
back-and-forth method.[2]
Minkowskis question mark function can be used to determine the order isomorphisms between the quadratic algebraic
numbers and the rational numbers, and between the rationals and the dyadic rationals.

10.3 Generalizations
Any binary relation R is said to be dense if, for all R-related x and y, there is a z such that x and z and also z and y are
R-related. Formally:

x y xRy (z xRz zRy).

Every reexive relation is dense. A strict partial order < is a dense order i < is a dense relation. A dense relation
that is also transitive is said to be idempotent.

10.4 See also


Dense set

Dense-in-itself

Kripke semantics

33
34 CHAPTER 10. DENSE ORDER

10.5 References
[1] Roitman, Judith (1990), Theorem 27, p. 123, Introduction to Modern Set Theory, Pure and Applied Mathematics, 8,
John Wiley & Sons, ISBN 9780471635192.

[2] Dasgupta, Abhijit (2013), Set Theory: With an Introduction to Real Point Sets, Springer-Verlag, p. 161, ISBN 9781461488545.

10.6 Additional reading


David Harel, Dexter Kozen, Jerzy Tiuryn, Dynamic logic, MIT Press, 2000, ISBN 0-262-08289-6, p. 6
Chapter 11

Dependence relation

Not to be confused with Dependency relation, which is a binary relation that is symmetric and reexive.

In mathematics, a dependence relation is a binary relation which generalizes the relation of linear dependence.
Let X be a set. A (binary) relation between an element a of X and a subset S of X is called a dependence relation,
written a S , if it satises the following properties:

if a S , then a S ;
if a S , then there is a nite subset S0 of S , such that a S0 ;

if T is a subset of X such that b S implies b T , then a S implies a T ;


if a S but a S {b} for some b S , then b (S {b}) {a} .

Given a dependence relation on X , a subset S of X is said to be independent if a S {a} for all a S. If S T


, then S is said to span T if t S for every t T. S is said to be a basis of X if S is independent and S spans X.
Remark. If X is a non-empty set with a dependence relation , then X always has a basis with respect to .
Furthermore, any two bases of X have the same cardinality.

11.1 Examples
Let V be a vector space over a eld F. The relation , dened by S if is in the subspace spanned by S ,
is a dependence relation. This is equivalent to the denition of linear dependence.

Let K be a eld extension of F. Dene by S if is algebraic over F (S). Then is a dependence relation.
This is equivalent to the denition of algebraic dependence.

11.2 See also


matroid

This article incorporates material from Dependence relation on PlanetMath, which is licensed under the Creative Com-
mons Attribution/Share-Alike License.

35
Chapter 12

Dependency relation

For other uses, see Dependency (disambiguation).


Not to be confused with Dependence relation, which is a generalization of the concept of linear dependence among
members of a vector space.

In mathematics and computer science, a dependency relation is a binary relation that is nite, symmetric, and
reexive; i.e. a nite tolerance relation. That is, it is a nite set of ordered pairs D , such that

If (a, b) D then (b, a) D (symmetric)


If a is an element of the set on which the relation is dened, then (a, a) D (reexive)

In general, dependency relations are not transitive; thus, they generalize the notion of an equivalence relation by
discarding transitivity.
Let denote the alphabet of all the letters of D . Then the independency induced by D is the binary relation I

I =\D
That is, the independency is the set of all ordered pairs that are not in D . The independency is symmetric and
irreexive.
The pairs (, D) and (, I) , or the triple (, D, I) (with I induced by D ) are sometimes called the concurrent
alphabet or the reliance alphabet.
The pairs of letters in an independency relation induce an equivalence relation on the free monoid of all possible
strings of nite length. The elements of the equivalence classes induced by the independency are called traces, and
are studied in trace theory.

12.1 Examples
Consider the alphabet = {a, b, c} . A possible dependency relation is

D = {a, b} {a, b} {a, c} {a, c}


= {a, b} {a, c}
2 2

= {(a, b), (b, a), (a, c), (c, a), (a, a), (b, b), (c, c)}
The corresponding independency is

ID = {(b, c) , (c, b)}


Therefore, the letters b, c commute, or are independent of one another.

36
12.1. EXAMPLES 37

A
a
c

b
Chapter 13

Directed set

In mathematics, a directed set (or a directed preorder or a ltered set) is a nonempty set A together with a reexive
and transitive binary relation (that is, a preorder), with the additional property that every pair of elements has an
upper bound.[1] In other words, for any a and b in A there must exist c in A with a c and b c.
The notion dened above is sometimes called an upward directed set. A downward directed set is dened
analogously,[2] meaning when every pair of elements is bounded below.[3] Some authors (and this article) assume
that a directed set is directed upward, unless otherwise stated. Beware that other authors call a set directed if and
only if it is directed both upward and downward.[4]
Directed sets are a generalization of nonempty totally ordered sets. That is, all totally ordered sets are directed sets
(contrast partially ordered sets, which need not be directed). Join semilattices (which are partially ordered sets) are
directed sets as well, but not conversely. Likewise, lattices are directed sets both upward and downward.
In topology, directed sets are used to dene nets, which generalize sequences and unite the various notions of limit
used in analysis. Directed sets also give rise to direct limits in abstract algebra and (more generally) category theory.

13.1 Equivalent denition


In addition to the denition above, there is an equivalent denition. A directed set is a set A with a preorder such
that every nite subset of A has an upper bound. In this denition, the existence of an upper bound of the empty
subset implies that A is nonempty.

13.2 Examples
Examples of directed sets include:

The set of natural numbers N with the ordinary order is a directed set (and so is every totally ordered set).

Let D1 and D2 be directed sets. Then the Cartesian product set D1 D2 can be made into a directed set by
dening (n1 , n2 ) (m1 , m2 ) if and only if n1 m1 and n2 m2 . In analogy to the product order this is the
product direction on the Cartesian product.

It follows from previous example that the set N N of pairs of natural numbers can be made into a directed
set by dening (n0 , n1 ) (m0 , m1 ) if and only if n0 m0 and n1 m1 .

If x0 is a real number, we can turn the set R {x0 } into a directed set by writing a b if and only if
|a x0 | |b x0 |. We then say that the reals have been directed towards x0 . This is an example of a directed
set that is not ordered (neither totally nor partially).

A (trivial) example of a partially ordered set that is not directed is the set {a, b}, in which the only order
relations are a a and b b. A less trivial example is like the previous example of the reals directed towards
x0 " but in which the ordering rule only applies to pairs of elements on the same side of x0 .

38
13.3. CONTRAST WITH SEMILATTICES 39

If T is a topological space and x0 is a point in T, we turn the set of all neighbourhoods of x0 into a directed set
by writing U V if and only if U contains V.

For every U: U U; since U contains itself.


For every U,V,W: if U V and V W, then U W; since if U contains V and V contains W then U
contains W.
For every U, V: there exists the set U V such that U U V and V U V; since both U and V
contain U V.

In a poset P, every lower closure of an element, i.e. every subset of the form {a| a in P, a x} where x is a
xed element from P, is directed.

13.3 Contrast with semilattices

Witness

Directed sets are a more general concept than (join) semilattices: every join semilattice is a directed set, as the join
or least upper bound of two elements is the desired c. The converse does not hold however, witness the directed set
40 CHAPTER 13. DIRECTED SET

{1000,0001,1101,1011,1111} ordered bitwise (e.g. 1000 1011 holds, but 0001 1000 does not, since in the last
bit 1 > 0), where {1000,0001} has three upper bounds but no least upper bound, cf. picture. (Also note that without
1111, the set is not directed.)

13.4 Directed subsets


The order relation in a directed set is not required to be antisymmetric, and therefore directed sets are not always
partial orders. However, the term directed set is also used frequently in the context of posets. In this setting, a subset
A of a partially ordered set (P,) is called a directed subset if it is a directed set according to the same partial order:
in other words, it is not the empty set, and every pair of elements has an upper bound. Here the order relation on the
elements of A is inherited from P; for this reason, reexivity and transitivity need not be required explicitly.
A directed subset of a poset is not required to be downward closed; a subset of a poset is directed if and only if its
downward closure is an ideal. While the denition of a directed set is for an upward-directed set (every pair of
elements has an upper bound), it is also possible to dene a downward-directed set in which every pair of elements
has a common lower bound. A subset of a poset is downward-directed if and only if its upper closure is a lter.
Directed subsets are used in domain theory, which studies directed complete partial orders.[5] These are posets in
which every upward-directed set is required to have a least upper bound. In this context, directed subsets again
provide a generalization of convergent sequences.

13.5 See also


Filtered category

Centered set
Linked set

13.6 Notes
[1] Kelley, p. 65.

[2] Robert S. Borden (1988). A Course in Advanced Calculus. Courier Corporation. p. 20. ISBN 978-0-486-15038-3.

[3] Arlen Brown; Carl Pearcy (1995). An Introduction to Analysis. Springer. p. 13. ISBN 978-1-4612-0787-0.

[4] Siegfried Carl; Seppo Heikkil (2010). Fixed Point Theory in Ordered Sets and Applications: From Dierential and Integral
Equations to Game Theory. Springer. p. 77. ISBN 978-1-4419-7585-0.

[5] Gierz, p. 2.

13.7 References
J. L. Kelley (1955), General Topology.
Gierz, Hofmann, Keimel, et al. (2003), Continuous Lattices and Domains, Cambridge University Press. ISBN
0-521-80338-1.
Chapter 14

Equality (mathematics)

In mathematics, equality is a relationship between two quantities or, more generally two mathematical expressions,
asserting that the quantities have the same value, or that the expressions represent the same mathematical object. The
equality between A and B is written A = B, and pronounced A equals B. The symbol "=" is called an "equals sign".
Thus there are three kinds of equality, which are formalized in dierent ways.

Two symbols refer to the same object.[1]

Two sets have the same elements.[2]

Two expressions evaluate to the same value, such as a number, vector, function or set.

These may be thought of as the logical, set-theoretic and algebraic concepts of equality respectively.

14.1 Etymology
The etymology of the word is from the Latin aequlis (equal, like, comparable, similar) from aequus (equal,
level, fair, just).

14.2 Equality in mathematical logic

14.2.1 Logical formulations


See also: First-order logic Equality and its axioms

Leibniz characterized the notion of equality as follows:

Given any x and y, x = y if and only if, given any predicate P, P(x) if and only if P(y).

In this law, "P(x) if and only if P(y)" can be weakened to "P(x) if P(y)"; the modied law is equivalent to the original.
Instead of considering Leibnizs law as a true statement that can be proven from the usual laws of logic (including
axioms about equality such as symmetry, reexivity and substitution), it can also be taken as the denition of equality.
The property of being an equivalence relation, as well as the properties given below, can then be proved: they become
theorems.

14.2.2 Some basic logical properties of equality


The substitution property states:

41
42 CHAPTER 14. EQUALITY (MATHEMATICS)

For any quantities a and b and any expression F(x), if a = b, then F(a) = F(b) (if both sides make sense, i.e.
are well-formed).

In rst-order logic, this is a schema, since we can't quantify over expressions like F (which would be a functional
predicate).
Some specic examples of this are:

For any real numbers a, b, and c, if a = b, then a + c = b + c (here F(x) is x + c);

For any real numbers a, b, and c, if a = b, then a c = b c (here F(x) is x c);

For any real numbers a, b, and c, if a = b, then ac = bc (here F(x) is xc);

For any real numbers a, b, and c, if a = b and c is not zero, then a/c = b/c (here F(x) is x/c).

The reexive property states:

For any quantity a, a = a.

This property is generally used in mathematical proofs as an intermediate step.


The symmetric property states:

For any quantities a and b, if a = b, then b = a.

The transitive property states:

For any quantities a, b, and c, if a = b and b = c, then a = c.

These three properties were originally included among the Peano axioms for natural numbers. Although the sym-
metric and transitive properties are often seen as fundamental, they can be proved if the substitution and reexive
properties are assumed instead.

14.2.3 Equalities as predicates


When A and B are not fully specied or depend on some variables, equality is a proposition, which may be true
for some values and false for some other values. Equality is a binary relation, or, in other words, a two-arguments
predicate, which may produce a truth value (false or true) from its arguments. In computer programming, its com-
putation from two expressions is known as comparison.

14.3 Equality in set theory


Main article: Axiom of extensionality

Equality of sets is axiomatized in set theory in two dierent ways, depending on whether the axioms are based on a
rst-order language with or without equality.

14.3.1 Set equality based on rst-order logic with equality


In rst-order logic with equality, the axiom of extensionality states that two sets which contain the same elements are
the same set.[3]

Logic axiom: x = y z, (z x z y)
14.4. EQUALITY IN ALGEBRA AND ANALYSIS 43

Logic axiom: x = y z, (x z y z)
Set theory axiom: (z, (z x z y)) x = y

Incorporating half of the work into the rst-order logic may be regarded as a mere matter of convenience, as noted
by Lvy.

The reason why we take up rst-order predicate calculus with equality is a matter of convenience; by
this we save the labor of dening equality and proving all its properties; this burden is now assumed by
the logic.[4]

14.3.2 Set equality based on rst-order logic without equality


In rst-order logic without equality, two sets are dened to be equal if they contain the same elements. Then the
axiom of extensionality states that two equal sets are contained in the same sets.[5]

Set theory denition: "x = y" means z, (z x z y)


Set theory axiom: x = y z, (x z y z)

14.4 Equality in algebra and analysis

14.4.1 Identities
Main article: Identity (mathematics)

When A and B may be viewed as functions of some variables, then A = B means that A and B dene the same function.
Such an equality of functions is sometimes called an identity. An example is (x + 1)2 = x2 + 2x + 1. Sometimes, but
not always, an identity is written with a triple bar: (x + 1)2 x2 + 2x + 1.

14.4.2 Equations
An equation is the problem of nding values of some variables, called unknowns, for which the specied equality is
true. Equation may also refer to an equality relation that is satised only for the values of the variables that one is
interested in. For example, x2 + y2 = 1 is the equation of the unit circle.
There is no standard notation that distinguishes an equation from an identity or other use of the equality relation: a
reader has to guess an appropriate interpretation from the semantics of expressions and the context. An identity is
asserted to be true for all values of variables in a given domain. An equation may sometimes mean an identity, but
more often it species a subset of the variable space to be the subset where the equation is true.

14.4.3 Congruences
Main articles: Congruence relation and Congruence (geometry)

In some cases, one may consider as equal two mathematical objects that are only equivalent for the properties that
are considered. This is, in particular the case in geometry, where two geometric shapes are said equal when one may
be moved to coincide with the other. The word congruence is also used for this kind of equality.

14.4.4 Approximate equality


There are some logic systems that do not have any notion of equality. This reects the undecidability of the equality
of two real numbers dened by formulas involving the integers, the basic arithmetic operations, the logarithm and the
exponential function. In other words, there cannot exist any algorithm for deciding such an equality.
44 CHAPTER 14. EQUALITY (MATHEMATICS)

The binary relation "is approximately equal" between real numbers or other things, even if more precisely dened,
is not transitive (it may seem so at rst sight, but many small dierences can add up to something big). However,
equality almost everywhere is transitive.

14.5 Relation with equivalence and isomorphism


Main articles: Equivalence relation and Isomorphism

Viewed as a relation, equality is the archetype of the more general concept of an equivalence relation on a set:
those binary relations that are reexive, symmetric, and transitive. The identity relation is an equivalence relation.
Conversely, let R be an equivalence relation, and let us denote by xR the equivalence class of x, consisting of all
elements z such that x R z. Then the relation x R y is equivalent with the equality xR = yR . It follows that equality is
the nest equivalence relation on any set S, in the sense that it is the relation that has the smallest equivalence classes
(every class is reduced to a single element).
In some contexts, equality is sharply distinguished from equivalence or isomorphism.[6] For example, one may distin-
guish fractions from rational numbers, the latter being equivalence classes of fractions: the fractions 1/2 and 2/4 are
distinct as fractions, as dierent strings of symbols, but they represent the same rational number, the same point
on a number line. This distinction gives rise to the notion of a quotient set.
Similarly, the sets

{A, B, C} and {1, 2, 3}

are not equal sets the rst consists of letters, while the second consists of numbers but they are both sets of three
elements, and thus isomorphic, meaning that there is a bijection between them, for example

A 7 1, B 7 2, C 7 3.

However, there are other choices of isomorphism, such as

A 7 3, B 7 2, C 7 1,

and these sets cannot be identied without making such a choice any statement that identies them depends
on choice of identication. This distinction, between equality and isomorphism, is of fundamental importance in
category theory, and is one motivation for the development of category theory.

14.6 See also


Equals sign

Inequality

Logical equality

Extensionality

14.7 Notes
[1] Rosser 2008, p. 163.

[2] Lvy 2002, pp. 13, 358. Mac Lane & Birkho 1999, p. 2. Mendelson 1964, p. 5.

[3] Kleene 2002, p. 189. Lvy 2002, p. 13. Shoeneld 2001, p. 239.
14.8. REFERENCES 45

[4] Lvy 2002, p. 4.

[5] Mendelson 1964, pp. 159161. Rosser 2008, pp. 211213

[6] (Mazur 2007)

14.8 References
Kleene, Stephen Cole (2002) [1967]. Mathematical Logic. Mineola, New York: Dover Publications. ISBN
978-0-486-42533-7.

Lvy, Azriel (2002) [1979]. Basic set theory. Mineola, New York: Dover Publications. ISBN 978-0-486-
42079-0.

Mac Lane, Saunders; Birkho, Garrett (1999) [1967]. Algebra (Third ed.). Providence, Rhode Island: Amer-
ican Mathematical Society.

Mazur, Barry (12 June 2007), When is one thing equal to some other thing? (PDF)
Mendelson, Elliott (1964). Introduction to Mathematical Logic. New York: Van Nostrand Reinhold.

Rosser, John Barkley (2008) [1953]. Logic for mathematicians. Mineola, New York: Dover Publication. ISBN
978-0-486-46898-3.

Shoeneld, Joseph Robert (2001) [1967]. Mathematical Logic (2nd ed.). A K Peters. ISBN 978-1-56881-
135-2.

14.9 External links


Hazewinkel, Michiel, ed. (2001) [1994], Equality axioms, Encyclopedia of Mathematics, Springer Sci-
ence+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Chapter 15

Equipollence (geometry)

In Euclidean geometry, equipollence is a binary relation between directed line segments. A line segment AB from
point A to point B has the opposite direction to line segment BA. Two directed line segments are equipollent when
they have the same length and direction.
The concept of equipollent line segments was advanced by Giusto Bellavitis in 1835. Subsequently the term vector
was adopted for a class of equipollent line segments. Bellavitiss use of the idea of a relation to compare dierent
but similar objects has become a common mathematical technique, particularly in the use of equivalence relations.
Bellavitis used a special notation for the equipollence of segments AB and CD:

AB CD.

The following passages, translated by Michael J. Crowe, show the anticipation that Bellavitis had of vector concepts:

Equipollences continue to hold when one substitutes for the lines in them, other lines which are respec-
tively equipollent to them, however they may be situated in space. From this it can be understood how
any number and any kind of lines may be summed, and that in whatever order these lines are taken, the
same equipollent-sum will be obtained...

In equipollences, just as in equations, a line may be transferred from one side to the other, provided that
the sign is changed...

Thus oppositely directed segments are negatives of each other: AB + BA 0.

The equipollence AB n.CD, where n stands for a positive number, indicates that AB is both parallel
to and has the same direction as CD, and that their lengths have the relation expressed by AB = n.CD.[1]

On a great circle of a sphere, two directed circular arcs are equipollent when they agree in direction and arc length.
An equivalence class of such arcs is associated with a quaternion versor

exp(ar) = cos a + r sin a, where a is arc length and r determines the plane of the great circle by
perpendicularity.

15.1 References

[1] Michael J. Crowe (1967) A History of Vector Analysis, Giusto Bellavitis and His Calculus of Equipollences, pp 524,
University of Notre Dame Press

46
15.2. FURTHER READING 47

15.2 Further reading


Giusto Bellavitis (1835) Saggio di applicazioni di un nuovo metodo di Geometria Analitica (Calcolo delle
equipollenze)", Annali delle Scienze del Regno Lombardo-Veneto, Padova 5: 24459.
Giusto Bellavitis (1854) Sposizione del Metodo della Equipollenze, link from Google Books.

Giusto Bellavitis (1858) Calcolo dei Quaternioni di W.R. Hamilton e sua Relazione col Metodo delle Equipol-
lenze, link from HathiTrust.
Charles-Ange Laisant (1887) Theorie et Applications des Equipollence, Gauthier-Villars, link from University
of Michigan Historical Math Collection.
Lena L. Severance (1930) The Theory of Equipollences; Method of Analytical Geometry of Sig. Bellavitis,
link from HathiTrust.

15.3 External links


Axiomatic denition of equipollence
Chapter 16

Equivalence class

This article is about equivalency in mathematics. For equivalency in music, see equivalence class (music).
In mathematics, when the elements of some set S have a notion of equivalence (formalized as an equivalence

Congruence is an example of an equivalence relation. The leftmost two triangles are congruent, while the third and fourth triangles
are not congruent to any other triangle shown here. Thus, the rst two triangles are in the same equivalence class, while the third
and fourth triangles are each in their own equivalence class.

relation) dened on them, then one may naturally split the set S into equivalence classes. These equivalence classes
are constructed so that elements a and b belong to the same equivalence class if and only if a and b are equivalent.
Formally, given a set S and an equivalence relation ~ on S, the equivalence class of an element a in S is the set

{x S | x a}
of elements which are equivalent to a. It may be proven from the dening properties of equivalence relations that
the equivalence classes form a partition of S. This partition the set of equivalence classes is sometimes called the
quotient set or the quotient space of S by ~ and is denoted by S / ~.
When the set S has some structure (such as a group operation or a topology) and the equivalence relation ~ is dened in
a manner suitably compatible with this structure, then the quotient set often inherits a similar structure from its parent
set. Examples include quotient spaces in linear algebra, quotient spaces in topology, quotient groups, homogeneous
spaces, quotient rings, quotient monoids, and quotient categories.

16.1 Examples
If X is the set of all cars, and ~ is the equivalence relation has the same color as, then one particular equivalence
class consists of all green cars. X/~ could be naturally identied with the set of all car colors.

48
16.2. NOTATION AND FORMAL DEFINITION 49

Let X be the set of all rectangles in a plane, and ~ the equivalence relation has the same area as. For each
positive real number A there will be an equivalence class of all the rectangles that have area A.[1]
Consider the modulo 2 equivalence relation on the set Z of integers: x ~ y if and only if their dierence x y
is an even number. This relation gives rise to exactly two equivalence classes: one class consisting of all even
numbers, and the other consisting of all odd numbers. Under this relation [7], [9], and [1] all represent the
same element of Z/~.[2]
Let X be the set of ordered pairs of integers (a,b) with b not zero, and dene an equivalence relation ~ on X
according to which (a,b) ~ (c,d) if and only if ad = bc. Then the equivalence class of the pair (a,b) can be
identied with the rational number a/b, and this equivalence relation and its equivalence classes can be used to
give a formal denition of the set of rational numbers.[3] The same construction can be generalized to the eld
of fractions of any integral domain.
If X consists of all the lines in, say the Euclidean plane, and L ~ M means that L and M are parallel lines, then
the set of lines that are parallel to each other form an equivalence class as long as a line is considered parallel
to itself. In this situation, each equivalence class determines a point at innity.

16.2 Notation and formal denition


An equivalence relation is a binary relation ~ satisfying three properties:[4]

For every element a in X, a ~ a (reexivity),


For every two elements a and b in X, if a ~ b, then b ~ a (symmetry),
For every three elements a, b, and c in X, if a ~ b and b ~ c, then a ~ c (transitivity).

The equivalence class of an element a is denoted [a] and is dened as the set

[a] = {x X | a x}

of elements that are related to a by ~. An alternative notation [a]R can be used to denote the equivalence class of the
element a, specically with respect to the equivalence relation R. This is said to be the R-equivalence class of a.
The set of all equivalence classes in X with respect to an equivalence relation R is denoted as X/R and called X modulo
R (or the quotient set of X by R).[5] The surjective map x 7 [x] from X onto X/R, which maps each element to its
equivalence class, is called the canonical surjection or the canonical projection map.
When an element is chosen (often implicitly) in each equivalence class, this denes an injective map called a section. If
this section is denoted by s, one has [s(c)] = c for every equivalence class c. The element s(c) is called a representative
of c. Any element of a class may be chosen as a representative of the class, by choosing the section appropriately.
Sometimes, there is a section that is more natural than the other ones. In this case, the representatives are called
canonical representatives. For example, in modular arithmetic, consider the equivalence relation on the integers
dened by a ~ b if a b is a multiple of a given integer n, called the modulus. Each class contains a unique non-
negative integer smaller than n, and these integers are the canonical representatives. The class and its representative
are more or less identied, as is witnessed by the fact that the notation a mod n may denote either the class or its
canonical representative (which is the remainder of the division of a by n).

16.3 Properties
Every element x of X is a member of the equivalence class [x]. Every two equivalence classes [x] and [y] are either
equal or disjoint. Therefore, the set of all equivalence classes of X forms a partition of X: every element of X belongs
to one and only one equivalence class.[6] Conversely every partition of X comes from an equivalence relation in this
way, according to which x ~ y if and only if x and y belong to the same set of the partition.[7]
It follows from the properties of an equivalence relation that
50 CHAPTER 16. EQUIVALENCE CLASS

x ~ y if and only if [x] = [y].

In other words, if ~ is an equivalence relation on a set X, and x and y are two elements of X, then these statements
are equivalent:

xy

[x] = [y]

[x] [y] = .

16.4 Graphical representation

Graph of an example equivalence with 7 classes

An undirected graph may be associated to any symmetric relation on a set X, where the vertices are the elements of
X, and two vertices s and t are joined if and only if s ~ t. Among these graphs are the graphs of equivalence relations;
they are characterized as the graphs such that the connected components are cliques.[8]
16.5. INVARIANTS 51

16.5 Invariants
If ~ is an equivalence relation on X, and P(x) is a property of elements of X such that whenever x ~ y, P(x) is true if
P(y) is true, then the property P is said to be an invariant of ~, or well-dened under the relation ~.
A frequent particular case occurs when f is a function from X to another set Y; if f(x1 ) = f(x2 ) whenever x1 ~ x2 ,
then f is said to be a morphism for ~, a class invariant under ~, or simply invariant under ~. This occurs, e.g. in the
character theory of nite groups. Some authors use compatible with ~" or just respects ~" instead of invariant
under ~".
Any function f : X Y itself denes an equivalence relation on X according to which x1 ~ x2 if and only if f(x1 )
= f(x2 ). The equivalence class of x is the set of all elements in X which get mapped to f(x), i.e. the class [x] is the
inverse image of f(x). This equivalence relation is known as the kernel of f.
More generally, a function may map equivalent arguments (under an equivalence relation ~X on X) to equivalent
values (under an equivalence relation ~Y on Y). Such a function is known as a morphism from ~X to ~Y.

16.6 Quotient space in topology


In topology, a quotient space is a topological space formed on the set of equivalence classes of an equivalence relation
on a topological space using the original spaces topology to create the topology on the set of equivalence classes.
In abstract algebra, congruence relations on the underlying set of an algebra allow the algebra to induce an algebra
on the equivalence classes of the relation, called a quotient algebra. In linear algebra, a quotient space is a vector
space formed by taking a quotient group where the quotient homomorphism is a linear map. By extension, in abstract
algebra, the term quotient space may be used for quotient modules, quotient rings, quotient groups, or any quotient
algebra. However, the use of the term for the more general cases can as often be by analogy with the orbits of a group
action.
The orbits of a group action on a set may be called the quotient space of the action on the set, particularly when the
orbits of the group action are the right cosets of a subgroup of a group, which arise from the action of the subgroup
on the group by left translations, or respectively the left cosets as orbits under right translation.
A normal subgroup of a topological group, acting on the group by translation action, is a quotient space in the senses
of topology, abstract algebra, and group actions simultaneously.
Although the term can be used for any equivalence relations set of equivalence classes, possibly with further structure,
the intent of using the term is generally to compare that type of equivalence relation on a set X either to an equivalence
relation that induces some structure on the set of equivalence classes from a structure of the same kind on X, or to the
orbits of a group action. Both the sense of a structure preserved by an equivalence relation and the study of invariants
under group actions lead to the denition of invariants of equivalence relations given above.

16.7 See also


Equivalence partitioning, a method for devising test sets in software testing based on dividing the possible
program inputs into equivalence classes according to the behavior of the program on those inputs
Homogeneous space, the quotient space of Lie groups
Transversal (combinatorics)

16.8 Notes
[1] Avelsgaard 1989, p. 127

[2] Devlin 2004, p. 123

[3] Maddox 2002, pp. 7778

[4] Devlin 2004, p. 122


52 CHAPTER 16. EQUIVALENCE CLASS

[5] Wolf 1998, p. 178

[6] Maddox 2002, p. 74, Thm. 2.5.15

[7] Avelsgaard 1989, p. 132, Thm. 3.16

[8] Devlin 2004, p. 123

16.9 References
Avelsgaard, Carol (1989), Foundations for Advanced Mathematics, Scott Foresman, ISBN 0-673-38152-8
Devlin, Keith (2004), Sets, Functions, and Logic: An Introduction to Abstract Mathematics (3rd ed.), Chapman
& Hall/ CRC Press, ISBN 978-1-58488-449-1
Maddox, Randall B. (2002), Mathematical Thinking and Writing, Harcourt/ Academic Press, ISBN 0-12-
464976-9
Morash, Ronald P. (1987), Bridge to Abstract Mathematics, Random House, ISBN 0-394-35429-X

Wolf, Robert S. (1998), Proof, Logic and Conjecture: A Mathematicians Toolbox, Freeman, ISBN 978-0-
7167-3050-7

16.10 Further reading


This material is basic and can be found in any text dealing with the fundamentals of proof technique, such as any of
the following:

Sundstrom (2003), Mathematical Reasoning: Writing and Proof, Prentice-Hall

Smith; Eggen; St.Andre (2006), A Transition to Advanced Mathematics (6th Ed.), Thomson (Brooks/Cole)
Schumacher, Carol (1996), Chapter Zero: Fundamental Notions of Abstract Mathematics, Addison-Wesley,
ISBN 0-201-82653-4
O'Leary (2003), The Structure of Proof: With Logic and Set Theory, Prentice-Hall

Lay (2001), Analysis with an introduction to proof, Prentice Hall


Gilbert; Vanstone (2005), An Introduction to Mathematical Thinking, Pearson Prentice-Hall

Fletcher; Patty, Foundations of Higher Mathematics, PWS-Kent


Iglewicz; Stoyle, An Introduction to Mathematical Reasoning, MacMillan

D'Angelo; West (2000), Mathematical Thinking: Problem Solving and Proofs, Prentice Hall
Cupillari, The Nuts and Bolts of Proofs, Wadsworth

Bond, Introduction to Abstract Mathematics, Brooks/Cole


Barnier; Feldman (2000), Introduction to Advanced Mathematics, Prentice Hall

Ash, A Primer of Abstract Mathematics, MAA


Chapter 17

Equivalence relation

This article is about the mathematical concept. For the patent doctrine, see Doctrine of equivalents.
In mathematics, an equivalence relation is a binary relation that is at the same time a reexive relation, a symmetric
relation and a transitive relation. As a consequence of these properties an equivalence relation provides a partition of
a set into equivalence classes.

17.1 Notation
Although various notations are used throughout the literature to denote that two elements a and b of a set are equivalent
with respect to an equivalence relation R, the most common are "a ~ b" and "a b", which are used when R is the
obvious relation being referenced, and variations of "a ~R b", "a R b", or "aRb" otherwise.

17.2 Denition
A given binary relation ~ on a set X is said to be an equivalence relation if and only if it is reexive, symmetric and
transitive. That is, for all a, b and c in X:

a ~ a. (Reexivity)
a ~ b if and only if b ~ a. (Symmetry)
if a ~ b and b ~ c then a ~ c. (Transitivity)

X together with the relation ~ is called a setoid. The equivalence class of a under ~, denoted [a] , is dened as
[a] = {b X | a b} .

17.3 Examples

17.3.1 Simple example


Let the set {a, b, c} have the equivalence relation {(a, a), (b, b), (c, c), (b, c), (c, b)} . The following sets are equivalence
classes of this relation:
[a] = {a}, [b] = [c] = {b, c} .
The set of all equivalence classes for this relation is {{a}, {b, c}} .

17.3.2 Equivalence relations


The following are all equivalence relations:

53
54 CHAPTER 17. EQUIVALENCE RELATION

Has the same birthday as on the set of all people.

Is similar to on the set of all triangles.

Is congruent to on the set of all triangles.

Is congruent to, modulo n" on the integers.

Has the same image under a function" on the elements of the domain of the function.

Has the same absolute value on the set of real numbers

Has the same cosine on the set of all angles.

17.3.3 Relations that are not equivalences


The relation "" between real numbers is reexive and transitive, but not symmetric. For example, 7 5 does
not imply that 5 7. It is, however, a total order.

The relation has a common factor greater than 1 with between natural numbers greater than 1, is reexive
and symmetric, but not transitive. (Example: The natural numbers 2 and 6 have a common factor greater than
1, and 6 and 3 have a common factor greater than 1, but 2 and 3 do not have a common factor greater than 1).

The empty relation R on a non-empty set X (i.e. aRb is never true) is vacuously symmetric and transitive, but
not reexive. (If X is also empty then R is reexive.)

The relation is approximately equal to between real numbers, even if more precisely dened, is not an equiv-
alence relation, because although reexive and symmetric, it is not transitive, since multiple small changes can
accumulate to become a big change. However, if the approximation is dened asymptotically, for example by
saying that two functions f and g are approximately equal near some point if the limit of f g is 0 at that point,
then this denes an equivalence relation.

17.4 Connections to other relations


A partial order is a relation that is reexive, antisymmetric, and transitive.

Equality is both an equivalence relation and a partial order. Equality is also the only relation on a set that
is reexive, symmetric and antisymmetric. In algebraic expressions, equal variables may be substituted for
one another, a facility that is not available for equivalence related variables. The equivalence classes of an
equivalence relation can substitute for one another, but not individuals within a class.

A strict partial order is irreexive, transitive, and asymmetric.

A partial equivalence relation is transitive and symmetric. Transitive and symmetric imply reexive if and only
if for all a X, there exists a b X such that a ~ b.

A reexive and symmetric relation is a dependency relation, if nite, and a tolerance relation if innite.

A preorder is reexive and transitive.

A congruence relation is an equivalence relation whose domain X is also the underlying set for an algebraic
structure, and which respects the additional structure. In general, congruence relations play the role of kernels
of homomorphisms, and the quotient of a structure by a congruence relation can be formed. In many important
cases congruence relations have an alternative representation as substructures of the structure on which they
are dened. E.g. the congruence relations on groups correspond to the normal subgroups.

Any equivalence relation is the negation of an apartness relation, though the converse statement only holds in
classical mathematics (as opposed to constructive mathematics), since it is equivalent to the law of excluded
middle.
17.5. WELL-DEFINEDNESS UNDER AN EQUIVALENCE RELATION 55

17.5 Well-denedness under an equivalence relation


If ~ is an equivalence relation on X, and P(x) is a property of elements of X, such that whenever x ~ y, P(x) is true if
P(y) is true, then the property P is said to be well-dened or a class invariant under the relation ~.
A frequent particular case occurs when f is a function from X to another set Y; if x1 ~ x2 implies f(x1 ) = f(x2 ) then
f is said to be a morphism for ~, a class invariant under ~, or simply invariant under ~. This occurs, e.g. in the
character theory of nite groups. The latter case with the function f can be expressed by a commutative triangle. See
also invariant. Some authors use compatible with ~" or just respects ~" instead of invariant under ~".
More generally, a function may map equivalent arguments (under an equivalence relation ~A) to equivalent values
(under an equivalence relation ~B). Such a function is known as a morphism from ~A to ~B.

17.6 Equivalence class, quotient set, partition


Let a, b X . Some denitions:

17.6.1 Equivalence class


Main article: Equivalence class

A subset Y of X such that a ~ b holds for all a and b in Y, and never for a in Y and b outside Y, is called an equivalence
class of X by ~. Let [a] := {x X | a x} denote the equivalence class to which a belongs. All elements of X
equivalent to each other are also elements of the same equivalence class.

17.6.2 Quotient set


Main article: Quotient set

The set of all possible equivalence classes of X by ~, denoted X/ := {[x] | x X} , is the quotient set of X by
~. If X is a topological space, there is a natural way of transforming X/~ into a topological space; see quotient space
for the details.

17.6.3 Projection
Main article: Projection (relational algebra)

The projection of ~ is the function : X X/ dened by (x) = [x] which maps elements of X into their
respective equivalence classes by ~.

Theorem on projections:[1] Let the function f: X B be such that a ~ b f(a) = f(b). Then there is a
unique function g : X/~ B, such that f = g. If f is a surjection and a ~ b f(a) = f(b), then g is a
bijection.

17.6.4 Equivalence kernel


The equivalence kernel of a function f is the equivalence relation ~ dened by x y f (x) = f (y) . The
equivalence kernel of an injection is the identity relation.

17.6.5 Partition
Main article: Partition of a set
56 CHAPTER 17. EQUIVALENCE RELATION

A partition of X is a set P of nonempty subsets of X, such that every element of X is an element of a single element
of P. Each element of P is a cell of the partition. Moreover, the elements of P are pairwise disjoint and their union
is X.

Counting possible partitions

Let X be a nite set with n elements. Since every equivalence relation over X corresponds to a partition of X, and
vice versa, the number of possible equivalence relations on X equals the number of distinct partitions of X, which is
the nth Bell number Bn:


1 kn
Bn = ,
e k!
k=0

where the above is one of the ways to write the nth Bell number.

17.7 Fundamental theorem of equivalence relations


A key result links equivalence relations and partitions:[2][3][4]

An equivalence relation ~ on a set X partitions X.

Conversely, corresponding to any partition of X, there exists an equivalence relation ~ on X.

In both cases, the cells of the partition of X are the equivalence classes of X by ~. Since each element of X belongs
to a unique cell of any partition of X, and since each cell of the partition is identical to an equivalence class of X by
~, each element of X belongs to a unique equivalence class of X by ~. Thus there is a natural bijection between the
set of all possible equivalence relations on X and the set of all partitions of X.

17.8 Comparing equivalence relations


See also: Partition of a set Renement of partitions

If ~ and are two equivalence relations on the same set S, and a~b implies ab for all a,b S, then is said to be a
coarser relation than ~, and ~ is a ner relation than . Equivalently,

~ is ner than if every equivalence class of ~ is a subset of an equivalence class of , and thus every equivalence
class of is a union of equivalence classes of ~.

~ is ner than if the partition created by ~ is a renement of the partition created by .

The equality equivalence relation is the nest equivalence relation on any set, while the trivial relation that makes all
pairs of elements related is the coarsest.
The relation "~ is ner than " on the collection of all equivalence relations on a xed set is itself a partial order
relation, which makes the collection a geometric lattice.[5]

17.9 Generating equivalence relations


Given any binary relation A X X on X , the equivalence relation generated by A is the intersection of the
equivalence relations on X that contain A . (Since X X is an equivalence relation, the intersection is nontrivial.)
17.10. ALGEBRAIC STRUCTURE 57

Given any set X, there is an equivalence relation over the set [XX] of all possible functions XX. Two such
functions are deemed equivalent when their respective sets of xpoints have the same cardinality, corresponding
to cycles of length one in a permutation. Functions equivalent in this manner form an equivalence class on
[XX], and these equivalence classes partition [XX].

An equivalence relation ~ on X is the equivalence kernel of its surjective projection : X X/~.[6] Conversely,
any surjection between sets determines a partition on its domain, the set of preimages of singletons in the
codomain. Thus an equivalence relation over X, a partition of X, and a projection whose domain is X, are three
equivalent ways of specifying the same thing.

The intersection of any collection of equivalence relations over X (binary relations viewed as a subset of X X)
is also an equivalence relation. This yields a convenient way of generating an equivalence relation: given any
binary relation R on X, the equivalence relation generated by R is the smallest equivalence relation containing
R. Concretely, R generates the equivalence relation a ~ b if and only if there exist elements x1 , x2 , ..., xn in X
such that a = x1 , b = xn, and (xi,xi )R or (xi,xi)R, i = 1, ..., n1.

Note that the equivalence relation generated in this manner can be trivial. For instance, the equivalence
relation ~ generated by any total order on X has exactly one equivalence class, X itself, because x ~ y for
all x and y. As another example, any subset of the identity relation on X has equivalence classes that are
the singletons of X.

Equivalence relations can construct new spaces by gluing things together. Let X be the unit Cartesian square
[0,1] [0,1], and let ~ be the equivalence relation on X dened by a, b [0,1] ((a, 0) ~ (a, 1) (0, b) ~ (1, b)).
Then the quotient space X/~ can be naturally identied (homeomorphism) with a torus: take a square piece of
paper, bend and glue together the upper and lower edge to form a cylinder, then bend the resulting cylinder so
as to glue together its two open ends, resulting in a torus.

17.10 Algebraic structure


Much of mathematics is grounded in the study of equivalences, and order relations. Lattice theory captures the
mathematical structure of order relations. Even though equivalence relations are as ubiquitous in mathematics as
order relations, the algebraic structure of equivalences is not as well known as that of orders. The former structure
draws primarily on group theory and, to a lesser extent, on the theory of lattices, categories, and groupoids.

17.10.1 Group theory


Just as order relations are grounded in ordered sets, sets closed under pairwise supremum and inmum, equivalence
relations are grounded in partitioned sets, which are sets closed under bijections and preserve partition structure.
Since all such bijections map an equivalence class onto itself, such bijections are also known as permutations. Hence
permutation groups (also known as transformation groups) and the related notion of orbit shed light on the mathe-
matical structure of equivalence relations.
Let '~' denote an equivalence relation over some nonempty set A, called the universe or underlying set. Let G denote
the set of bijective functions over A that preserve the partition structure of A: x A g G (g(x) [x]). Then the
following three connected theorems hold:[7]

~ partitions A into equivalence classes. (This is the Fundamental Theorem of Equivalence Relations, mentioned
above);

Given a partition of A, G is a transformation group under composition, whose orbits are the cells of the parti-
tion;

Given a transformation group G over A, there exists an equivalence relation ~ over A, whose equivalence classes
are the orbits of G.[8][9]

In sum, given an equivalence relation ~ over A, there exists a transformation group G over A whose orbits are the
equivalence classes of A under ~.
58 CHAPTER 17. EQUIVALENCE RELATION

This transformation group characterisation of equivalence relations diers fundamentally from the way lattices char-
acterize order relations. The arguments of the lattice theory operations meet and join are elements of some universe
A. Meanwhile, the arguments of the transformation group operations composition and inverse are elements of a set
of bijections, A A.
Moving to groups in general, let H be a subgroup of some group G. Let ~ be an equivalence relation on G, such that a
~ b (ab1 H). The equivalence classes of ~also called the orbits of the action of H on Gare the right cosets
of H in G. Interchanging a and b yields the left cosets.
Proof.[10] Let function composition interpret group multiplication, and function inverse interpret group inverse.
Then G is a group under composition, meaning that x A g G ([g(x)] = [x]), because G satises the following
four conditions:

G is closed under composition. The composition of any two elements of G exists, because the domain and
codomain of any element of G is A. Moreover, the composition of bijections is bijective;[11]

Existence of identity function. The identity function, I(x)=x, is an obvious element of G;

Existence of inverse function. Every bijective function g has an inverse g1 , such that gg1 = I;

Composition associates. f(gh) = (fg)h. This holds for all functions over all domains.[12]

Let f and g be any two elements of G. By virtue of the denition of G, [g(f(x))] = [f(x)] and [f(x)] = [x], so that
[g(f(x))] = [x]. Hence G is also a transformation group (and an automorphism group) because function composition
preserves the partitioning of A.
Related thinking can be found in Rosen (2008: chpt. 10).

17.10.2 Categories and groupoids


Let G be a set and let "~" denote an equivalence relation over G. Then we can form a groupoid representing this
equivalence relation as follows. The objects are the elements of G, and for any two elements x and y of G, there exists
a unique morphism from x to y if and only if x~y.
The advantages of regarding an equivalence relation as a special case of a groupoid include:

Whereas the notion of free equivalence relation does not exist, that of a free groupoid on a directed graph
does. Thus it is meaningful to speak of a presentation of an equivalence relation, i.e., a presentation of the
corresponding groupoid;

Bundles of groups, group actions, sets, and equivalence relations can be regarded as special cases of the notion
of groupoid, a point of view that suggests a number of analogies;

In many contexts quotienting, and hence the appropriate equivalence relations often called congruences, are
important. This leads to the notion of an internal groupoid in a category.[13]

17.10.3 Lattices
The possible equivalence relations on any set X, when ordered by set inclusion, form a complete lattice, called Con
X by convention. The canonical map ker: X^X Con X, relates the monoid X^X of all functions on X and Con X.
ker is surjective but not injective. Less formally, the equivalence relation ker on X, takes each function f: XX to
its kernel ker f. Likewise, ker(ker) is an equivalence relation on X^X.

17.11 Equivalence relations and mathematical logic


Equivalence relations are a ready source of examples or counterexamples. For example, an equivalence relation with
exactly two innite equivalence classes is an easy example of a theory which is -categorical, but not categorical for
any larger cardinal number.
17.12. EUCLIDEAN RELATIONS 59

An implication of model theory is that the properties dening a relation can be proved independent of each other
(and hence necessary parts of the denition) if and only if, for each property, examples can be found of relations
not satisfying the given property while satisfying all the other properties. Hence the three dening properties of
equivalence relations can be proved mutually independent by the following three examples:

Reexive and transitive: The relation on N. Or any preorder;


Symmetric and transitive: The relation R on N, dened as aRb ab 0. Or any partial equivalence relation;
Reexive and symmetric: The relation R on Z, dened as aRb "a b is divisible by at least one of 2 or 3.
Or any dependency relation.

Properties denable in rst-order logic that an equivalence relation may or may not possess include:

The number of equivalence classes is nite or innite;


The number of equivalence classes equals the (nite) natural number n;
All equivalence classes have innite cardinality;
The number of elements in each equivalence class is the natural number n.

17.12 Euclidean relations


Euclid's The Elements includes the following Common Notion 1":

Things which equal the same thing also equal one another.

Nowadays, the property described by Common Notion 1 is called Euclidean (replacing equal by are in relation
with). By relation is meant a binary relation, in which aRb is generally distinct from bRa. A Euclidean relation
thus comes in two forms:

(aRc bRc) aRb (Left-Euclidean relation)


(cRa cRb) aRb (Right-Euclidean relation)

The following theorem connects Euclidean relations and equivalence relations:

Theorem If a relation is (left or right) Euclidean and reexive, it is also symmetric and transitive.
Proof for a left-Euclidean relation
(aRc bRc) aRb [a/c] = (aRa bRa) aRb [reexive; erase T] = bRa aRb. Hence R is symmetric.
(aRc bRc) aRb [symmetry] = (aRc cRb) aRb. Hence R is transitive.

with an analogous proof for a right-Euclidean relation. Hence an equivalence relation is a relation that is Euclidean
and reexive. The Elements mentions neither symmetry nor reexivity, and Euclid probably would have deemed the
reexivity of equality too obvious to warrant explicit mention.

17.13 See also


Apartness relation
Conjugacy class
Equipollence (geometry)
Topological conjugacy
Up to
60 CHAPTER 17. EQUIVALENCE RELATION

17.14 Notes
[1] Garrett Birkho and Saunders Mac Lane, 1999 (1967). Algebra, 3rd ed. p. 35, Th. 19. Chelsea.

[2] Wallace, D. A. R., 1998. Groups, Rings and Fields. p. 31, Th. 8. Springer-Verlag.

[3] Dummit, D. S., and Foote, R. M., 2004. Abstract Algebra, 3rd ed. p. 3, Prop. 2. John Wiley & Sons.

[4] Karel Hrbacek & Thomas Jech (1999) Introduction to Set Theory, 3rd edition, pages 2932, Marcel Dekker

[5] Birkho, Garrett (1995), Lattice Theory, Colloquium Publications, 25 (3rd ed.), American Mathematical Society, ISBN
9780821810255. Sect. IV.9, Theorem 12, page 95

[6] Garrett Birkho and Saunders Mac Lane, 1999 (1967). Algebra, 3rd ed. p. 33, Th. 18. Chelsea.

[7] Rosen (2008), pp. 24345. Less clear is 10.3 of Bas van Fraassen, 1989. Laws and Symmetry. Oxford Univ. Press.

[8] Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 202, Th. 6.

[9] Dummit, D. S., and Foote, R. M., 2004. Abstract Algebra, 3rd ed. John Wiley & Sons: 114, Prop. 2.

[10] Bas van Fraassen, 1989. Laws and Symmetry. Oxford Univ. Press: 246.

[11] Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 22, Th. 6.

[12] Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 24, Th. 7.

[13] Borceux, F. and Janelidze, G., 2001. Galois theories, Cambridge University Press, ISBN 0-521-80309-8

17.15 References
Brown, Ronald, 2006. Topology and Groupoids. Booksurge LLC. ISBN 1-4196-2722-8.

Castellani, E., 2003, Symmetry and equivalence in Brading, Katherine, and E. Castellani, eds., Symmetries
in Physics: Philosophical Reections. Cambridge Univ. Press: 422-433.

Robert Dilworth and Crawley, Peter, 1973. Algebraic Theory of Lattices. Prentice Hall. Chpt. 12 discusses
how equivalence relations arise in lattice theory.

Higgins, P.J., 1971. Categories and groupoids. Van Nostrand. Downloadable since 2005 as a TAC Reprint.
John Randolph Lucas, 1973. A Treatise on Time and Space. London: Methuen. Section 31.

Rosen, Joseph (2008) Symmetry Rules: How Science and Nature are Founded on Symmetry. Springer-Verlag.
Mostly chpts. 9,10.

Raymond Wilder (1965) Introduction to the Foundations of Mathematics 2nd edition, Chapter 2-8: Axioms
dening equivalence, pp 4850, John Wiley & Sons.

17.16 External links


Hazewinkel, Michiel, ed. (2001) [1994], Equivalence relation, Encyclopedia of Mathematics, Springer Sci-
ence+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Bogomolny, A., "Equivalence Relationship" cut-the-knot. Accessed 1 September 2009

Equivalence relation at PlanetMath


Binary matrices representing equivalence relations at OEIS.
17.16. EXTERNAL LINKS 61
Chapter 18

Euclidean relation

In mathematics, Euclidean relations are a class of binary relations that formalizes Euclid's Common Notion 1 in
The Elements: things which equal the same thing also equal one another.

18.1 Denition
A binary relation R on a set X is Euclidean (sometimes called right Euclidean) if it satises the following: for every
a, b, c in X, if a is related to b and c, then b is related to c.[1]
To write this in predicate logic:

a, b, c X (a R b a R c b R c).

Dually, a relation R on X is left Euclidean if for every a, b, c in X, if b is related to a and c is related to a, then b is
related to c:

a, b, c X (b R a c R a b R c).

18.2 Relation to transitivity


The property of being Euclidean is dierent from transitivity. A transitive relation is Euclidean only if it is also
symmetric. Only a symmetric Euclidean relation is transitive.
A relation which is both Euclidean and reexive is also symmetric and therefore an equivalence relation.[1]

18.3 References
[1] Fagin, Ronald (2003), Reasoning About Knowledge, MIT Press, p. 60, ISBN 978-0-262-56200-3.

62
Chapter 19

Foundational relation

In set theory, a foundational relation on a set or proper class lets each nonempty subset admit a relational minimal
element.
Formally, let (A, R) be a binary relation structure, where A is a class (set or proper class), and R is a binary relation
dened on A. Then (A, R) is a foundational relation if and only if any nonempty subset in A has a R-minimal element.
In predicate logic,
( )
(S) S A S = (x S)(S R1 {x} = ) , [1]

in which denotes the empty set, and R1 {x} denotes the class of the elements that precede x in the relation R. That
is,

R1 {x} = {y|yRx}. [2]

Here x is an R-minimal element in the subset S, since none of its R-predecessors is in S.

19.1 See also


Binary relation
Well-order

19.2 References
[1] See Denition 6.21 in Zaring W.M., G. Takeuti (1971). Introduction to axiomatic set theory (2nd, rev. ed.). New York:
Springer-Verlag. ISBN 0387900241.

[2] See Theorem 6.19 and Denition 6.20 in Zaring W.M., G. Takeuti (1971). Introduction to axiomatic set theory (2nd, rev.
ed.). New York: Springer-Verlag. ISBN 0387900241.

63
Chapter 20

Idempotent relation

In mathematics, an idempotent binary relation R X X is one for which R R = R.[1][2] This notion generalizes
that of an idempotent function to relations. Each idempotent relation is necessarily transitive, as the latter means R
R R.
For example, the relation < on is idempotent. In contrast, < on is not, since (<) (<) (<) does not hold: e.g. 1
< 2, but 1 < x < 2 is false for every x .
Idempotent relations have been used as an example to illustrate the application of Mechanized Formalisation of math-
ematics using the interactive theorem prover Isabelle/HOL. Besides checking the mathematical properties of nite
idempotent relations, an algorithm for counting the number of idempotent relations has been derived in Isabelle/HOL.
[3][4]

20.1 References
[1] Florian Kammller, J. W. Sanders (2004). Idempotent Relation in Isabelle/HOL (PDF) (Technical report). TU Berlin. p.
27. 2004-04. Here:p.3

[2] Florian Kammller (2011). Mechanical Analysis of Finite Idempotent Relations. Fundamenta Informaticae. 107. pp.
4365. doi:10.3233/FI-2011-392.

[3] Florian Kammller (2006). Number of idempotent relations on n labeled elements. The On-Line Ecyclopedea of Integer
Sequences (A12137).

[4] Florian Kammller (2008). Counting Idempotent Relations (PDF) (Technical report). TU Berlin. p. 27. 2008-15.

Berstel, Jean; Perrin, Dominique; Reutenauer, Christophe (2010). Codes and automata. Encyclopedia of
Mathematics and its Applications. 129. Cambridge: Cambridge University Press. p. 330. ISBN 978-0-521-
88831-8. Zbl 1187.94001.

64
Chapter 21

Intransitivity

This article is about intransitivity in mathematics. For other uses, see Intransitive (disambiguation).

In mathematics, intransitivity (sometimes called nontransitivity) is a property of binary relations that are not
transitive relations. This may include any relation that is not transitive, or the stronger property of antitransitiv-
ity, which describes a relation that is never transitive.

21.1 Intransitivity
A relation is transitive if, whenever it relates some A to some B, and that B to some C, it also relates that A to that C.
Some authors call a relation intransitive if it is not transitive, i.e. (if the relation in question is named R )

(a, b, c : aRb bRc aRc) .

This statement is equivalent to

a, b, c : aRb bRc (aRc)

For instance, in the food chain, wolves feed on deer, and deer feed on grass, but wolves do not feed on grass.[1] Thus,
the feed on relation among life forms is intransitive, in this sense.
Another example that does not involve preference loops arises in freemasonry: in some instances lodge A recognizes
lodge B, and lodge B recognizes lodge C, but lodge A does not recognize lodge C. Thus the recognition relation
among Masonic lodges is intransitive.

21.2 Antitransitivity
Often the term intransitive is used to refer to the stronger property of antitransitivity.
We just saw that the feed on relation is not transitive, but it still contains some transitivity: for instance: humans feed
on rabbits, rabbits feed on carrots, and humans also feed on carrots.
A relation is antitransitive if this never occurs at all, i.e.,

a, b, c : aRb bRc aRc

Many authors use the term intransitivity to mean antitransitivity.[2][3]


An example of an antitransitive relation: the defeated relation in knockout tournaments. If player A defeated player
B and player B defeated player C, A can have never played C, and therefore, A has not defeated C.

65
66 CHAPTER 21. INTRANSITIVITY

21.3 Cycles
The term intransitivity is often used when speaking of scenarios in which a relation describes the relative preferences
between pairs of options, and weighing several options produces a loop of preference:

A is preferred to B
B is preferred to C
C is preferred to A

Rock, paper, scissors; Nontransitive dice; and Penneys game are examples. Real combative relations of competing
species [4] and strategies of individual animals [5] can be cyclic as well.
Assuming no option is preferred to itself i.e. the relation is irreexive, a preference relation with a loop is not transitive.
For if it is, each option in the loop is preferred to each option, including itself. This can be illustrated for this example
of a loop among A, B, and C. Assume the relation is transitive. Then, since A is preferred to B and B is preferred to
C, also A is preferred to C. But then, since C is preferred to A, also A is preferred to A.
Therefore such a preference loop (or "cycle") is known as an intransitivity.
Notice that a cycle is neither necessary nor sucient for a binary relation to be not transitive. For example, an
equivalence relation possesses cycles but is transitive. Now, consider the relation is an enemy of and suppose that
the relation is symmetric and satises the condition that for any country, any enemy of an enemy of the country is
not itself an enemy of the country. This is an example of an antitransitive relation that does not have any cycles. In
particular, by virtue of being antitransitive the relation is not transitive.
Finally, let us work with the example of rock, paper, scissors, calling the three options A, B, and C. Now, the relation
over A, B, and C is defeats and the standard rules of the game are such that A defeats B, B defeats C, and C defeats
A. Furthermore, it is also true that B does not defeat A, C does not defeat B, and A does not defeat C. Finally, it is
also true that no option defeats itself. This information can be depicted in a table:
The rst argument of the relation is a row and the second one is a column. Ones indicate the relation holds, zero
indicates that it does not hold. Now, notice that the following statement is true for any pair of elements x and y drawn
(with replacement) from the set {A, B, C}: If x defeats y, and y defeats z, then x does not defeat z. Hence the relation
is antitransitive.
Thus, a cycle is neither necessary nor sucient for a binary relation to be antitransitive.

21.4 Occurrences in preferences


Intransitivity can occur under majority rule, in probabilistic outcomes of game theory, and in the Condorcet
voting method in which ranking several candidates can produce a loop of preference when the weights are
compared (see voting paradox). Intransitive dice demonstrate that probabilities are not necessarily transitive.
In psychology, intransitivity often occurs in a persons system of values (or preferences, or tastes), potentially
leading to unresolvable conicts.
Analogously, in economics intransitivity can occur in a consumers preferences. This may lead to consumer
behaviour that does not conform to perfect economic rationality. In recent years, economists and philosophers
have questioned whether violations of transitivity must necessarily lead to 'irrational behaviour' (see Anand
(1993)).

21.5 Likelihood
It has been suggested that Condorcet voting tends to eliminate intransitive loops when large numbers of voters
participate because the overall assessment criteria for voters balances out. For instance, voters may prefer candidates
on several dierent units of measure such as by order of social consciousness or by order of most scally conservative.
In such cases intransitivity reduces to a broader equation of numbers of people and the weights of their units of
measure in assessing candidates.
21.6. REFERENCES 67

Such as:

30% favor 60/40 weighting between social consciousness and scal conservatism

50% favor 50/50 weighting between social consciousness and scal conservatism
20% favor a 40/60 weighting between social consciousness and scal conservatism

While each voter may not assess the units of measure identically, the trend then becomes a single vector on which
the consensus agrees is a preferred balance of candidate criteria.

21.6 References
[1] Wolves do in fact eat grass see Engel, Cindy (2003). Wild Health: Lessons in Natural Wellness from the Animal Kingdom
(paperback ed.). Houghton Miin. p. 141. ISBN 0-618-34068-8..

[2] Guide to Logic, Relations II

[3] IntransitiveRelation

[4] Kerr B., Riley M.A., Feldman M.W., & Bohannan B.J.M. (2002). Local dispersal promotes biodiversity in a real-life game
of rockpaperscissors. Nature. 418(171174)

[5] Leutwyler, K. (2000). Mating Lizards Play a Game of Rock-Paper-Scissors. Scientic American.

21.7 Further reading


Anand, P (1993). Foundations of Rational Choice Under Risk. Oxford: Oxford University Press..
Poddiakov, A., & Valsiner, J. (2013). Intransitivity cycles and their transformations: How dynamically adapting
systems function. In: L. Rudolph (Ed.). Qualitative mathematics for the social sciences: Mathematical models
for research on cultural dynamics. Abingdon, NY: Routledge. Pp. 343391.
Chapter 22

Inverse relation

For inverse relationships in statistics, see negative relationship.

In mathematics, the inverse relation of a binary relation is the relation that occurs when the order of the elements is
switched in the relation. For example, the inverse of the relation 'child of' is the relation 'parent of'. In formal terms,
if X and Y are sets and L X Y is a relation from X to Y, then L1 is the relation dened so that y L1 x if and only
if x L y. In set-builder notation, L1 = {(y, x) Y X | (x, y) L}.
The notation is analogous with that for an inverse function. Although many functions do not have an inverse, every
relation does have a unique inverse. Despite the notation and terminology, the inverse relation is not an inverse in
the sense of group inverse. However, the unary operation that maps a relation to the inverse relation is an involution,
so it induces the structure of a semigroup with involution on the binary relations on a set, or, more generally, induces
a dagger category on the category of relations as detailed below. As a unary operation, taking the inverse (some-
times called inversion) commutes with the order-related operations of relation algebra, i.e., it commutes with union,
intersection, complement etc.
The inverse relation is also called the converse relation or transpose relation the latter in view of its similarity
with the transpose of a matrix.[1] It has also been called the opposite or dual of the original relation.[2] Other notations
for the inverse relation include LC , LT , L~ or L or L or L .

22.1 Examples

For usual (maybe strict or partial) order relations, the converse is the naively expected opposite order, e.g. 1 =
, <1 = > , etc.

22.1.1 Inverse relation of a function

A function is invertible if and only if its inverse relation is a function, in which case the inverse relation is the inverse
function.
The inverse relation of a function f : X Y is the relation f 1 : Y X dened by graph f 1 = {(y, x) | y =
f (x)} .
This is not necessarily a function: One necessary condition is that f be injective, since else f 1 is multi-valued. This
condition is sucient for f 1 being a partial function, and it is clear that f 1 then is a (total) function if and only if
f is surjective. In that case, i.e. if f is bijective, f 1 may be called the inverse function of f.
For example, the function f (x) = 2x + 2 has the inverse function f 1 (x) = x/2 1 .

68
22.2. PROPERTIES 69

22.2 Properties
In the monoid of binary endorelations on a set (with the binary operation on relations being the composition of
relations), the inverse relation does not satisfy the denition of an inverse from group theory, i.e. if L is an arbitrary
relation on X, then L L1 does not equal the identity relation on X in general. The inverse relation does satisfy the
(weaker) axioms of a semigroup with involution: (L1 )1 = L and (L R)1 = R1 L1 .[3]
Since one may generally consider relations between dierent sets (which form a category rather than a monoid,
namely the category of relations Rel), in this context the inverse relation conforms to the axioms of a dagger category
(aka category with involution).[3] A relation equal to its inverse is a symmetric relation; in the language of dagger
categories, it is self-adjoint.
Furthermore, the semigroup of endorelations on a set is also a partially ordered structure (with inclusion of relations
as sets), and actually an involutive quantale. Similarly, the category of heterogenous relations, Rel is also an ordered
category.[3]
In relation algebra (which is an abstraction of the properties of the algebra of endorelations on a set), inversion (the
operation of taking the inverse relation) commutes with other binary operations of union and intersection. Inversion
also commutes with unary operation of complementation as well as with taking suprema and inma. Inversion is also
compatible with the ordering of relations by inclusion.[1]
If a relation is reexive, irreexive, symmetric, antisymmetric, asymmetric, transitive, total, trichotomous, a partial
order, total order, strict weak order, total preorder (weak order), or an equivalence relation, its inverse is too.

22.3 See also


Bijection
Function (mathematics)

Inverse function
Relation (mathematics)

Transpose graph

22.4 References
[1] Gunther Schmidt; Thomas Strhlein (1993). Relations and Graphs: Discrete Mathematics for Computer Scientists. Springer
Berlin Heidelberg. pp. 910. ISBN 978-3-642-77970-1.

[2] Celestina Cotti Ferrero; Giovanni Ferrero (2002). Nearrings: Some Developments Linked to Semigroups and Groups.
Kluwer Academic Publishers. p. 3. ISBN 978-1-4613-0267-4.

[3] Joachim Lambek (2001). Relations Old and New. In Ewa Orlowska, Andrzej Szalas. Relational Methods for Computer
Science Applications. Springer Science & Business Media. pp. 135146. ISBN 978-3-7908-1365-4.

Halmos, Paul R. (1974), Naive Set Theory, p. 40, ISBN 978-0-387-90092-6


Chapter 23

Partially ordered set

{x,y,z}

{x,y} {x,z} {y,z}

{x} {y} {z}

The Hasse diagram of the set of all subsets of a three-element set {x, y, z}, ordered by inclusion. Sets on the same horizontal level
are incomparable with each other. Some other pairs, such as {x} and {y,z}, are also incomparable.

In mathematics, especially order theory, a partially ordered set (also poset) formalizes and generalizes the intuitive
concept of an ordering, sequencing, or arrangement of the elements of a set. A poset consists of a set together with
a binary relation indicating that, for certain pairs of elements in the set, one of the elements precedes the other in
the ordering. The word partial in the names partial order or partially ordered set is used as an indication that
not every pair of elements need be comparable. That is, there may be pairs of elements for which neither element
precedes the other in the poset. Partial orders thus generalize total orders, in which every pair is comparable.
To be a partial order, a binary relation must be reexive (each element is comparable to itself), antisymmetric (no
two dierent elements precede each other), and transitive (the start of a chain of precedence relations must precede
the end of the chain).
One familiar example of a partially ordered set is a collection of people ordered by genealogical descendancy. Some

70
23.1. FORMAL DEFINITION 71

pairs of people bear the descendant-ancestor relationship, but other pairs of people are incomparable, with neither
being a descendent of the other.
A poset can be visualized through its Hasse diagram, which depicts the ordering relation.[1]

23.1 Formal denition


A (non-strict) partial order[2] is a binary relation over a set P satisfying particular axioms which are discussed
below. When a b, we say that a is related to b. (This does not imply that b is also related to a, because the relation
need not be symmetric.)
The axioms for a non-strict partial order state that the relation is reexive, antisymmetric, and transitive. That is,
for all a, b, and c in P, it must satisfy:

1. a a (reexivity: every element is related to itself).


2. if a b and b a, then a = b (antisymmetry: two distinct elements cannot be related in both directions).
3. if a b and b c, then a c (transitivity: if a rst element is related to a second element, and, in turn, that
element is related to a third element, then the rst element is related to the third element).

In other words, a partial order is an antisymmetric preorder.


A set with a partial order is called a partially ordered set (also called a poset). The term ordered set is sometimes
also used, as long as it is clear from the context that no other kind of order is meant. In particular, totally ordered sets
can also be referred to as ordered sets, especially in areas where these structures are more common than posets.
For a, b, elements of a partially ordered set P, if a b or b a, then a and b are comparable. Otherwise they are
incomparable. In the gure on top-right, e.g. {x} and {x,y,z} are comparable, while {x} and {y} are not. A partial
order under which every pair of elements is comparable is called a total order or linear order; a totally ordered
set is also called a chain (e.g., the natural numbers with their standard order). A subset of a poset in which no two
distinct elements are comparable is called an antichain (e.g. the set of singletons {{x}, {y}, {z}} in the top-right
gure). An element a is said to be covered by another element b, written a<:b, if a is strictly less than b and no third
element c ts between them; formally: if both ab and ab are true, and acb is false for each c with acb. A
more concise denition will be given below using the strict order corresponding to "". For example, {x} is covered
by {x,z} in the top-right gure, but not by {x,y,z}.

23.2 Examples
Standard examples of posets arising in mathematics include:

The real numbers ordered by the standard less-than-or-equal relation (a totally ordered set as well).
The set of subsets of a given set (its power set) ordered by inclusion (see the gure on top-right). Similarly, the
set of sequences ordered by subsequence, and the set of strings ordered by substring.
The set of natural numbers equipped with the relation of divisibility.
The vertex set of a directed acyclic graph ordered by reachability.
The set of subspaces of a vector space ordered by inclusion.
For a partially ordered set P, the sequence space containing all sequences of elements from P, where sequence
a precedes sequence b if every item in a precedes the corresponding item in b. Formally, (an)n (bn)n
if and only if an bn for all n in , i.e. a componentwise order.
For a set X and a partially ordered set P, the function space containing all functions from X to P, where f g
if and only if f(x) g(x) for all x in X.
A fence, a partially ordered set dened by an alternating sequence of order relations a < b > c < d ...
The set of events in special relativity, where for two events X and Y, X Y if and only if Y is in the future
light cone of X. An event Y can only be causally aected by X if X Y.
72 CHAPTER 23. PARTIALLY ORDERED SET

23.3 Extrema
There are several notions of greatest and least element in a poset P, notably:

Greatest element and least element: An element g in P is a greatest element if for every element a in P, a g.
An element m in P is a least element if for every element a in P, a m. A poset can only have one greatest or
least element.
Maximal elements and minimal elements: An element g in P is a maximal element if there is no element a in
P such that a > g. Similarly, an element m in P is a minimal element if there is no element a in P such that a <
m. If a poset has a greatest element, it must be the unique maximal element, but otherwise there can be more
than one maximal element, and similarly for least elements and minimal elements.
Upper and lower bounds: For a subset A of P, an element x in P is an upper bound of A if a x, for each
element a in A. In particular, x need not be in A to be an upper bound of A. Similarly, an element x in P is a
lower bound of A if a x, for each element a in A. A greatest element of P is an upper bound of P itself, and
a least element is a lower bound of P.

For example, consider the positive integers, ordered by divisibility: 1 is a least element, as it divides all other elements;
on the other hand this poset does not have a greatest element (although if one would include 0 in the poset, which
is a multiple of any integer, that would be a greatest element; see gure). This partially ordered set does not even
have any maximal elements, since any g divides for instance 2g, which is distinct from it, so g is not maximal. If the
number 1 is excluded, while keeping divisibility as ordering on the elements greater than 1, then the resulting poset
does not have a least element, but any prime number is a minimal element for it. In this poset, 60 is an upper bound
(though not a least upper bound) of the subset {2,3,5,10}, which does not have any lower bound (since 1 is not in the
poset); on the other hand 2 is a lower bound of the subset of powers of 2, which does not have any upper bound.

23.4 Orders on the Cartesian product of partially ordered sets


In order of increasing strength, i.e., decreasing sets of pairs, three of the possible partial orders on the Cartesian
product of two partially ordered sets are (see gures):

the lexicographical order: (a,b) (c,d) if a < c or (a = c and b d);


the product order: (a,b) (c,d) if a c and b d;
the reexive closure of the direct product of the corresponding strict orders: (a,b) (c,d) if (a < c and b < d)
or (a = c and b = d).

All three can similarly be dened for the Cartesian product of more than two sets.
Applied to ordered vector spaces over the same eld, the result is in each case also an ordered vector space.
See also orders on the Cartesian product of totally ordered sets.

23.5 Sums of partially ordered sets


Another way to combine two posets is the ordinal sum[3] (or linear sum[4] ), Z = X Y, dened on the union of the
underlying sets X and Y by the order a Z b if and only if:

a, b X with a X b, or
a, b Y with a Y b, or
a X and b Y.

If two posets are well-ordered, then so is their ordinal sum.[5] The ordinal sum operation is one of two operations
used to form series-parallel partial orders, and in this context is called series composition. The other operation used
to form these orders, the disjoint union of two partially ordered sets (with no order relation between elements of one
set and elements of the other set) is called in this context parallel composition.
23.6. STRICT AND NON-STRICT PARTIAL ORDERS 73

Series composition

Parallel composition

Hasse diagram of a series-parallel partial order, formed as the ordinal sum of three smaller partial orders.

23.6 Strict and non-strict partial orders


In some contexts, the partial order dened above is called a non-strict (or reexive, or weak) partial order. In these
contexts, a strict (or irreexive) partial order "<" is a binary relation that is irreexive, transitive and asymmetric,
i.e. which satises for all a, b, and c in P:

not a < a (irreexivity),


if a < b and b < c then a < c (transitivity), and
if a < b then not b < a (asymmetry; implied by irreexivity and transitivity[6] ).

Strict and non-strict partial orders are closely related. A non-strict partial order may be converted to a strict partial
order by removing all relationships of the form a a. Conversely, a strict partial order may be converted to a non-
strict partial order by adjoining all relationships of that form. Thus, if "" is a non-strict partial order, then the
corresponding strict partial order "<" is the irreexive kernel given by:

a < b if a b and a b
74 CHAPTER 23. PARTIALLY ORDERED SET

Conversely, if "<" is a strict partial order, then the corresponding non-strict partial order "" is the reexive closure
given by:

a b if a < b or a = b.

This is the reason for using the notation "".


Using the strict order "<", the relation "a is covered by b" can be equivalently rephrased as "a<b, but not a<c<b for
any c". Strict partial orders are useful because they correspond more directly to directed acyclic graphs (dags): every
strict partial order is a dag, and the transitive closure of a dag is both a strict partial order and also a dag itself.

23.7 Inverse and order dual


The inverse or converse of a partial order relation satises xy if and only if yx. The inverse of a partial
order relation is reexive, transitive, and antisymmetric, and hence itself a partial order relation. The order dual of a
partially ordered set is the same set with the partial order relation replaced by its inverse. The irreexive relation > is
to as < is to .
Any one of the four relations , <, , and > on a given set uniquely determines the other three.
In general two elements x and y of a partial order may stand in any of four mutually exclusive relationships to each
other: either x < y, or x = y, or x > y, or x and y are incomparable (none of the other three). A totally ordered set is one
that rules out this fourth possibility: all pairs of elements are comparable and we then say that trichotomy holds. The
natural numbers, the integers, the rationals, and the reals are all totally ordered by their algebraic (signed) magnitude
whereas the complex numbers are not. This is not to say that the complex numbers cannot be totally ordered; we
could for example order them lexicographically via x+iy < u+iv if and only if x < u or (x = u and y < v), but this is not
ordering by magnitude in any reasonable sense as it makes 1 greater than 100i. Ordering them by absolute magnitude
yields a preorder in which all pairs are comparable, but this is not a partial order since 1 and i have the same absolute
magnitude but are not equal, violating antisymmetry.

23.8 Mappings between partially ordered sets


Given two partially ordered sets (S,) and (T,), a function f: S T is called order-preserving, or monotone,
or isotone, if for all x and y in S, xy implies f(x) f(y). If (U,) is also a partially ordered set, and both f: S
T and g: T U are order-preserving, their composition (gf): S U is order-preserving, too. A function f:
S T is called order-reecting if for all x and y in S, f(x) f(y) implies xy. If f is both order-preserving and
order-reecting, then it is called an order-embedding of (S,) into (T,). In the latter case, f is necessarily injective,
since f(x) = f(y) implies x y and y x. If an order-embedding between two posets S and T exists, one says that S
can be embedded into T. If an order-embedding f: S T is bijective, it is called an order isomorphism, and the
partial orders (S,) and (T,) are said to be isomorphic. Isomorphic orders have structurally similar Hasse diagrams
(cf. right picture). It can be shown that if order-preserving maps f: S T and g: T S exist such that gf and fg
yields the identity function on S and T, respectively, then S and T are order-isomorphic. [7]
For example, a mapping f: () from the set of natural numbers (ordered by divisibility) to the power set of
natural numbers (ordered by set inclusion) can be dened by taking each number to the set of its prime divisors. It
is order-preserving: if x divides y, then each prime divisor of x is also a prime divisor of y. However, it is neither
injective (since it maps both 12 and 6 to {2,3}) nor order-reecting (since besides 12 doesn't divide 6). Taking
instead each number to the set of its prime power divisors denes a map g: () that is order-preserving, order-
reecting, and hence an order-embedding. It is not an order-isomorphism (since it e.g. doesn't map any number to
the set {4}), but it can be made one by restricting its codomain to g(). The right picture shows a subset of and its
isomorphic image under g. The construction of such an order-isomorphism into a power set can be generalized to a
wide class of partial orders, called distributive lattices, see "Birkhos representation theorem".

23.9 Number of partial orders


Sequence A001035 in OEIS gives the number of partial orders on a set of n labeled elements:
23.10. LINEAR EXTENSION 75

The number of strict partial orders is the same as that of partial orders.
If the count is made only up to isomorphism, the seqeuence 1, 1, 2, 5, 16, 63, 318, (sequence A000112 in the
OEIS) is obtained.

23.10 Linear extension


A partial order * on a set X is an extension of another partial order on X provided that for all elements x and y
of X, whenever x y, it is also the case that x * y. A linear extension is an extension that is also a linear (i.e., total)
order. Every partial order can be extended to a total order (order-extension principle).[8]
In computer science, algorithms for nding linear extensions of partial orders (represented as the reachability orders
of directed acyclic graphs) are called topological sorting.

23.11 In category theory


Every poset (and every preorder) may be considered as a category in which every hom-set has at most one element.
More explicitly, let hom(x, y) = {(x, y)} if x y (and otherwise the empty set) and (y, z)(x, y) = (x, z). Such
categories are sometimes called posetal.
Posets are equivalent to one another if and only if they are isomorphic. In a poset, the smallest element, if it exists,
is an initial object, and the largest element, if it exists, is a terminal object. Also, every preordered set is equivalent
to a poset. Finally, every subcategory of a poset is isomorphism-closed.

23.12 Partial orders in topological spaces


Main article: Partially ordered space

If P is a partially ordered set that has also been given the structure of a topological space, then it is customary to
assume that {(a, b) : a b} is a closed subset of the topological product space P P . Under this assumption partial
order relations are well behaved at limits in the sense that if ai a , bi b and ai bi for all i, then a b.[9]

23.13 Interval
For a b, the closed interval [a,b] is the set of elements x satisfying a x b (i.e. a x and x b). It contains at
least the elements a and b.
Using the corresponding strict relation "<", the open interval (a,b) is the set of elements x satisfying a < x < b (i.e. a
< x and x < b). An open interval may be empty even if a < b. For example, the open interval (1,2) on the integers is
empty since there are no integers i such that 1 < i < 2.
Sometimes the denitions are extended to allow a > b, in which case the interval is empty.
The half-open intervals [a,b) and (a,b] are dened similarly.
A poset is locally nite if every interval is nite. For example, the integers are locally nite under their natural order-
ing. The lexicographical order on the cartesian product is not locally nite, since e.g. (1,2)(1,3)(1,4)(1,5)...(2,1).
Using the interval notation, the property "a is covered by b" can be rephrased equivalently as [a,b] = {a,b}.
This concept of an interval in a partial order should not be confused with the particular class of partial orders known
as the interval orders.

23.14 See also


antimatroid, a formalization of orderings on a set that allows more general families of orderings than posets
76 CHAPTER 23. PARTIALLY ORDERED SET

causal set

comparability graph

complete partial order

directed set

graded poset

incidence algebra

lattice

locally nite poset

Mbius function on posets

ordered group

poset topology, a kind of topological space that can be dened from any poset

Scott continuity continuity of a function between two partial orders.

semilattice

semiorder

stochastic dominance

strict weak ordering strict partial order "<" in which the relation neither a < b nor b < a" is transitive.

Zorns lemma

23.15 Notes
[1] Merrield, Richard E.; Simmons, Howard E. (1989). Topological Methods in Chemistry. New York: John Wiley & Sons.
p. 28. ISBN 0-471-83817-9. Retrieved 27 July 2012. A partially ordered set is conveniently represented by a Hasse
diagram...

[2] Simovici, Dan A. & Djeraba, Chabane (2008). Partially Ordered Sets. Mathematical Tools for Data Mining: Set Theory,
Partial Orders, Combinatorics. Springer. ISBN 9781848002012.

[3] Neggers, J.; Kim, Hee Sik (1998), 4.2 Product Order and Lexicographic Order, Basic Posets, World Scientic, pp. 6263,
ISBN 9789810235895

[4] Davey, B. A.; Priestley, H. A., Introduction to Lattices and Order (Second Edition), 2002, p. 17-18

[5] P. R. Halmos (1974). Naive Set Theory. Springer. p. 82. ISBN 978-1-4757-1645-0.

[6] Flaka, V.; Jeek, J.; Kepka, T.; Kortelainen, J. (2007). Transitive Closures of Binary Relations I (PDF). Prague: School
of Mathematics - Physics Charles University. p. 1. Lemma 1.1 (iv). Note that this source refers to asymmetric relations
as strictly antisymmetric.

[7] Davey, B. A.; Priestley, H. A. (2002). Maps between ordered sets. Introduction to Lattices and Order (2nd ed.). New
York: Cambridge University Press. pp. 2324. ISBN 0-521-78451-4. MR 1902334.

[8] Jech, Thomas (2008) [1973]. The Axiom of Choice. Dover Publications. ISBN 0-486-46624-8.

[9] Ward, L. E. Jr (1954). Partially Ordered Topological Spaces. Proceedings of the American Mathematical Society. 5 (1):
144161. doi:10.1090/S0002-9939-1954-0063016-5
23.16. REFERENCES 77

23.16 References
Deshpande, Jayant V. (1968). On Continuity of a Partial Order. Proceedings of the American Mathematical
Society. 19 (2): 383386. doi:10.1090/S0002-9939-1968-0236071-7.
Schmidt, Gunther (2010). Relational Mathematics. Encyclopedia of Mathematics and its Applications. 132.
Cambridge University Press. ISBN 978-0-521-76268-7.

Schrder, Bernd S. W. (2003). Ordered Sets: An Introduction. Birkhuser, Boston.


Stanley, Richard P. Enumerative Combinatorics 1. Cambridge Studies in Advanced Mathematics. 49. Cam-
bridge University Press. ISBN 0-521-66351-2.

23.17 External links


A001035: Number of posets with n labeled elements in the OEIS

A000112: Number of posets with n unlabeled elements in the OEIS


Chapter 24

Preorder

This article is about binary relations. For the graph vertex ordering, see Depth-rst search. For purchase orders for
unreleased products, see Pre-order. For other uses, see Preorder (disambiguation).
Quasiorder redirects here. For irreexive transitive relations, see strict order.

In mathematics, especially in order theory, a preorder or quasiorder is a binary relation that is reexive and transitive.
Preorders are more general than equivalence relations and (non-strict) partial orders, both of which are special cases
of a preorder.
The name 'preorder' comes from the idea that preorders (that are not partial orders) are 'almost' (partial) orders,
but not quite; they're neither necessarily anti-symmetric nor symmetric. Because a preorder is a binary relation,
the symbol can be used as the notational device for the relation. However, because they are not necessarily anti-
symmetric, some of the ordinary intuition associated to the symbol may not apply. On the other hand, a preorder
can be used, in a straightforward fashion, to dene a partial order and an equivalence relation. Doing so, however, is
not always useful or worthwhile, depending on the problem domain being studied.
In words, when a b, one may say that b covers a or that a precedes b, or that b reduces to a. Occasionally, the
notation or is used instead of .
To every preorder, there corresponds a directed graph, with elements of the set corresponding to vertices, and the
order relation between pairs of elements corresponding to the directed edges between vertices. The converse is not
true: most directed graphs are neither reexive nor transitive. In general, the corresponding graphs may contain cycles.
A preorder that is antisymmetric no longer has cycles; it is a partial order, and corresponds to a directed acyclic graph.
A preorder that is symmetric is an equivalence relation; it can be thought of as having lost the direction markers on the
edges of the graph. In general, a preorders corresponding directed graph may have many disconnected components.

24.1 Formal denition


Consider some set P and a binary relation on P. Then is a preorder, or quasiorder, if it is reexive and transitive,
i.e., for all a, b and c in P, we have that:

a a (reexivity)
if a b and b c then a c (transitivity)

A set that is equipped with a preorder is called a preordered set (or proset).[1]
If a preorder is also antisymmetric, that is, a b and b a implies a = b, then it is a partial order.
On the other hand, if it is symmetric, that is, if a b implies b a, then it is an equivalence relation.
Equivalently, the notion of a preordered set P can be formulated in a categorical framework as a thin category, i.e.
as a category with at most one morphism from an object to another. Here the objects correspond to the elements
of P, and there is one morphism for objects which are related, zero otherwise. Alternately, a preordered set can be
understood as an enriched category, enriched over the category 2 = (01).

78
24.2. EXAMPLES 79

A preordered class is a class equipped with a preorder. Every set is a class and so every preordered set is a preordered
class.

24.2 Examples
The reachability relationship in any directed graph (possibly containing cycles) gives rise to a preorder, where
x y in the preorder if and only if there is a path from x to y in the directed graph. Conversely, every preorder
is the reachability relationship of a directed graph (for instance, the graph that has an edge from x to y for every
pair (x, y) with x y). However, many dierent graphs may have the same reachability preorder as each other.
In the same way, reachability of directed acyclic graphs, directed graphs with no cycles, gives rise to partially
ordered sets (preorders satisfying an additional anti-symmetry property).
Every nite topological space gives rise to a preorder on its points by dening x y if and only if x belongs to
every neighborhood of y. Every nite preorder can be formed as the specialization preorder of a topological
space in this way. That is, there is a one-to-one correspondence between nite topologies and nite preorders.
However, the relation between innite topological spaces and their specialization preorders is not one-to-one.
A net is a directed preorder, that is, each pair of elements has an upper bound. The denition of convergence
via nets is important in topology, where preorders cannot be replaced by partially ordered sets without losing
important features.
The relation dened by x y if f (x) f (y) , where f is a function into some preorder.
The relation dened by x y if there exists some injection from x to y. Injection may be replaced by surjection,
or any type of structure-preserving function, such as ring homomorphism, or permutation.
The embedding relation for countable total orderings.
The graph-minor relation in graph theory.
A category with at most one morphism from any object x to any other object y is a preorder. Such categories
are called thin. In this sense, categories generalize preorders by allowing more than one relation between
objects: each morphism is a distinct (named) preorder relation.

In computer science, one can nd examples of the following preorders.

Many-one and Turing reductions are preorders on complexity classes.


The subtyping relations are usually preorders.
Simulation preorders are preorders (hence the name).
Reduction relations in abstract rewriting systems.
The encompassment preorder on the set of terms, dened by st if a subterm of t is a substitution instance of
s.

Example of a total preorder:

Preference, according to common models.

24.3 Uses
Preorders play a pivotal role in several situations:

Every preorder can be given a topology, the Alexandrov topology; and indeed, every preorder on a set is in
one-to-one correspondence with an Alexandrov topology on that set.
Preorders may be used to dene interior algebras.
Preorders provide the Kripke semantics for certain types of modal logic.
80 CHAPTER 24. PREORDER

24.4 Constructions
Every binary relation R on a set S can be extended to a preorder on S by taking the transitive closure and reexive
closure, R+= . The transitive closure indicates path connection in R: x R+ y if and only if there is an R-path from x to
y.
Given a preorder on S one may dene an equivalence relation ~ on S such that a ~ b if and only if a b and b
a. (The resulting relation is reexive since a preorder is reexive, transitive by applying transitivity of the preorder
twice, and symmetric by denition.)
Using this relation, it is possible to construct a partial order on the quotient set of the equivalence, S / ~, the set of
all equivalence classes of ~. Note that if the preorder is R+= , S / ~ is the set of R-cycle equivalence classes: x [y]
if and only if x = y or x is in an R-cycle with y. In any case, on S / ~ we can dene [x] [y] if and only if x y.
By the construction of ~, this denition is independent of the chosen representatives and the corresponding relation
is indeed well-dened. It is readily veried that this yields a partially ordered set.
Conversely, from a partial order on a partition of a set S one can construct a preorder on S. There is a 1-to-1 corre-
spondence between preorders and pairs (partition, partial order).
For a preorder " ", a relation "<" can be dened as a < b if and only if (a b and not b a), or equivalently, using
the equivalence relation introduced above, (a b and not a ~ b). It is a strict partial order; every strict partial order
can be the result of such a construction. If the preorder is anti-symmetric, hence a partial order "", the equivalence
is equality, so the relation "<" can also be dened as a < b if and only if (a b and a b).
(We do not dene the relation "<" as a < b if and only if (a b and a b). Doing so would cause problems if
the preorder was not anti-symmetric, as the resulting relation "<" would not be transitive (think of how equivalent
non-equal elements relate).)
Conversely we have a b if and only if a < b or a ~ b. This is the reason for using the notation " "; "" can be
confusing for a preorder that is not anti-symmetric, it may suggest that a b implies that a < b or a = b.
Note that with this construction multiple preorders " " can give the same relation "<", so without more information,
such as the equivalence relation, " " cannot be reconstructed from "<". Possible preorders include the following:

Dene a b as a < b or a = b (i.e., take the reexive closure of the relation). This gives the partial order
associated with the strict partial order "<" through reexive closure; in this case the equivalence is equality, so
we don't need the notations and ~.
Dene a b as not b < a" (i.e., take the inverse complement of the relation), which corresponds to dening a
~ b as neither a < b nor b < a"; these relations and ~ are in general not transitive; however, if they are, ~ is
an equivalence; in that case "<" is a strict weak order. The resulting preorder is total, that is, a total preorder.

24.5 Number of preorders


As explained above, there is a 1-to-1 correspondence between preorders and pairs (partition, partial order). Thus the
number of preorders is the sum of the number of partial orders on every partition. For example:

for n=3:
1 partition of 3, giving 1 preorder
3 partitions of 2+1, giving 3 3 = 9 preorders
1 partition of 1+1+1, giving 19 preorders

i.e. together 29 preorders.

for n=4:
1 partition of 4, giving 1 preorder
7 partitions with two classes (4 of 3+1 and 3 of 2+2), giving 7 3 = 21 preorders
6 partitions of 2+1+1, giving 6 19 = 114 preorders
24.6. INTERVAL 81

1 partition of 1+1+1+1, giving 219 preorders

i.e. together 355 preorders.

24.6 Interval
For a b, the interval [a,b] is the set of points x satisfying a x and x b, also written a x b. It contains at
least the points a and b. One may choose to extend the denition to all pairs (a,b). The extra intervals are all empty.
Using the corresponding strict relation "<", one can also dene the interval (a,b) as the set of points x satisfying a <
x and x < b, also written a < x < b. An open interval may be empty even if a < b.
Also [a,b) and (a,b] can be dened similarly.

24.7 See also


Partial order preorder that is antisymmetric

Equivalence relation preorder that is symmetric

Total preorder preorder that is total


Total order preorder that is antisymmetric and total

Directed set
Category of preordered sets

Prewellordering
Well-quasi-ordering

24.8 Notes
[1] For proset, see e.g. Eklund, Patrik; Ghler, Werner (1990), Generalized Cauchy spaces, Mathematische Nachrichten,
147: 219233, MR 1127325, doi:10.1002/mana.19901470123.

24.9 References
Schmidt, Gunther, Relational Mathematics, Encyclopedia of Mathematics and its Applications, vol. 132,
Cambridge University Press, 2011, ISBN 978-0-521-76268-7

Schrder, Bernd S. W. (2002), Ordered Sets: An Introduction, Boston: Birkhuser, ISBN 0-8176-4128-9
Chapter 25

Prewellordering

In set theory, a prewellordering is a binary relation that is transitive, total, and wellfounded (more precisely, the
relation x y y x is wellfounded). In other words, if is a prewellordering on a set X , and if we dene by

x y x y y x

then is an equivalence relation on X , and induces a wellordering on the quotient X/ . The order-type of this
induced wellordering is an ordinal, referred to as the length of the prewellordering.
A norm on a set X is a map from X into the ordinals. Every norm induces a prewellordering; if : X Ord is a
norm, the associated prewellordering is given by

x y (x) (y)

Conversely, every prewellordering is induced by a unique regular norm (a norm : X Ord is regular if, for any
x X and any < (x) , there is y X such that (y) = ).

25.1 Prewellordering property


If is a pointclass of subsets of some collection F of Polish spaces, F closed under Cartesian product, and if is a
prewellordering of some subset P of some element X of F , then is said to be a -prewellordering of P if the
relations < and are elements of , where for x, y X ,

1. x < y x P [y
/ P {x y y x}]

2. x y x P [y
/ P x y]

is said to have the prewellordering property if every set in admits a -prewellordering.


The prewellordering property is related to the stronger scale property; in practice, many pointclasses having the
prewellordering property also have the scale property, which allows drawing stronger conclusions.

25.1.1 Examples
11 and 12 both have the prewellordering property; this is provable in ZFC alone. Assuming sucient large cardinals,
for every n , 12n+1 and 12n+2 have the prewellordering property.

25.1.2 Consequences

82
25.2. SEE ALSO 83

Reduction

If is an adequate pointclass with the prewellordering property, then it also has the reduction property: For any
space X F and any sets A, B X , A and B both in , the union A B may be partitioned into sets A , B ,
both in , such that A A and B B .

Separation

If is an adequate pointclass whose dual pointclass has the prewellordering property, then has the separation
property: For any space X F and any sets A, B X , A and B disjoint sets both in , there is a set C X
such that both C and its complement X \ C are in , with A C and B C = .
For example, 11 has the prewellordering property, so 11 has the separation property. This means that if A and B
are disjoint analytic subsets of some Polish space X , then there is a Borel subset C of X such that C includes A and
is disjoint from B .

25.2 See also


Descriptive set theory
Scale property

Graded poset a graded poset is analogous to a prewellordering with a norm, replacing a map to the ordinals
with a map to the integers

25.3 References
Moschovakis, Yiannis N. (1980). Descriptive Set Theory. North Holland. ISBN 0-444-70199-0.
Chapter 26

Quasitransitive relation

Quasitransitivity is a weakened version of transitivity that is used in social choice theory or microeconomics. In-
formally, a relation is quasitransitive if it is symmetric for some values and transitive elsewhere. The concept was
introduced by Sen (1969) to study the consequences of Arrows theorem.

26.1 Formal denition


A binary relation T over a set X is quasitransitive if for all a, b, and c in X the following holds:

(a T b) (b T a) (b T c) (c T b) (a T c) (c T a).

If the relation is also antisymmetric, T is transitive.


Alternately, for a relation T, dene the asymmetric or strict part P:

(a P b) (a T b) (b T a).

Then T is quasitransitive i P is transitive.

26.2 Examples
Preferences are assumed to be quasitransitive (rather than transitive) in some economic contexts. The classic example
is a person indierent between 10 and 11 grams of sugar and indierent between 11 and 12 grams of sugar, but who
prefers 12 grams of sugar to 10. Similarly, the Sorites paradox can be resolved by weakening assumed transitivity of
certain relations to quasitransitivity.

26.3 Properties
Every transitive relation is quasitransitive; every quasitransitive relation is an acyclic relation. In each case the
converse does not hold in general.

26.4 See also


Intransitivity

Reexive relation

84
26.5. REFERENCES 85

26.5 References
Bossert, Walter; Suzumura, Ktar (2010). Consistency, choice and rationality. Harvard University Press.
ISBN 0674052994.
Sen, A. (1969). Quasi-transitivity, rational choice and collective decisions. Rev. Econ. Stud. 36: 381393.
Zbl 0181.47302. doi:10.2307/2296434.
Chapter 27

Quotient by an equivalence relation

This article is about a generalization to category theory, used in scheme theory. For the common meaning, see
Equivalence class.

In mathematics, given a category C, a quotient of an object X by an equivalence relation f : R X X is a


coequalizer for the pair of maps

f pr
RX X X,
i
i = 1, 2,

where R is an object in C and "f is an equivalence relation means that, for any object T in C, the image (which is
a set) of f : R(T ) = Mor(T, R) X(T ) X(T ) is an equivalence relation; that is, (x, y) is in it if and only if
(y, x) is in it, etc.
The basic case in practice is when C is the category of all schemes over some scheme S. But the notion is exible and
one can also take C to be the category of sheaves.

27.1 Examples

Let X be a set and consider some equivalence relation on it. Let Q be the set of all equivalence classes in X.
Then the map q : X Q that sends an element x to an equivalence class to which x belong is a quotient.

In the above example, Q is a subset of the power set H of X. In algebraic geometry, one might replace H
by a Hilbert scheme or disjoint union of Hilbert schemes. In fact, Grothendieck constructed a relative Picard
scheme of a at projective scheme X[1] as a quotient Q (of the scheme Z parametrizing relative eective divisors
on X) that is a closed scheme of a Hilbert scheme H. The quotient map q : Z Q can then be thought of as
a relative version of the Abel map.

27.2 See also

categorical quotient, a special case

27.3 Notes

[1] One also needs to assume the geometric bers are integral schemes; Mumfords example shows the integral cannot be
omitted.

86
27.4. REFERENCES 87

27.4 References
Nitsure, N. Construction of Hilbert and Quot schemes. Fundamental algebraic geometry: Grothendiecks FGA
explained, Mathematical Surveys and Monographs 123, American Mathematical Society 2005, 105137.
Chapter 28

Rational consequence relation

In logic, a rational consequence relation is a non-monotonic consequence relation satisfying certain properties listed
below.

28.1 Properties
A rational consequence relation satises:

REF Reexivity

and the so-called Gabbay-Makinson rules:


LLE Left Logical Equivalence
|=
RWE Right-hand weakening

CMO Cautious monotonicity

DIS Logical or (ie disjunction) on left hand side

AND Logical and on right hand side

RMO Rational monotonicity

28.2 Uses
The rational consequence relation is non-monotonic, and the relation is intended to carry the meaning theta
usually implies phi or phi usually follows from theta. In this sense it is more useful for modeling some everyday
situations than a monotone consequence relation because the latter relation models facts in a more strict boolean
fashion - something either follows under all circumstances or it does not.

28.2.1 Example
The statement If a cake contains sugar then it tastes good implies under a monotone consequence relation the state-
ment If a cake contains sugar and soap then it tastes good. Clearly this doesn't match our own understanding of
cakes. By asserting If a cake contains sugar then it usually tastes good a rational consequence relation allows for
a more realistic model of the real world, and certainly it does not automatically follow that If a cake contains sugar
and soap then it usually tastes good.
Note that if we also have the information If a cake contains sugar then it usually contains butter then we may legally
conclude (under CMO) that If a cake contains sugar and butter then it usually tastes good.. Equally in the absence

88
28.3. CONSEQUENCES 89

of a statement such as If a cake contains sugar then usually it contains no soap" then we may legally conclude from
RMO that If the cake contains sugar and soap then it usually tastes good.
If this latter conclusion seems ridiculous to you then it is likely that you are subconsciously asserting your own pre-
conceived knowledge about cakes when evaluating the validity of the statement. That is, from your experience you
know that cakes which contain soap are likely to taste bad so you add to the system your own knowledge such as
Cakes which contain sugar do not usually contain soap., even though this knowledge is absent from it. If the conclu-
sion seems silly to you then you might consider replacing the word soap with the word eggs to see if it changes your
feelings.

28.2.2 Example
Consider the sentences:

Young people are usually happy


Drug abusers are usually not happy
Drug abusers are usually young

We may consider it reasonable to conclude:

Young drug abusers are usually not happy

This would not be a valid conclusion under a monotonic deduction system (omitting of course the word 'usually'),
since the third sentence would contradict the rst two. In contrast the conclusion follows immediately using the
Gabbay-Makinson rules: applying the rule CMO to the last two sentences yields the result.

28.3 Consequences
The following consequences follow from the above rules:

()
MP Modus ponens
MP is proved via the rules AND and RWE.


CON Conditionalisation ()


CC Cautious Cut
The notion of Cautious Cut simply encapsulates the operation of conditionalisation, followed by MP.
It may seem redundant in this sense, but it is often used in proofs so it is useful to have a name for
it to act as a shortcut.

|=
SCL Supraclassity
SCL is proved trivially via REF and RWE.

28.4 Rational consequence relations via atom preferences


n
Let L = {p1 , . . . , pn } be a nite language. An atom is a formula of the form i=1 pi (where p1 = p and p1 = p
). Notice that there is a unique valuation which makes any given atom true (and conversely each valuation satises
precisely one atom). Thus an atom can be used to represent a preference about what we believe ought to be true.
Let AtL be the set of all atoms in L. For SL, dene S = { AtL | |=SC } .
Let s = s1 , . . . , sm be a sequence of subsets of AtL . For , in SL, let the relation s be such that s if one
of the following holds:
90 CHAPTER 28. RATIONAL CONSEQUENCE RELATION

1. S si = for each 1 i m

2. S si = for some 1 i m and for the least such i, S si S .

Then the relation s is a rational consequence relation. This may easily be veried by checking directly that it satises
the GM-conditions.
The idea behind the sequence of atom sets is that the earlier sets account for the most likely situations such as young
people are usually law abiding whereas the later sets account for the less likely situations such as young joyriders
are usually not law abiding.

28.4.1 Notes
1. By the denition of the relation s , the relation is unchanged if we replace s2 with s2 \ s1 , s3 with s3 \ s2 \ s1
m1
... and sm with sm \ i=1 si . In this way we make each si disjoint. Conversely it makes no dierence to the
rcr s if we add to subsequent si atoms from any of the preceding si .

28.5 The representation theorem


It can be proven that any rational consequence relation on a nite language is representable via a sequence of atom
preferences above. That is, for any such rational consequence relation there is a sequence s = s1 , . . . , sm of subsets
of AtL such that the associated rcr s is the same relation: s =

28.5.1 Notes
1. By the above property of s , the representation of an rcr need not be unique - if the si are not disjoint then
they can be made so without changing the rcr and conversely if they are disjoint then each subsequent set can
contain any of the atoms of the previous sets without changing the rcr.

28.6 References
A mathematical paper in which the GM rules are dened
Chapter 29

Reexive closure

In mathematics, the reexive closure of a binary relation R on a set X is the smallest reexive relation on X that
contains R.
For example, if X is a set of distinct numbers and x R y means "x is less than y", then the reexive closure of R is the
relation "x is less than or equal to y".

29.1 Denition
The reexive closure S of a relation R on a set X is given by

S = R {(x, x) : x X}
In words, the reexive closure of R is the union of R with the identity relation on X.

29.2 Example
As an example, if

X = {1, 2, 3, 4}
R = {(1, 1), (2, 2), (3, 3), (4, 4)}
then the relation R is already reexive by itself, so it doesn't dier from its reexive closure.
However, if any of the pairs in R was absent, it would be inserted for the reexive closure. For example, if

X = {1, 2, 3, 4}
R = {(1, 1), (2, 2), (4, 4)}
then reexive closure is, by the denition of a reexive closure:

S = R {(x, x) : x X} = {(1, 1), (2, 2), (3, 3), (4, 4)}

29.3 See also


Transitive closure
Symmetric closure

91
92 CHAPTER 29. REFLEXIVE CLOSURE

29.4 References
Franz Baader and Tobias Nipkow, Term Rewriting and All That, Cambridge University Press, 1998, p. 8
Chapter 30

Reexive relation

In mathematics, a binary relation R over a set X is reexive if every element of X is related to itself.[1][2] Formally,
this may be written x X : x R x.
An example of a reexive relation is the relation "is equal to" on the set of real numbers, since every real number is
equal to itself. A reexive relation is said to have the reexive property or is said to possess reexivity. Along with
symmetry and transitivity, reexivity is one of three properties dening equivalence relations.

30.1 Related terms

A relation that is irreexive, or anti-reexive, is a binary relation on a set where no element is related to itself. An
example is the greater than relation (x > y) on the real numbers. Note that not every relation which is not reexive
is irreexive; it is possible to dene relations where some elements are related to themselves but others are not (i.e.,
neither all nor none are). For example, the binary relation the product of x and y is even is reexive on the set of
even numbers, irreexive on the set of odd numbers, and neither reexive nor irreexive on the set of natural numbers.
A relation ~ on a set S is called quasi-reexive if every element that is related to some element is also related to
itself, formally: x, y S : x ~ y (x ~ x y ~ y). An example is the relation has the same limit as on the set of
sequences of real numbers: not every sequence has a limit, and thus the relation is not reexive, but if a sequence has
the same limit as some sequence, then it has the same limit as itself.
The reexive closure of a binary relation ~ on a set S is the smallest reexive relation on S that is a superset of ~.
Equivalently, it is the union of ~ and the identity relation on S, formally: () = (~) (=). For example, the reexive
closure of x < y is x y.
The reexive reduction, or irreexive kernel, of a binary relation ~ on a set S is the smallest relation such that
shares the same reexive closure as ~. It can be seen in a way as the opposite of the reexive closure. It is equivalent
to the complement of the identity relation on S with regard to ~, formally: () = (~) \ (=). That is, it is equivalent to
~ except for where x~x is true. For example, the reexive reduction of x y is x < y.

93
94 CHAPTER 30. REFLEXIVE RELATION

30.2 Examples

Examples of reexive relations include:

is equal to (equality)
is a subset of (set inclusion)
divides (divisibility)
is greater than or equal to
is less than or equal to

Examples of irreexive relations include:

is not equal to
is coprime to (for the integers>1, since 1 is coprime to itself)
is a proper subset of
is greater than
is less than

30.3 Number of reexive relations


2
The number of reexive relations on an n-element set is 2n n [3]
.

30.4 Philosophical logic


Authors in philosophical logic often use dierent terminology. Reexive relations in the mathematical sense are
called totally reexive in philosophical logic, and quasi-reexive relations are called reexive.[4][5]

30.5 See also


Antisymmetric relation
30.6. NOTES 95

30.6 Notes
[1] Levy 1979:74

[2] Relational Mathematics, 2010

[3] On-Line Encyclopedia of Integer Sequences A053763

[4] Alan Hausman; Howard Kahane; Paul Tidman (2013). Logic and Philosophy A Modern Introduction. Wadsworth. ISBN
1-133-05000-X. Here: p.327-328

[5] D.S. Clarke; Richard Behling (1998). Deductive Logic An Introduction to Evaluation Techniques and Logical Theory.
University Press of America. ISBN 0-7618-0922-8. Here: p.187

30.7 References
Levy, A. (1979) Basic Set Theory, Perspectives in Mathematical Logic, Springer-Verlag. Reprinted 2002,
Dover. ISBN 0-486-42079-5

Lidl, R. and Pilz, G. (1998). Applied abstract algebra, Undergraduate Texts in Mathematics, Springer-Verlag.
ISBN 0-387-98290-6

Quine, W. V. (1951). Mathematical Logic, Revised Edition. Reprinted 2003, Harvard University Press. ISBN
0-674-55451-5

Gunther Schmidt, 2010. Relational Mathematics. Cambridge University Press, ISBN 978-0-521-76268-7.

30.8 External links


Hazewinkel, Michiel, ed. (2001) [1994], Reexivity, Encyclopedia of Mathematics, Springer Science+Business
Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Chapter 31

Semiorder

In order theory, a branch of mathematics, a semiorder is a type of ordering that may be determined for a set of
items with numerical scores by declaring two items to be incomparable when their scores are within a given margin
of error of each other, and by using the numerical comparison of their scores when those scores are suciently far
apart. Semiorders were introduced and applied in mathematical psychology by Luce (1956) as a model of human
preference without the assumption that indierence is transitive. They generalize strict weak orderings, form a special
case of partial orders and interval orders, and can be characterized among the partial orders by two forbidden four-
item suborders.

31.1 Denition
Let X be a set of items, and let < be a binary relation on X. Items x and y are said to be incomparable, written here as
x ~ y, if neither x < y nor y < x is true. Then the pair (X,<) is a semiorder if it satises the following three axioms:[1]

For all x and y, it is not possible for both x < y and y < x to be true. That is, < must be an irreexive,
antisymmetric relation

For all x, y, z, and w, if it is true that x < y, y ~ z, and z < w, then it must also be true that x < w.

For all x, y, z, and w, if it is true that x < y, y < z, and y ~ w, then it cannot also be true that x ~ w and z ~ w
simultaneously.

It follows from the rst axiom that x ~ x, and therefore the second axiom (with y = z) implies that < is a transitive
relation.
One may dene a partial order (X,) from a semiorder (X,<) by declaring that x y whenever either x < y or x = y. Of
the axioms that a partial order is required to obey, reexivity follows automatically from this denition, antisymmetry
follows from the rst semiorder axiom, and transitivity follows from the second semiorder axiom. Conversely, from
a partial order dened in this way, the semiorder may be recovered by declaring that x < y whenever x y and x
y. The rst of the semiorder axioms listed above follows automatically from the axioms dening a partial order, but
the others do not. The second and third semiorder axioms forbid partial orders of four items forming two disjoint
chains: the second axiom forbids two chains of two items each, while the third item forbids a three-item chain with
one unrelated item.

31.2 Utility
The original motivation for introducing semiorders was to model human preferences without assuming (as strict weak
orderings do) that incomparability is a transitive relation. For instance, if x, y, and z represent three quantities of the
same material, and x and z dier by the smallest amount that is perceptible as a dierence, while y is halfway between
the two of them, then it is reasonable for a preference to exist between x and z but not between the other two pairs,
violating transitivity.[2]

96
31.2. UTILITY 97

An example of a semiorder, shown by its Hasse diagram. The horizontal blue lines indicate the spacing of the y-coordinates of the
points; two points are comparable when their y coordinates dier by at least one unit.

Thus, suppose that X is a set of items, and u is a utility function that maps the members of X to real numbers. A strict
weak ordering can be dened on x by declaring two items to be incomparable when they have equal utilities, and
otherwise using the numerical comparison, but this necessarily leads to a transitive incomparability relation. Instead,
if one sets a numerical threshold (which may be normalized to 1) such that utilities within that threshold of each other
are declared incomparable, then a semiorder arises.
Specically, dene a binary relation < from X and u by setting x < y whenever u(x) u(y) 1. Then (X,<) is a
semiorder.[3] It may equivalently be dened as the interval order dened by the intervals [u(x),u(x) + 1].[4]
In the other direction, not every semiorder can be dened from numerical utilities in this way. For instance, if a
semiorder (X,<) includes an uncountable totally ordered subset then there do not exist suciently many suciently
well-spaced real-numbers to represent this subset numerically. However, every nite semiorder can be dened from
a utility function.[5] Fishburn (1973) supplies a precise characterization of the semiorders that may be dened nu-
98 CHAPTER 31. SEMIORDER

merically.

31.3 Other results


The number of distinct semiorders on n unlabeled items is given by the Catalan numbers

2n
( )
1
n+1 n , [6]

while the number of semiorders on n labeled items is given by the sequence

1, 1, 3, 19, 183, 2371, 38703, 763099, 17648823, ... (sequence A006531 in the OEIS).[7]

Any nite semiorder has order dimension at most three.[8]


Among all partial orders with a xed number of elements and a xed number of comparable pairs, the partial orders
that have the largest number of linear extensions are semiorders.[9]
Semiorders are known to obey the 1/32/3 conjecture: in any nite semiorder that is not a total order, there exists
a pair of elements x and y such that x appears earlier than y in between 1/3 and 2/3 of the linear extensions of the
semiorder.[10]
The set of semiorders on an n-element set is well-graded: if two semiorders on the same set dier from each other
by the addition or removal of k order relations, then it is possible to nd a path of k steps from the rst semiorder to
the second one, in such a way that each step of the path adds or removes a single order relation and each intermediate
state in the path is itself a semiorder.[11]
The incomparability graphs of semiorders are called indierence graphs, and are a special case of the interval
graphs.[12]
If a semiorder is given only in terms of the order relation between its pairs of elements, then it is possible to construct
a utility function that represents the order in time O(n2 ), where n is the number of elements in the semiorder.[13]

31.4 Notes
[1] Luce (1956) describes an equivalent set of four axioms, the rst two of which combine the denition of incomparability
and the rst axiom listed here.

[2] Luce (1956), p. 179.

[3] Luce (1956), Theorem 3 describes a more general situation in which the threshold for comparability between two utilities
is a function of the utility rather than being identically 1.

[4] Fishburn (1970).

[5] This result is typically credited to Scott & Suppes (1958); see, e.g., Rabinovitch (1977). However, Luce (1956), Theorem
2 proves a more general statement, that a nite semiorder can be dened from a utility function and a threshold function
whenever a certain underlying weak order can be dened numerically. For nite semiorders, it is trivial that the weak order
can be dened numerically with a unit threshold function.

[6] Kim & Roush (1978).

[7] Chandon, Lemaire & Pouget (1978).

[8] Rabinovitch (1978).

[9] Fishburn & Trotter (1992).

[10] Brightwell (1989).

[11] Doignon & Falmagne (1997).

[12] Roberts (1969).

[13] Avery (1992).


31.5. REFERENCES 99

31.5 References
Avery, Peter (1992), An algorithmic proof that semiorders are representable, Journal of Algorithms, 13 (1):
144147, MR 1146337, doi:10.1016/0196-6774(92)90010-A.
Brightwell, Graham R. (1989), Semiorders and the 1/32/3 conjecture, Order, 5 (4): 369380, doi:10.1007/BF00353656.

Chandon, J.-L.; Lemaire, J.; Pouget, J. (1978), Dnombrement des quasi-ordres sur un ensemble ni, Centre
de Mathmatique Sociale. cole Pratique des Hautes tudes. Mathmatiques et Sciences Humaines (62): 6180,
83, MR 517680.

Doignon, Jean-Paul; Falmagne, Jean-Claude (1997), Well-graded families of relations, Discrete Mathematics,
173 (1-3): 3544, MR 1468838, doi:10.1016/S0012-365X(96)00095-7.

Fishburn, Peter C. (1970), Intransitive indierence with unequal indierence intervals, J. Mathematical
Psychology, 7: 144149, MR 0253942, doi:10.1016/0022-2496(70)90062-3.

Fishburn, Peter C. (1973), Interval representations for interval orders and semiorders, J. Mathematical Psy-
chology, 10: 91105, MR 0316322, doi:10.1016/0022-2496(73)90007-2.

Fishburn, Peter C.; Trotter, W. T. (1992), Linear extensions of semiorders: a maximization problem, Discrete
Mathematics, 103 (1): 2540, MR 1171114, doi:10.1016/0012-365X(92)90036-F.

Kim, K. H.; Roush, F. W. (1978), Enumeration of isomorphism classes of semiorders, Journal of Combina-
torics, Information &System Sciences, 3 (2): 5861, MR 538212.

Luce, R. Duncan (1956), Semiorders and a theory of utility discrimination, Econometrica, 24: 178191,
JSTOR 1905751, MR 0078632, doi:10.2307/1905751.

Rabinovitch, Issie (1977), The Scott-Suppes theorem on semiorders, J. Mathematical Psychology, 15 (2):
209212, MR 0437404, doi:10.1016/0022-2496(77)90030-x.

Rabinovitch, Issie (1978), The dimension of semiorders, Journal of Combinatorial Theory. Series A, 25 (1):
5061, MR 0498294, doi:10.1016/0097-3165(78)90030-4.

Roberts, Fred S. (1969), Indierence graphs, Proof Techniques in Graph Theory (Proc. Second Ann Arbor
Graph Theory Conf., Ann Arbor, Mich., 1968), Academic Press, New York, pp. 139146, MR 0252267.

Scott, Dana; Suppes, Patrick (1958), Foundational aspects of theories of measurement, The Journal of Sym-
bolic Logic, 23: 113128, MR 0115919, doi:10.2307/2964389.

31.6 Additional reading


Pirlot, M.; Vincke, Ph. (1997), Semiorders: Properties, representations, applications, Theory and Decision
Library. Series B: Mathematical and Statistical Methods, 36, Dordrecht: Kluwer Academic Publishers Group,
ISBN 0-7923-4617-3, MR 1472236.
Chapter 32

Separoid

In mathematics, a separoid is a binary relation between disjoint sets which is stable as an ideal in the canonical order
induced by inclusion. Many mathematical objects which appear to be quite dierent, nd a common generalisation in
the framework of separoids; e.g., graphs, congurations of convex sets, oriented matroids, and polytopes. Any count-
able category is an induced subcategory of separoids when they are endowed with homomorphisms (viz., mappings
that preserve the so-called minimal Radon partitions).
In this general framework, some results and invariants of dierent categories turn out to be special cases of the
same aspect; e.g., the pseudoachromatic number from graph theory and the Tverberg theorem from combinatorial
convexity are simply two faces of the same aspect, namely, complete colouring of separoids.

32.1 The axioms


A separoid is a set S endowed with a binary relation | 2S 2S on its power set, which satises the following
simple properties for A, B S :

A | B B | A,

A | B A B = ,

A | B and A A A | B.
A related pair A | B is called a separation and we often say that A is separated from B. It is enough to know the
maximal separations to reconstruct the separoid.
A mapping : S T is a morphism of separoids if the preimages of separations are separations; that is, for
A, B T

A | B 1 (A) | 1 (B).

32.2 Examples
Examples of separoids can be found in almost every branch of mathematics. Here we list just a few.
1. Given a graph G=(V,E), we can dene a separoid on its vertices by saying that two (disjoint) subsets of V, say A
and B, are separated if there are no edges going from one to the other; i.e.,

A | B a A and b B : ab E.

100
32.3. THE BASIC LEMMA 101

2. Given an oriented matroid M = (E,T), given in terms of its topes T, we can dene a separoid on E by saying that
two subsets are separated if they are contained in opposite signs of a tope. In other words, the topes of an oriented
matroid are the maximal separations of a separoid. This example includes, of course, all directed graphs.
3. Given a family of objects in an Euclidean space, we can dene a separoid in it by saying that two subsets are
separated if there exists a hyperplane that separates them; i.e., leaving them in the two opposite sides of it.
4. Given a topological space, we can dene a separoid saying that two subsets are separated if there exist two disjoint
open sets which contains them (one for each of them).

32.3 The basic lemma


Every separoid can be represented with a family of convex sets in some Euclidean space and their separations by
hyperplanes.

32.4 References
Strausz Ricardo; Separoides. Situs, serie B, no. 5 (1998), Universidad Nacional Autnoma de Mxico.

Arocha Jorge Luis, Bracho Javier, Montejano Luis, Oliveros Deborah, Strausz Ricardo; Separoids, their cat-
egories and a Hadwiger-type theorem for transversals. Discrete and Computational Geometry 27 (2002), no.
3, 377385.
Strausz Ricardo; Separoids and a Tverberg-type problem. Geombinatorics 15 (2005), no. 2, 7992.

Montellano-Ballesteros Juan Jose, Por Attila, Strausz Ricardo; Tverberg-type theorems for separoids. Dis-
crete and Computational Geometry 35 (2006), no.3, 513523.

Neetil Jaroslav, Strausz Ricardo; Universality of separoids. Archivum Mathematicum (Brno) 42 (2006), no.
1, 85101.

Bracho Javier, Strausz Ricardo; Two geometric representations of separoids. Periodica Mathematica Hun-
garica 53 (2006), no. 1-2, 115120.

Strausz Ricardo; Homomorphisms of separoids. 6th Czech-Slovak International Symposium on Combina-


torics, Graph Theory, Algorithms and Applications, 461468, Electronic Notes on Discrete Mathematics 28,
Elsevier, Amsterdam, 2007.
Strausz Ricardo; Edrs-Szekeres 'happy end'-type theorems for separoids. European Journal of Combina-
torics 29 (2008), no. 4, 10761085.
Chapter 33

Series-parallel partial order

Series composition

Parallel composition

A series-parallel partial order, shown as a Hasse diagram.

In order-theoretic mathematics, a series-parallel partial order is a partially ordered set built up from smaller series-

102
33.1. DEFINITION 103

parallel partial orders by two simple composition operations.[1][2]


The series-parallel partial orders may be characterized as the N-free nite partial orders; they have order dimension at
most two.[1][3] They include weak orders and the reachability relationship in directed trees and directed series-parallel
graphs.[2][3] The comparability graphs of series-parallel partial orders are cographs.[2][4]
Series-parallel partial orders have been applied in job shop scheduling,[5] machine learning of event sequencing in
time series data,[6] transmission sequencing of multimedia data,[7] and throughput maximization in dataow pro-
gramming.[8]
Series-parallel partial orders have also been called multitrees;[4] however, that name is ambiguous: multitrees also
refer to partial orders with no four-element diamond suborder[9] and to other structures formed from multiple trees.

33.1 Denition
Consider P and Q, two partially ordered sets. The series composition of P and Q, written P; Q,[7] P * Q,[2] or P
Q,[1] is the partially ordered set whose elements are the disjoint union of the elements of P and Q. In P; Q, two elements
x and y that both belong to P or that both belong to Q have the same order relation that they do in P or Q respectively.
However, for every pair x, y where x belongs to P and y belongs to Q, there is an additional order relation x y in the
series composition. Series composition is an associative operation: one can write P; Q; R as the series composition
of three orders, without ambiguity about how to combine them pairwise, because both of the parenthesizations (P;
Q); R and P; (Q; R) describe the same partial order. However, it is not a commutative operation, because switching
the roles of P and Q will produce a dierent partial order that reverses the order relations of pairs with one element
in P and one in Q.[1]
The parallel composition of P and Q, written P || Q,[7] P + Q,[2] or P Q,[1] is dened similarly, from the disjoint
union of the elements in P and the elements in Q, with pairs of elements that both belong to P or both to Q having
the same order as they do in P or Q respectively. In P || Q, a pair x, y is incomparable whenever x belongs to P and y
belongs to Q. Parallel composition is both commutative and associative.[1]
The class of series-parallel partial orders is the set of partial orders that can be built up from single-element partial
orders using these two operations. Equivalently, it is the smallest set of partial orders that includes the single-element
partial order and is closed under the series and parallel composition operations.[1][2]
A weak order is the series parallel partial order obtained from a sequence of composition operations in which all
of the parallel compositions are performed rst, and then the results of these compositions are combined using only
series compositions.[2]

33.2 Forbidden suborder characterization


The partial order N with the four elements a, b, c, and d and exactly the three order relations a b c d is an
example of a fence or zigzag poset; its Hasse diagram has the shape of the capital letter N. It is not series-parallel,
because there is no way of splitting it into the series or parallel composition of two smaller partial orders. A partial
order P is said to be N-free if there does not exist a set of four elements in P such that the restriction of P to those
elements is order-isomorphic to N. The series-parallel partial orders are exactly the nonempty nite N-free partial
orders.[1][2][3]
It follows immediately from this (although it can also be proven directly) that any nonempty restriction of a series-
parallel partial order is itself a series-parallel partial order.[1]

33.3 Order dimension


The order dimension of a partial order P is the minimum size of a realizer of P, a set of linear extensions of P with
the property that, for every two distinct elements x and y of P, x y in P if and only if x has an earlier position than
y in every linear extension of the realizer. Series-parallel partial orders have order dimension at most two. If P and
Q have realizers {L1 , L2 } and {L3 , L4 }, respectively, then {L1 L3 , L2 L4 } is a realizer of the series composition P; Q,
and {L1 L3 , L4 L2 } is a realizer of the parallel composition P || Q.[2][3] A partial order is series-parallel if and only if
it has a realizer in which one of the two permutations is the identity and the other is a separable permutation.
104 CHAPTER 33. SERIES-PARALLEL PARTIAL ORDER

It is known that a partial order P has order dimension two if and only if there exists a conjugate order Q on the
same elements, with the property that any two distinct elements x and y are comparable on exactly one of these two
orders. In the case of series parallel partial orders, a conjugate order that is itself series parallel may be obtained by
performing a sequence of composition operations in the same order as the ones dening P on the same elements,
but performing a series composition for each parallel composition in the decomposition of P and vice versa. More
strongly, although a partial order may have many dierent conjugates, every conjugate of a series parallel partial order
must itself be series parallel.[2]

33.4 Connections to graph theory


Any partial order may be represented (usually in more than one way) by a directed acyclic graph in which there is a
path from x to y whenever x and y are elements of the partial order with x y. The graphs that represent series-parallel
partial orders in this way have been called vertex series parallel graphs, and their transitive reductions (the graphs
of the covering relations of the partial order) are called minimal vertex series parallel graphs.[3] Directed trees and
(two-terminal) series parallel graphs are examples of minimal vertex series parallel graphs; therefore, series parallel
partial orders may be used to represent reachability relations in directed trees and series parallel graphs.[2][3]
The comparability graph of a partial order is the undirected graph with a vertex for each element and an undirected
edge for each pair of distinct elements x, y with either x y or y x. That is, it is formed from a minimal vertex
series parallel graph by forgetting the orientation of each edge. The comparability graph of a series-partial order is a
cograph: the series and parallel composition operations of the partial order give rise to operations on the comparability
graph that form the disjoint union of two subgraphs or that connect two subgraphs by all possible edges; these two
operations are the basic operations from which cographs are dened. Conversely, every cograph is the comparability
graph of a series-parallel partial order. If a partial order has a cograph as its comparability graph, then it must be a
series-parallel partial order, because every other kind of partial order has an N suborder that would correspond to an
induced four-vertex path in its comparability graph, and such paths are forbidden in cographs.[2][4]

33.5 Computational complexity


It is possible to use the forbidden suborder characterization of series-parallel partial orders as a basis for an algorithm
that tests whether a given binary relation is a series-parallel partial order, in an amount of time that is linear in
the number of related pairs.[2][3] Alternatively, if a partial order is described as the reachability order of a directed
acyclic graph, it is possible to test whether it is a series-parallel partial order, and if so compute its transitive closure,
in time proportional to the number of vertices and edges in the transitive closure; it remains open whether the time
to recognize series-parallel reachability orders can be improved to be linear in the size of the input graph.[10]
If a series-parallel partial order is represented as an expression tree describing the series and parallel composition
operations that formed it, then the elements of the partial order may be represented by the leaves of the expression
tree. A comparison between any two elements may be performed algorithmically by searching for the lowest common
ancestor of the corresponding two leaves; if that ancestor is a parallel composition, the two elements are incomparable,
and otherwise the order of the series composition operands determines the order of the elements. In this way, a series-
parallel partial order on n elements may be represented in O(n) space with O(1) time to determine any comparison
value.[2]
It is NP-complete to test, for two given series-parallel partial orders P and Q, whether P contains a restriction iso-
morphic to Q.[3]
Although the problem of counting the number of linear extensions of an arbitrary partial order is #P-complete,[11] it
may be solved in polynomial time for series-parallel partial orders. Specically, if L(P) denotes the number of linear
extensions of a partial order P, then L(P; Q) = L(P)L(Q) and

(|P | + |Q|)!
L(P ||Q) = L(P )L(Q),
|P |!|Q|!

so the number of linear extensions may be calculated using an expression tree with the same form as the decomposition
tree of the given series-parallel order.[2]
33.6. APPLICATIONS 105

33.6 Applications
Mannila & Meek (2000) use series-parallel partial orders as a model for the sequences of events in time series data.
They describe machine learning algorithms for inferring models of this type, and demonstrate its eectiveness at
inferring course prerequisites from student enrollment data and at modeling web browser usage patterns.[6]
Amer et al. (1994) argue that series-parallel partial orders are a good t for modeling the transmission sequencing
requirements of multimedia presentations. They use the formula for computing the number of linear extensions of a
series-parallel partial order as the basis for analyzing multimedia transmission algorithms.[7]
Choudhary et al. (1994) use series-parallel partial orders to model the task dependencies in a dataow model of
massive data processing for computer vision. They show that, by using series-parallel orders for this problem, it is
possible to eciently construct an optimized schedule that assigns dierent tasks to dierent processors of a parallel
computing system in order to optimize the throughput of the system.[8]
A class of orderings somewhat more general than series-parallel partial orders is provided by PQ trees, data structures
that have been applied in algorithms for testing whether a graph is planar and recognizing interval graphs.[12] A P
node of a PQ tree allows all possible orderings of its children, like a parallel composition of partial orders, while a Q
node requires the children to occur in a xed linear ordering, like a series composition of partial orders. However,
unlike series-parallel partial orders, PQ trees allow the linear ordering of any Q node to be reversed.

33.7 See also


Series and parallel circuits

33.8 References
[1] Bechet, Denis; De Groote, Philippe; Retor, Christian (1997), A complete axiomatisation for the inclusion of series-
parallel partial orders, Rewriting Techniques and Applications, Lecture Notes in Computer Science, 1232, Springer-Verlag,
pp. 230240, doi:10.1007/3-540-62950-5_74.
[2] Mhring, Rolf H. (1989), Computationally tractable classes of ordered sets, in Rival, Ivan, Algorithms and Order: Pro-
ceedings of the NATO Advanced Study Institute on Algorithms and Order, Ottawa, Canada, May 31-June 13, 1987, NATO
Science Series C, 255, Springer-Verlag, pp. 105194, ISBN 978-0-7923-0007-6.
[3] Valdes, Jacobo; Tarjan, Robert E.; Lawler, Eugene L. (1982), The recognition of series parallel digraphs, SIAM Journal
on Computing, 11 (2): 298313, doi:10.1137/0211023.
[4] Jung, H. A. (1978), On a class of posets and the corresponding comparability graphs, Journal of Combinatorial Theory,
Series B, 24 (2): 125133, MR 0491356, doi:10.1016/0095-8956(78)90013-8.
[5] Lawler, Eugene L. (1978), Sequencing jobs to minimize total weighted completion time subject to precedence constraints,
Annals of Discrete Mathematics, 2: 7590, MR 0495156, doi:10.1016/S0167-5060(08)70323-6.
[6] Mannila, Heikki; Meek, Christopher (2000), Global partial orders from sequential data, Proc. 6th ACM SIGKDD Inter-
national Conference on Knowledge Discovery and Data Mining (KDD 2000), pp. 161168, doi:10.1145/347090.347122.
[7] Amer, Paul D.; Chassot, Christophe; Connolly, Thomas J.; Diaz, Michel; Conrad, Phillip (1994), Partial-order transport
service for multimedia and other applications, IEEE/ACM Transactions on Networking, 2 (5): 440456, doi:10.1109/90.336326.
[8] Choudhary, A. N.; Narahari, B.; Nicol, D. M.; Simha, R. (1994), Optimal processor assignment for a class of pipelined
computations, IEEE Transactions on Parallel and Distributed Systems, 5 (4): 439445, doi:10.1109/71.273050.
[9] Furnas, George W.; Zacks, Je (1994), Multitrees: enriching and reusing hierarchical structure, Proc. SIGCHI conference
on Human Factors in Computing Systems (CHI '94), pp. 330336, doi:10.1145/191666.191778.
[10] Ma, Tze-Heng; Spinrad, Jeremy (1991), Transitive closure for restricted classes of partial orders, Order, 8 (2): 175183,
doi:10.1007/BF00383402.
[11] Brightwell, Graham R.; Winkler, Peter (1991), Counting linear extensions, Order, 8 (3): 225242, doi:10.1007/BF00383444.
[12] Booth, Kellogg S.; Lueker, George S. (1976), Testing for the consecutive ones property, interval graphs, and graph
planarity using PQ-tree algorithms, Journal of Computer and System Sciences, 13 (3): 335379, doi:10.1016/S0022-
0000(76)80045-1.
Chapter 34

Symmetric closure

In mathematics, the symmetric closure of a binary relation R on a set X is the smallest symmetric relation on X that
contains R.
For example, if X is a set of airports and xRy means there is a direct ight from airport x to airport y", then the
symmetric closure of R is the relation there is a direct ight either from x to y or from y to x". Or, if X is the set of
humans and R is the relation 'parent of', then the symmetric closure of R is the relation "x is a parent or a child of y".

34.1 Denition
The symmetric closure S of a relation R on a set X is given by

S = R {(x, y) : (y, x) R} .

In other words, the symmetric closure of R is the union of R with its inverse relation, R 1 .

34.2 See also


Transitive closure
Reexive closure

34.3 References
Franz Baader and Tobias Nipkow, Term Rewriting and All That, Cambridge University Press, 1998, p. 8

106
Chapter 35

Tolerance relation

In mathematics, a tolerance relation is a relation that is reexive and symmetric. It does not need to be transitive.

35.1 External links


Gerasin, S. N., Shlyakhov, V. V., and Yakovlev, S. V. 2008. Set coverings and tolerance relations. Cybernetics
and Sys. Anal. 44, 3 (May 2008), 333340. doi:10.1007/s10559-008-9007-y
Hryniewiecki, K. 1991, Relations of Tolerance

FORMALIZED MATHEMATICS, Vol. 2, No. 1, JanuaryFebruary 1991.

107
Chapter 36

Total order

In mathematics, a linear order, total order, simple order, or (non-strict) ordering is a binary relation on some set
X , which is antisymmetric, transitive, and total (this relation is denoted here by inx ). A set paired with a total
order is called a totally ordered set, a linearly ordered set, a simply ordered set, or a chain.
More symbolically, a set X is totally ordered under if the following statements hold for all a, b and c in X :

If a b and b a then a = b (antisymmetry);


If a b and b c then a c (transitivity);
a b or b a (totality).

Antisymmetry eliminates uncertain cases when both a precedes b and b precedes a .[1] A relation having the property
of totality means that any pair of elements in the set of the relation are comparable under the relation. This also
means that the set can be diagrammed as a line of elements, giving it the name linear.[2] Totality also implies reexivity,
i.e., a a. Therefore, a total order is also a partial order, as, for a partial order, the totality condition is replaced by
the weaker condition of reexivity. An extension of a given partial order to a total order is called a linear extension
of that partial order.

36.1 Strict total order


For each (non-strict) total order there is an associated asymmetric (hence irreexive) relation <, called a strict total
order, which can equivalently be dened in two ways:

a < b if and only if a b and a b


a < b if and only if not b a (i.e., < is the inverse of the complement of )

Properties:

The relation is transitive: a < b and b < c implies a < c.


The relation is trichotomous: exactly one of a < b, b < a and a = b is true.
The relation is a strict weak order, where the associated equivalence is equality.

We can work the other way and start by choosing < as a transitive trichotomous binary relation; then a total order
can equivalently be dened in two ways:

a b if and only if a < b or a = b


a b if and only if not b < a

Two more associated orders are the complements and >, completing the quadruple {<, >, , }.
We can dene or explain the way a set is totally ordered by any of these four relations; the notation implies whether
we are talking about the non-strict or the strict total order.

108
36.2. EXAMPLES 109

36.2 Examples

The letters of the alphabet ordered by the standard dictionary order, e.g., A < B < C etc.

Any subset of a totally ordered set X is totally ordered for the restriction of the order on X.

The unique order on the empty set, , is a total order.

Any set of cardinal numbers or ordinal numbers (more strongly, these are well-orders).

If X is any set and f an injective function from X to a totally ordered set then f induces a total ordering on X
by setting x1 < x2 if and only if f(x1 ) < f(x2 ).

The lexicographical order on the Cartesian product of a family of totally ordered sets, indexed by a well ordered
set, is itself a total order.

The set of real numbers ordered by the usual less than (<) or greater than (>) relations is totally ordered, hence
also the subsets of natural numbers, integers, and rational numbers. Each of these can be shown to be the unique
(to within isomorphism) smallest example of a totally ordered set with a certain property, (a total order A is
the smallest with a certain property if whenever B has the property, there is an order isomorphism from A to a
subset of B):

The natural numbers comprise the smallest totally ordered set with no upper bound.
The integers comprise the smallest totally ordered set with neither an upper nor a lower bound.
The rational numbers comprise the smallest totally ordered set which is dense in the real numbers. The
denition of density used here says that for every a and b in the real numbers such that a < b, there is a q
in the rational numbers such that a < q < b.
The real numbers comprise the smallest unbounded totally ordered set that is connected in the order
topology (dened below).

Ordered elds are totally ordered by denition. They include the rational numbers and the real numbers. Every
ordered eld contains an ordered subeld that is isomorphic to the rational numbers. Any Dedekind-complete
ordered eld is isomorphic to the real numbers.

36.3 Further concepts

36.3.1 Chains

While chain is sometimes merely a synonym for totally ordered set, it can also refer to a totally ordered subset of
some partially ordered set. The latter denition has a crucial role in Zorns lemma. The height of a poset denotes the
cardinality of its largest chain in this sense.
For example, consider the set of all subsets of the integers partially ordered by inclusion. Then the set { In : n is a
natural number}, where In is the set of natural numbers below n, is a chain in this ordering, as it is totally ordered
under inclusion: If nk, then In is a subset of Ik.

36.3.2 Lattice theory

One may dene a totally ordered set as a particular kind of lattice, namely one in which we have

{a b, a b} = {a, b} for all a, b.

We then write a b if and only if a = a b . Hence a totally ordered set is a distributive lattice.
110 CHAPTER 36. TOTAL ORDER

36.3.3 Finite total orders


A simple counting argument will verify that any non-empty nite totally ordered set (and hence any non-empty subset
thereof) has a least element. Thus every nite total order is in fact a well order. Either by direct proof or by observing
that every well order is order isomorphic to an ordinal one may show that every nite total order is order isomorphic
to an initial segment of the natural numbers ordered by <. In other words, a total order on a set with k elements
induces a bijection with the rst k natural numbers. Hence it is common to index nite total orders or well orders
with order type by natural numbers in a fashion which respects the ordering (either starting with zero or with one).

36.3.4 Category theory


Totally ordered sets form a full subcategory of the category of partially ordered sets, with the morphisms being maps
which respect the orders, i.e. maps f such that if a b then f(a) f(b).
A bijective map between two totally ordered sets that respects the two orders is an isomorphism in this category.

36.3.5 Order topology


For any totally ordered set X we can dene the open intervals (a, b) = {x : a < x and x < b}, (, b) = {x : x < b},
(a, ) = {x : a < x} and (, ) = X. We can use these open intervals to dene a topology on any ordered set, the
order topology.
When more than one order is being used on a set one talks about the order topology induced by a particular order.
For instance if N is the natural numbers, < is less than and > greater than we might refer to the order topology on
N induced by < and the order topology on N induced by > (in this case they happen to be identical but will not in
general).
The order topology induced by a total order may be shown to be hereditarily normal.

36.3.6 Completeness
A totally ordered set is said to be complete if every nonempty subset that has an upper bound, has a least upper
bound. For example, the set of real numbers R is complete but the set of rational numbers Q is not.
There are a number of results relating properties of the order topology to the completeness of X:

If the order topology on X is connected, X is complete.

X is connected under the order topology if and only if it is complete and there is no gap in X (a gap is two
points a and b in X with a < b such that no c satises a < c < b.)

X is complete if and only if every bounded set that is closed in the order topology is compact.

A totally ordered set (with its order topology) which is a complete lattice is compact. Examples are the closed intervals
of real numbers, e.g. the unit interval [0,1], and the anely extended real number system (extended real number line).
There are order-preserving homeomorphisms between these examples.

36.3.7 Sums of orders


For any two disjoint total orders (A1 , 1 ) and (A2 , 2 ) , there is a natural order + on the set A1 A2 , which is
called the sum of the two orders or sometimes just A1 + A2 :

For x, y A1 A2 , x + y holds if and only if one of the following holds:

1. x, y A1 and x 1 y
2. x, y A2 and x 2 y
3. x A1 and y A2
36.4. ORDERS ON THE CARTESIAN PRODUCT OF TOTALLY ORDERED SETS 111

Intutitively, this means that the elements of the second set are added on top of the elements of the rst set.
More generally, if (I, ) is a totally ordered index set, and for each i I the structure (Ai , i ) is a linear order,
where the sets Ai are pairwise disjoint, then the natural total order on i Ai is dened by

For x, y iI Ai , x y holds if:

1. Either there is some i I with x i y


2. or there are some i < j in I with x Ai , y Aj

36.4 Orders on the Cartesian product of totally ordered sets


In order of increasing strength, i.e., decreasing sets of pairs, three of the possible orders on the Cartesian product of
two totally ordered sets are:

Lexicographical order: (a,b) (c,d) if and only if a < c or (a = c and b d). This is a total order.

(a,b) (c,d) if and only if a c and b d (the product order). This is a partial order.

(a,b) (c,d) if and only if (a < c and b < d) or (a = c and b = d) (the reexive closure of the direct product of
the corresponding strict total orders). This is also a partial order.

All three can similarly be dened for the Cartesian product of more than two sets.
Applied to the vector space Rn , each of these make it an ordered vector space.
See also examples of partially ordered sets.
A real function of n real variables dened on a subset of Rn denes a strict weak order and a corresponding total
preorder on that subset.

36.5 Related structures


A binary relation that is antisymmetric, transitive, and reexive (but not necessarily total) is a partial order.
A group with a compatible total order is a totally ordered group.
There are only a few nontrivial structures that are (interdenable as) reducts of a total order. Forgetting the orientation
results in a betweenness relation. Forgetting the location of the ends results in a cyclic order. Forgetting both data
results in a separation relation.[3]

36.6 See also


Order theory

Well-order

Suslins problem

Countryman line

Prex order a downward total partial order

36.7 Notes
[1] Nederpelt, Rob (2004). Chapter 20.2: Ordered Sets. Orderings. Logical Reasoning: A First Course. Texts in Computing.
3 (3rd, Revised ed.). Kings College Publications. p. 325. ISBN 0-9543006-7-X.
112 CHAPTER 36. TOTAL ORDER

[2] Nederpelt, Rob (2004). Chapter 20.3: Ordered Sets. Linear orderings. Logical Reasoning: A First Course. Texts in
Computing. 3 (3rd, Revisied ed.). Kings College Publications. p. 330. ISBN 0-9543006-7-X.

[3] Macpherson, H. Dugald (2011), A survey of homogeneous structures (PDF), Discrete Mathematics, doi:10.1016/j.disc.2011.01.024,
retrieved 28 April 2011

36.8 References
George Grtzer (1971). Lattice theory: rst concepts and distributive lattices. W. H. Freeman and Co. ISBN
0-7167-0442-0

John G. Hocking and Gail S. Young (1961). Topology. Corrected reprint, Dover, 1988. ISBN 0-486-65676-4
Chapter 37

Total relation

In mathematics, a binary relation R over a set X is total or complete if for all a and b in X, a is related to b or b is
related to a (or both).
In mathematical notation, this is

a, b X(aRb bRa)

Total relations are sometimes said to have comparability.

37.1 Examples
For example, is less than or equal to is a total relation over the set of real numbers, because for two numbers either
the rst is less than or equal to the second, or the second is less than or equal to the rst. On the other hand, is less
than is not a total relation, since one can pick two equal numbers, and then neither the rst is less than the second, nor
is the second less than the rst. (But note that is less than is a weak order which gives rise to a total order, namely
is less than or equal to. The relationship between strict orders and weak orders is discussed at partially ordered set.)
The relation is a subset of is also not total because, for example, neither of the sets {1,2} and {3,4} is a subset of
the other.

37.2 Properties and related notions


Totality implies reexivity.
If a transitive relation is also total, it is a total preorder. If a partial order is also total, it is a total order.
A binary relation R over X is called connex if for all a and b in X such that a b, a is related to b or b is related to a
(or both):[1]

a, b X(aRb bRa (a = b))

Connexity does not imply reexivity. A strict partial order is a strict total order if and only if it is connex.

37.3 See also

Total order

113
114 CHAPTER 37. TOTAL RELATION

37.4 References
[1] Rautenberg, Wolfgang (2010), A Concise Introduction to Mathematical Logic (3rd ed.), New York: Springer Science+Business
Media, ISBN 978-1-4419-1220-6, doi:10.1007/978-1-4419-1221-3
Chapter 38

Transitive closure

For other uses, see Closure (disambiguation).


This article is about the transitive closure of a binary relation. For the transitive closure of a set, see transitive set
Transitive closure.

In mathematics, the transitive closure of a binary relation R on a set X is the smallest relation on X that contains R
and is transitive.
For example, if X is a set of airports and x R y means there is a direct ight from airport x to airport y" (for x and y
in X), then the transitive closure of R on X is the relation R+ such that x R+ y means it is possible to y from x to y in
one or more ights. Informally, the transitive closure gives you the set of all places you can get to from any starting
place.
More formally, the transitive closure of a binary relation R on a set X is the transitive relation R+ on set X such that R+
contains R and R+ is minimal (Lidl & Pilz (1998), p. 337). If the binary relation itself is transitive, then the transitive
closure is that same binary relation; otherwise, the transitive closure is a dierent relation.

38.1 Transitive relations and examples


A relation R on a set X is transitive if, for all x, y, z in X, whenever x R y and y R z then x R z. Examples of transitive
relations include the equality relation on any set, the less than or equal relation on any linearly ordered set, and the
relation "x was born before y" on the set of all people. Symbolically, this can be denoted as: if x < y and y < z then x
< z.
One example of a non-transitive relation is city x can be reached via a direct ight from city y" on the set of all cities.
Simply because there is a direct ight from one city to a second city, and a direct ight from the second city to the
third, does not imply there is a direct ight from the rst city to the third. The transitive closure of this relation is a
dierent relation, namely there is a sequence of direct ights that begins at city x and ends at city y". Every relation
can be extended in a similar way to a transitive relation.
An example of a non-transitive relation with a less meaningful transitive closure is "x is the day of the week after y".
The transitive closure of this relation is some day x comes after a day y on the calendar, which is trivially true for all
days of the week x and y (and thus equivalent to the Cartesian square, which is "x and y are both days of the week).

38.2 Existence and description


For any relation R, the transitive closure of R always exists. To see this, note that the intersection of any family of
transitive relations is again transitive. Furthermore, there exists at least one transitive relation containing R, namely
the trivial one: X X. The transitive closure of R is then given by the intersection of all transitive relations containing
R.
For nite sets, we can construct the transitive closure step by step, starting from R and adding transitive edges. This
gives the intuition for a general construction. For any set X, we can prove that transitive closure is given by the

115
116 CHAPTER 38. TRANSITIVE CLOSURE

following expression


R+ = Ri .
i{1,2,3,...}

where Ri is the i-th power of R, dened inductively by

R1 = R

and, for i > 0 ,

Ri+1 = R Ri

where denotes composition of relations.


To show that the above denition of R+ is the least transitive relation containing R, we show that it contains R, that it
is transitive, and that it is the smallest set with both of those characteristics.

R R+ : R+ contains all of the Ri , so in particular R+ contains R .


R+ is transitive: every element of R+ is in one of the Ri , so R+ must be transitive by the following reasoning:
if (s1 , s2 ) Rj and (s2 , s3 ) Rk , then from compositions associativity, (s1 , s3 ) Rj+k (and thus in R+
) because of the denition of Ri .
R+ is minimal: Let G be any transitive relation containing R , we want to show that R+ G . It is sucient to
show that for every i > 0 , Ri G . Well, since G contains R , R1 G . And since G is transitive, whenever
Ri G , Ri+1 G according to the construction of Ri and what it means to be transitive. Therefore, by
induction, G contains every Ri , and thus also R+ .

38.3 Properties
The intersection of two transitive relations is transitive.
The union of two transitive relations need not be transitive. To preserve transitivity, one must take the transitive
closure. This occurs, for example, when taking the union of two equivalence relations or two preorders. To obtain a
new equivalence relation or preorder one must take the transitive closure (reexivity and symmetryin the case of
equivalence relationsare automatic).

38.4 In graph theory


In computer science, the concept of transitive closure can be thought of as constructing a data structure that makes
it possible to answer reachability questions. That is, can one get from node a to node d in one or more hops? A
binary relation tells you only that node a is connected to node b, and that node b is connected to node c, etc. After
the transitive closure is constructed, as depicted in the following gure, in an O(1) operation one may determine that
node d is reachable from node a. The data structure is typically stored as a matrix, so if matrix[1][4] = 1, then it is
the case that node 1 can reach node 4 through one or more hops.
The transitive closure of a directed acyclic graph (DAG) is the reachability relation of the DAG and a strict partial
order.

38.5 In logic and computational complexity


The transitive closure of a binary relation cannot, in general, be expressed in rst-order logic (FO). This means that
one cannot write a formula using predicate symbols R and T that will be satised in any model if and only if T is
38.6. IN DATABASE QUERY LANGUAGES 117

Input

Output

Transitive closure constructs the output graph from the input graph.

the transitive closure of R. In nite model theory, rst-order logic (FO) extended with a transitive closure operator
is usually called transitive closure logic, and abbreviated FO(TC) or just TC. TC is a sub-type of xpoint logics.
The fact that FO(TC) is strictly more expressive than FO was discovered by Ronald Fagin in 1974; the result was
then rediscovered by Alfred Aho and Jerey Ullman in 1979, who proposed to use xpoint logic as a database query
language (Libkin 2004:vii). With more recent concepts of nite model theory, proof that FO(TC) is strictly more
expressive than FO follows immediately from the fact that FO(TC) is not Gaifman-local (Libkin 2004:49).
In computational complexity theory, the complexity class NL corresponds precisely to the set of logical sentences
expressible in TC. This is because the transitive closure property has a close relationship with the NL-complete
problem STCON for nding directed paths in a graph. Similarly, the class L is rst-order logic with the commutative,
transitive closure. When transitive closure is added to second-order logic instead, we obtain PSPACE.

38.6 In database query languages


Further information: Hierarchical and recursive queries in SQL

Since the 1980s Oracle Database has implemented a proprietary SQL extension CONNECT BY... START WITH
that allows the computation of a transitive closure as part of a declarative query. The SQL 3 (1999) standard added
a more general WITH RECURSIVE construct also allowing transitive closures to be computed inside the query
processor; as of 2011 the latter is implemented in IBM DB2, Microsoft SQL Server, Oracle, and PostgreSQL,
although not in MySQL (Benedikt and Senellart 2011:189).
Datalog also implements transitive closure computations (Silberschatz et al. 2010:C.3.6).

38.7 Algorithms
Ecient algorithms for computing the transitive closure of a graph can be found in Nuutila (1995). The fastest worst-
case methods, which are not practical, reduce the problem to matrix multiplication. The problem can also be solved
by the FloydWarshall algorithm, or by repeated breadth-rst search or depth-rst search starting from each node of
the graph.
More recent research has explored ecient ways of computing transitive closure on distributed systems based on the
MapReduce paradigm (Afrati et al. 2011).
118 CHAPTER 38. TRANSITIVE CLOSURE

38.8 See also


Ancestral relation

Deductive closure
Reexive closure

Symmetric closure

Transitive reduction (a smallest relation having the transitive closure of R as its transitive closure)

38.9 References
Lidl, R. and Pilz, G., 1998, Applied abstract algebra, 2nd edition, Undergraduate Texts in Mathematics,
Springer, ISBN 0-387-98290-6
Keller, U., 2004, Some Remarks on the Denability of Transitive Closure in First-order Logic and Datalog
(unpublished manuscript)
Erich Grdel; Phokion G. Kolaitis; Leonid Libkin; Maarten Marx; Joel Spencer; Moshe Y. Vardi; Yde Venema;
Scott Weinstein (2007). Finite Model Theory and Its Applications. Springer. pp. 151152. ISBN 978-3-540-
68804-4.

Libkin, Leonid (2004), Elements of Finite Model Theory, Springer, ISBN 978-3-540-21202-7
Heinz-Dieter Ebbinghaus; Jrg Flum (1999). Finite Model Theory (2nd ed.). Springer. pp. 123124, 151161,
220235. ISBN 978-3-540-28787-2.
Aho, A. V.; Ullman, J. D. (1979). Universality of data retrieval languages. Proceedings of the 6th ACM
SIGACT-SIGPLAN Symposium on Principles of programming languages - POPL '79. p. 110. doi:10.1145/567752.567763.
Benedikt, M.; Senellart, P. (2011). Databases. In Blum, Edward K.; Aho, Alfred V. Computer Science.
The Hardware, Software and Heart of It. pp. 169229. ISBN 978-1-4614-1167-3. doi:10.1007/978-1-4614-
1168-0_10.
Nuutila, E., Ecient Transitive Closure Computation in Large Digraphs. Acta Polytechnica Scandinavica,
Mathematics and Computing in Engineering Series No. 74, Helsinki 1995, 124 pages. Published by the
Finnish Academy of Technology. ISBN 951-666-451-2, ISSN 1237-2404, UDC 681.3.

Abraham Silberschatz; Henry Korth; S. Sudarshan (2010). Database System Concepts (6th ed.). McGraw-Hill.
ISBN 978-0-07-352332-3. Appendix C (online only)

Foto N. Afrati, Vinayak Borkar, Michael Carey, Neoklis Polyzotis, Jerey D. Ullman, Map-Reduce Extensions
and Recursive Queries, EDBT 2011, March 2224, 2011, Uppsala, Sweden, ISBN 978-1-4503-0528-0

38.10 External links


"Transitive closure and reduction", The Stony Brook Algorithm Repository, Steven Skiena .
"Apti Algoritmi", An example and some C++ implementations of algorithms that calculate the transitive closure
of a given binary relation, Vreda Pieterse.
Chapter 39

Trichotomy (mathematics)

In mathematics, the law of trichotomy states that every real number is either positive, negative, or zero.[1] More
generally, trichotomy is the property of an order relation < on a set X that for any x and y, exactly one of the
following holds: x < y, x = y , or x > y .[2][3]
In mathematical notation, this is

x X y X ((x < y (y < x) (x = y) ) ((x < y) y < x (x = y) ) ((x < y) (y < x) x = y )) .

Assuming that the ordering is irreexive and transitive, this can be simplied to

x X y X ((x < y) (y < x) (x = y)) .

In classical logic, this axiom of trichotomy holds for ordinary comparison between real numbers and therefore also
for comparisons between integers and between rational numbers. The law does not hold in general in intuitionistic
logic.
In ZermeloFraenkel set theory and Bernays set theory, the law of trichotomy holds between the cardinal numbers
of well-orderable sets even without the axiom of choice. If the axiom of choice holds, then trichotomy holds between
arbitrary cardinal numbers (because they are all well-orderable in that case).[4]
More generally, a binary relation R on X is trichotomous if for all x and y in X exactly one of xRy, yRx or x=y holds.
If such a relation is also transitive it is a strict total order; this is a special case of a strict weak order. For example, in
the case of three element set {a,b,c} the relation R given by aRb, aRc, bRc is a strict total order, while the relation R
given by the cyclic aRb, bRc, cRa is a non-transitive trichotomous relation.
In the denition of an ordered integral domain or ordered eld, the law of trichotomy is usually taken as more
foundational than the law of total order.
A trichotomous relation cannot be reexive, since xRx must be false. If a trichotomous relation is transitive, it is
trivially antisymmetric and also asymmetric, since xRy and yRx cannot both hold.

39.1 See also

Begrisschrift contains an early formulation of the law of trichotomy

Dichotomy

Law of noncontradiction

Law of excluded middle

119
120 CHAPTER 39. TRICHOTOMY (MATHEMATICS)

39.2 References
[1] Trichotomy Law at MathWorld

[2] Jerrold E. Marsden & Michael J. Homan (1993) Elementary Classical Analysis, page 27, W. H. Freeman and Company
ISBN 0-7167-2105-8

[3] H.S. Bear (1997) An Introduction to Mathematical Analysis, page 11, Academic Press ISBN 0-12-083940-7

[4] Bernays, Paul (1991). Axiomatic Set Theory. Dover Publications. ISBN 0-486-66637-9.
Chapter 40

Weak ordering

Not to be confused with weak order of permutations.


In mathematics, especially order theory, a weak ordering is a mathematical formalization of the intuitive notion of

a<c<b

a<b a<b
c<a<b c<b a<c a<b<c

c<a a,b,c a<c


c<b b<c

c<b<a b<a b<a b<a<c


c<a b<c

b<c<a

The 13 possible strict weak orderings on a set of three elements {a, b, c}. The only partially ordered sets are coloured, while totally
ordered ones are in black. Two orderings are shown as connected by an edge if they dier by a single dichotomy.

121
122 CHAPTER 40. WEAK ORDERING

a ranking of a set, some of whose members may be tied with each other. Weak orders are a generalization of totally
ordered sets (rankings without ties) and are in turn generalized by partially ordered sets and preorders.[1]
There are several common ways of formalizing weak orderings, that are dierent from each other but cryptomorphic
(interconvertable with no loss of information): they may be axiomatized as strict weak orderings (partially ordered
sets in which incomparability is a transitive relation), as total preorders (transitive binary relations in which at least
one of the two possible relations exists between every pair of elements), or as ordered partitions (partitions of the
elements into disjoint subsets, together with a total order on the subsets). In many cases another representation called
a preferential arrangement based on a utility function is also possible.
Weak orderings are counted by the ordered Bell numbers. They are used in computer science as part of partition
renement algorithms, and in the C++ Standard Library.[2]

40.1 Examples
In horse racing, the use of photo nishes has eliminated some, but not all, ties or (as they are called in this context)
dead heats, so the outcome of a horse race may be modeled by a weak ordering.[3] In an example from the Maryland
Hunt Cup steeplechase in 2007, The Bruce was the clear winner, but two horses Bug River and Lear Charm tied for
second place, with the remaining horses farther back; three horses did not nish.[4] In the weak ordering describing
this outcome, The Bruce would be rst, Bug River and Lear Charm would be ranked after The Bruce but before all
the other horses that nished, and the three horses that did not nish would be placed last in the order but tied with
each other.
The points of the Euclidean plane may be ordered by their distance from the origin, giving another example of a weak
ordering with innitely many elements, innitely many subsets of tied elements (the sets of points that belong to a
common circle centered at the origin), and innitely many points within these subsets. Although this ordering has a
smallest element (the origin itself), it does not have any second-smallest elements, nor any largest element.
Opinion polling in political elections provides an example of a type of ordering that resembles weak orderings, but is
better modeled mathematically in other ways. In the results of a poll, one candidate may be clearly ahead of another,
or the two candidates may be statistically tied, meaning not that their poll results are equal but rather that they are
within the margin of error of each other. However, if candidate x is statistically tied with y, and y is statistically tied
with z, it might still be possible for x to be clearly better than z, so being tied is not in this case a transitive relation.
Because of this possibility, rankings of this type are better modeled as semiorders than as weak orderings.[5]

40.2 Axiomatizations

40.2.1 Strict weak orderings

A strict weak ordering is a binary relation < on a set S that is a strict partial order (a transitive relation that is
irreexive, or equivalently,[6] that is asymmetric) in which the relation neither a < b nor b < a" is transitive.[1]
Therefore, a strict weak ordering has the following properties:

For all x in S, it is not the case that x < x (irreexivity).

For all x, y in S, if x < y then it is not the case that y < x (asymmetry).

For all x, y, z in S, if x < y and y < z then x < z (transitivity).

For all x, y, z in S, if x is incomparable with y (neither x < y nor y < x hold), and y is incomparable with z, then
x is incomparable with z (transitivity of incomparability).

This list of properties is somewhat redundant, in that asymmetry implies irreexivity, and in that irreexivity and
transitivity together imply asymmetry.
The incomparability relation is an equivalence relation, and its equivalence classes partition the elements of S, and
are totally ordered by <. Conversely, any total order on a partition of S gives rise to a strict weak ordering in which x
< y if and only if there exists sets A and B in the partition with x in A, y in B, and A < B in the total order.
40.2. AXIOMATIZATIONS 123

Not every partial order obeys the transitive law for incomparability. For instance, consider the partial order in the
set {a, b, c} dened by the relationship b < c. The pairs a, b and a, c are incomparable but b and c are related, so
incomparability does not form an equivalence relation and this example is not a strict weak ordering.
Transitivity of incomparability (together with transitivity) can also be stated in the following forms:

If x < y, then for all z, either x < z or z < y or both.

Or:

If x is incomparable with y, then for all z x, z y, either (x < z and y < z) or (z < x and z < y) or (z is
incomparable with x and z is incomparable with y).

40.2.2 Total preorders

Strict weak orders are very closely related to total preorders or (non-strict) weak orders, and the same mathematical
concepts that can be modeled with strict weak orderings can be modeled equally well with total preorders. A total
preorder or weak order is a preorder that is total; that is, no pair of items is incomparable. A total preorder satises
the following properties:

For all x, y, and z, if x y and y z then x z (transitivity).

For all x and y, x y or y x (totality).

Hence, for all x, x x (reexivity).

A total order is a total preorder which is antisymmetric, in other words, which is also a partial order. Total preorders
are sometimes also called preference relations.
The complement of a strict weak order is a total preorder, and vice versa, but it seems more natural to relate strict
weak orders and total preorders in a way that preserves rather than reverses the order of the elements. Thus we take
the inverse of the complement: for a strict weak ordering <, dene a total preorder by setting x y whenever it is
not the case that y < x. In the other direction, to dene a strict weak ordering < from a total preorder , set x < y
whenever it is not the case that y x.[7]
In any preorder there is a corresponding equivalence relation where two elements x and y are dened as equivalent if
x y and y x. In the case of a total preorder the corresponding partial order on the set of equivalence classes is a
total order. Two elements are equivalent in a total preorder if and only if they are incomparable in the corresponding
strict weak ordering.

40.2.3 Ordered partitions

A partition of a set S is a family of disjoint subsets of S that have S as their union. A partition, together with a
total order on the sets of the partition, gives a structure called by Richard P. Stanley an ordered partition[8] and by
Theodore Motzkin a list of sets.[9] An ordered partition of a nite set may be written as a nite sequence of the sets
in the partition: for instance, the three ordered partitions of the set {a, b} are

{a}, {b},
{b}, {a}, and
{a, b}.

In a strict weak ordering, the equivalence classes of incomparability give a set partition, in which the sets inherit a
total ordering from their elements, giving rise to an ordered partition. In the other direction, any ordered partition
gives rise to a strict weak ordering in which two elements are incomparable when they belong to the same set in the
partition, and otherwise inherit the order of the sets that contain them.
124 CHAPTER 40. WEAK ORDERING

40.2.4 Representation by functions

For sets of suciently small cardinality, a third axiomatization is possible, based on real-valued functions. If X is any
set and f a real-valued function on X then f induces a strict weak order on X by setting a < b if and only if f(a) < f(b).
The associated total preorder is given by setting a b if and only if f(a) f(b), and the associated equivalence by
setting a b if and only if f(a) = f(b).
The relations do not change when f is replaced by g o f (composite function), where g is a strictly increasing real-
valued function dened on at least the range of f. Thus e.g. a utility function denes a preference relation. In this
context, weak orderings are also known as preferential arrangements.[10]
If X is nite or countable, every weak order on X can be represented by a function in this way.[11] However, there
exist strict weak orders that have no corresponding real function. For example, there is no such function for the
lexicographic order on Rn . Thus, while in most preference relation models the relation denes a utility function up to
order-preserving transformations, there is no such function for lexicographic preferences.
More generally, if X is a set, and Y is a set with a strict weak ordering "<", and f a function from X to Y, then f
induces a strict weak ordering on X by setting a < b if and only if f(a) < f(b). As before, the associated total preorder
is given by setting a b if and only if f(a) f(b), and the associated equivalence by setting a b if and only
if f(a) f(b). It is not assumed here that f is an injective function, so a class of two equivalent elements on Y
may induce a larger class of equivalent elements on X. Also, f is not assumed to be an surjective function, so a class
of equivalent elements on Y may induce a smaller or empty class on X. However, the function f induces an injective
function that maps the partition on X to that on Y. Thus, in the case of nite partitions, the number of classes in X is
less than or equal to the number of classes on Y.

40.3 Related types of ordering

Semiorders generalize strict weak orderings, but do not assume transitivity of incomparability.[12] A strict weak order
that is trichotomous is called a strict total order.[13] The total preorder which is the inverse of its complement is in
this case a total order.
For a strict weak order "<" another associated reexive relation is its reexive closure, a (non-strict) partial order "".
The two associated reexive relations dier with regard to dierent a and b for which neither a < b nor b < a: in the
total preorder corresponding to a strict weak order we get a b and b a, while in the partial order given by the
reexive closure we get neither a b nor b a. For strict total orders these two associated reexive relations are
the same: the corresponding (non-strict) total order.[13] The reexive closure of a strict weak ordering is a type of
series-parallel partial order.

40.4 All weak orders on a nite set

40.4.1 Combinatorial enumeration

Main article: ordered Bell number

The number of distinct weak orders (represented either as strict weak orders or as total preorders) on an n-element
set is given by the following sequence (sequence A000670 in the OEIS):
These numbers are also called the Fubini numbers or ordered Bell numbers.
For example, for a set of three labeled items, there is one weak order in which all three items are tied. There are three
ways of partitioning the items into one singleton set and one group of two tied items, and each of these partitions
gives two weak orders (one in which the singleton is smaller than the group of two, and one in which this ordering is
reversed), giving six weak orders of this type. And there is a single way of partitioning the set into three singletons,
which can be totally ordered in six dierent ways. Thus, altogether, there are 13 dierent weak orders on three items.
40.4. ALL WEAK ORDERS ON A FINITE SET 125

(4,1,2,3)
(3,1,2,4) (4,2,1,3)
(3,2,1,4)

(4,1,3,2)
(2,1,3,4) (4,3,1,2)
(2,3,1,4)

(3,1,4,2) (4,2,3,1)
(2,1,4,3) (4,3,2,1)

(1,2,3,4) (3,4,1,2)
(1,3,2,4) (2,4,1,3)
(3,2,4,1)

(1,2,4,3) (3,4,2,1)

(1,4,2,3)
(2,3,4,1)

(1,3,4,2) (2,4,3,1)

(1,4,3,2)

The permutohedron on four elements, a three-dimensional convex polyhedron. It has 24 vertices, 36 edges, and 14 two-dimensional
faces, which all together with the whole three-dimensional polyhedron correspond to the 75 weak orderings on four elements.

40.4.2 Adjacency structure

Unlike for partial orders, the family of weak orderings on a given nite set is not in general connected by moves that
add or remove a single order relation to a given ordering. For instance, for three elements, the ordering in which all
three elements are tied diers by at least two pairs from any other weak ordering on the same set, in either the strict
weak ordering or total preorder axiomatizations. However, a dierent kind of move is possible, in which the weak
orderings on a set are more highly connected. Dene a dichotomy to be a weak ordering with two equivalence classes,
and dene a dichotomy to be compatible with a given weak ordering if every two elements that are related in the
ordering are either related in the same way or tied in the dichotomy. Alternatively, a dichotomy may be dened as a
Dedekind cut for a weak ordering. Then a weak ordering may be characterized by its set of compatible dichotomies.
For a nite set of labeled items, every pair of weak orderings may be connected to each other by a sequence of moves
that add or remove one dichotomy at a time to or from this set of dichotomies. Moreover, the undirected graph that
has the weak orderings as its vertices, and these moves as its edges, forms a partial cube.[14]
Geometrically, the total orderings of a given nite set may be represented as the vertices of a permutohedron, and the
dichotomies on this same set as the facets of the permutohedron. In this geometric representation, the weak orderings
on the set correspond to the faces of all dierent dimensions of the permutohedron (including the permutohedron
itself, but not the empty set, as a face). The codimension of a face gives the number of equivalence classes in the
corresponding weak ordering.[15] In this geometric representation the partial cube of moves on weak orderings is the
graph describing the covering relation of the face lattice of the permutohedron.
For instance, for n = 3, the permutohedron on three elements is just a regular hexagon. The face lattice of the hexagon
(again, including the hexagon itself as a face, but not including the empty set) has thirteen elements: one hexagon,
six edges, and six vertices, corresponding to the one completely tied weak ordering, six weak orderings with one tie,
and six total orderings. The graph of moves on these 13 weak orderings is shown in the gure.
126 CHAPTER 40. WEAK ORDERING

40.5 Applications
As mentioned above, weak orders have applications in utility theory.[11] In linear programming and other types of
combinatorial optimization problem, the prioritization of solutions or of bases is often given by a weak order, de-
termined by a real-valued objective function; the phenomenon of ties in these orderings is called degeneracy, and
several types of tie-breaking rule have been used to rene this weak ordering into a total ordering in order to prevent
problems caused by degeneracy.[16]
Weak orders have also been used in computer science, in partition renement based algorithms for lexicographic
breadth-rst search and lexicographic topological ordering. In these algorithms, a weak ordering on the vertices of a
graph (represented as a family of sets that partition the vertices, together with a doubly linked list providing a total
order on the sets) is gradually rened over the course of the algorithm, eventually producing a total ordering that is
the output of the algorithm.[17]
In the Standard Library for the C++ programming language, the set and multiset data types sort their input by a
comparison function that is specied at the time of template instantiation, and that is assumed to implement a strict
weak ordering.[2]

40.6 References
[1] Roberts, Fred; Tesman, Barry (2011), Applied Combinatorics (2nd ed.), CRC Press, Section 4.2.4 Weak Orders, pp. 254
256, ISBN 9781420099836.

[2] Josuttis, Nicolai M. (2012), The C++ Standard Library: A Tutorial and Reference, Addison-Wesley, p. 469, ISBN
9780132977739.

[3] de Koninck, J. M. (2009), Those Fascinating Numbers, American Mathematical Society, p. 4, ISBN 9780821886311.

[4] Baker, Kent (April 29, 2007), The Bruce hangs on for Hunt Cup victory: Bug River, Lear Charm nish in dead heat for
second, The Baltimore Sun, (Subscription required (help)).

[5] Regenwetter, Michel (2006), Behavioral Social Choice: Probabilistic Models, Statistical Inference, and Applications, Cam-
bridge University Press, pp. 113, ISBN 9780521536660.

[6] Flaka, V.; Jeek, J.; Kepka, T.; Kortelainen, J. (2007). Transitive Closures of Binary Relations I (PDF). Prague: School
of Mathematics - Physics Charles University. p. 1. Lemma 1.1 (iv). Note that this source refers to asymmetric relations
as strictly antisymmetric.

[7] Ehrgott, Matthias (2005), Multicriteria Optimization, Springer, Proposition 1.9, p. 10, ISBN 9783540276593.

[8] Stanley, Richard P. (1997), Enumerative Combinatorics, Vol. 2, Cambridge Studies in Advanced Mathematics, 62, Cam-
bridge University Press, p. 297.

[9] Motzkin, Theodore S. (1971), Sorting numbers for cylinders and other classication numbers, Combinatorics (Proc.
Sympos. Pure Math., Vol. XIX, Univ. California, Los Angeles, Calif., 1968), Providence, R.I.: Amer. Math. Soc., pp.
167176, MR 0332508.

[10] Gross, O. A. (1962), Preferential arrangements, The American Mathematical Monthly, 69: 48, MR 0130837, doi:10.2307/2312725.

[11] Roberts, Fred S. (1979), Measurement Theory, with Applications to Decisionmaking, Utility, and the Social Sciences, Ency-
clopedia of Mathematics and its Applications, 7, Addison-Wesley, Theorem 3.1, ISBN 978-0-201-13506-0.

[12] Luce, R. Duncan (1956), Semiorders and a theory of utility discrimination, Econometrica, 24: 178191, JSTOR 1905751,
MR 0078632, doi:10.2307/1905751.

[13] Velleman, Daniel J. (2006), How to Prove It: A Structured Approach, Cambridge University Press, p. 204, ISBN 9780521675994.

[14] Eppstein, David; Falmagne, Jean-Claude; Ovchinnikov, Sergei (2008), Media Theory: Interdisciplinary Applied Mathemat-
ics, Springer, Section 9.4, Weak Orders and Cubical Complexes, pp. 188196.

[15] Ziegler, Gnter M. (1995), Lectures on Polytopes, Graduate Texts in Mathematics, 152, Springer, p. 18.

[16] Chvtal, Vaek (1983), Linear Programming, Macmillan, pp. 2938, ISBN 9780716715870.

[17] Habib, Michel; Paul, Christophe; Viennot, Laurent (1999), Partition renement techniques: an interesting algorithmic tool
kit, International Journal of Foundations of Computer Science, 10 (2): 147170, MR 1759929, doi:10.1142/S0129054199000125.
Chapter 41

Well-founded relation

Noetherian induction redirects here. For the use in topology, see Noetherian topological space.

In mathematics, a binary relation, R, is called well-founded (or wellfounded) on a class X if every non-empty subset
S X has a minimal element with respect to R; that is, some element m not related by sRm (for instance, "m is not
smaller than s) for any s S.

S X (S = m S s S (s, m)
/ R).

(Some authors include an extra condition that R is set-like, i.e., that the elements less than any given element form a
set.)
Equivalently, assuming some choice, a relation is well-founded if it contains no countable innite descending chains:
that is, there is no innite sequence x0 , x1 , x2 , ... of elements of X such that xn R x for every natural number n.
In order theory, a partial order is called well-founded if the corresponding strict order is a well-founded relation. If
the order is a total order then it is called a well-order.
In set theory, a set x is called a well-founded set if the set membership relation is well-founded on the transitive
closure of x. The axiom of regularity, which is one of the axioms of ZermeloFraenkel set theory, asserts that all sets
are well-founded.
A relation R is converse well-founded, upwards well-founded or Noetherian on X, if the converse relation R1
is well-founded on X. In this case R is also said to satisfy the ascending chain condition. In the context of rewriting
systems, a Noetherian relation is also called terminating.

41.1 Induction and recursion


An important reason that well-founded relations are interesting is because a version of transnite induction can be
used on them: if (X, R) is a well-founded relation, P(x) is some property of elements of X, and we want to show that

P(x) holds for all elements x of X,

it suces to show that:

If x is an element of X and P(y) is true for all y such that y R x, then P(x) must also be true.

That is,

x X [(y X (y R x P (y))) P (x)] x X P (x).

127
128 CHAPTER 41. WELL-FOUNDED RELATION

Well-founded induction is sometimes called Noetherian induction,[1] after Emmy Noether.


On par with induction, well-founded relations also support construction of objects by transnite recursion. Let (X,
R) be a set-like well-founded relation and F a function that assigns an object F(x, g) to each pair of an element x X
and a function g on the initial segment {y: y R x} of X. Then there is a unique function G such that for every x X,

G(x) = F (x, G|{y:y R x} )

That is, if we want to construct a function G on X, we may dene G(x) using the values of G(y) for y R x.
As an example, consider the well-founded relation (N, S), where N is the set of all natural numbers, and S is the graph
of the successor function x x + 1. Then induction on S is the usual mathematical induction, and recursion on S
gives primitive recursion. If we consider the order relation (N, <), we obtain complete induction, and course-of-values
recursion. The statement that (N, <) is well-founded is also known as the well-ordering principle.
There are other interesting special cases of well-founded induction. When the well-founded relation is the usual
ordering on the class of all ordinal numbers, the technique is called transnite induction. When the well-founded set
is a set of recursively-dened data structures, the technique is called structural induction. When the well-founded
relation is set membership on the universal class, the technique is known as -induction. See those articles for more
details.

41.2 Examples
Well-founded relations which are not totally ordered include:

the positive integers {1, 2, 3, ...}, with the order dened by a < b if and only if a divides b and a b.
the set of all nite strings over a xed alphabet, with the order dened by s < t if and only if s is a proper
substring of t.
the set N N of pairs of natural numbers, ordered by (n1 , n2 ) < (m1 , m2 ) if and only if n1 < m1 and n2 < m2 .
the set of all regular expressions over a xed alphabet, with the order dened by s < t if and only if s is a proper
subexpression of t.
any class whose elements are sets, with the relation (is an element of). This is the axiom of regularity.
the nodes of any nite directed acyclic graph, with the relation R dened such that a R b if and only if there is
an edge from a to b.

Examples of relations that are not well-founded include:

the negative integers {1, 2, 3, }, with the usual order, since any unbounded subset has no least element.
The set of strings over a nite alphabet with more than one element, under the usual (lexicographic) order,
since the sequence B > AB > AAB > AAAB > is an innite descending chain. This relation fails
to be well-founded even though the entire set has a minimum element, namely the empty string.
the rational numbers (or reals) under the standard ordering, since, for example, the set of positive rationals (or
reals) lacks a minimum.

41.3 Other properties


If (X, <) is a well-founded relation and x is an element of X, then the descending chains starting at x are all nite, but
this does not mean that their lengths are necessarily bounded. Consider the following example: Let X be the union of
the positive integers and a new element , which is bigger than any integer. Then X is a well-founded set, but there
are descending chains starting at of arbitrary great (nite) length; the chain , n 1, n 2, ..., 2, 1 has length n
for any n.
41.4. REFLEXIVITY 129

The Mostowski collapse lemma implies that set membership is a universal among the extensional well-founded rela-
tions: for any set-like well-founded relation R on a class X which is extensional, there exists a class C such that (X,R)
is isomorphic to (C,).

41.4 Reexivity
A relation R is said to be reexive if a R a holds for every a in the domain of the relation. Every reexive relation on a
nonempty domain has innite descending chains, because any constant sequence is a descending chain. For example,
in the natural numbers with their usual order , we have 1 1 1 . To avoid these trivial descending
sequences, when working with a reexive relation R it is common to use (perhaps implicitly) the alternate relation R
dened such that a R b if and only if a R b and a b. In the context of the natural numbers, this means that the
relation <, which is well-founded, is used instead of the relation , which is not. In some texts, the denition of a
well-founded relation is changed from the denition above to include this convention.

41.5 References
[1] Bourbaki, N. (1972) Elements of mathematics. Commutative algebra, Addison-Wesley.

Just, Winfried and Weese, Martin, Discovering Modern Set theory. I, American Mathematical Society (1998)
ISBN 0-8218-0266-6.
Chapter 42

Well-order

In mathematics, a well-order (or well-ordering or well-order relation) on a set S is a total order on S with the
property that every non-empty subset of S has a least element in this ordering. The set S together with the well-order
relation is then called a well-ordered set. In some academic articles and textbooks these terms are instead written as
wellorder, wellordered, and wellordering or well order, well ordered, and well ordering.
Every non-empty well-ordered set has a least element. Every element s of a well-ordered set, except a possible greatest
element, has a unique successor (next element), namely the least element of the subset of all elements greater than
s. There may be elements besides the least element which have no predecessor (see Natural numbers below for an
example). In a well-ordered set S, every subset T which has an upper bound has a least upper bound, namely the least
element of the subset of all upper bounds of T in S.
If is a non-strict well ordering, then < is a strict well ordering. A relation is a strict well ordering if and only if it is
a well-founded strict total order. The distinction between strict and non-strict well orders is often ignored since they
are easily interconvertible.
Every well-ordered set is uniquely order isomorphic to a unique ordinal number, called the order type of the well-
ordered set. The well-ordering theorem, which is equivalent to the axiom of choice, states that every set can be well
ordered. If a set is well ordered (or even if it merely admits a well-founded relation), the proof technique of transnite
induction can be used to prove that a given statement is true for all elements of the set.
The observation that the natural numbers are well ordered by the usual less-than relation is commonly called the
well-ordering principle (for natural numbers).

42.1 Ordinal numbers

Main article: Ordinal number

Every well-ordered set is uniquely order isomorphic to a unique ordinal number, called the order type of the well-
ordered set. The position of each element within the ordered set is also given by an ordinal number. In the case of a
nite set, the basic operation of counting, to nd the ordinal number of a particular object, or to nd the object with
a particular ordinal number, corresponds to assigning ordinal numbers one by one to the objects. The size (number
of elements, cardinal number) of a nite set is equal to the order type. Counting in the everyday sense typically starts
from one, so it assigns to each object the size of the initial segment with that object as last element. Note that these
numbers are one more than the formal ordinal numbers according to the isomorphic order, because these are equal
to the number of earlier objects (which corresponds to counting from zero). Thus for nite n, the expression "n-th
element of a well-ordered set requires context to know whether this counts from zero or one. In a notation "-th
element where can also be an innite ordinal, it will typically count from zero.
For an innite set the order type determines the cardinality, but not conversely: well-ordered sets of a particular
cardinality can have many dierent order types. For a countably innite set, the set of possible order types is even
uncountable.

130
42.2. EXAMPLES AND COUNTEREXAMPLES 131

42.2 Examples and counterexamples

42.2.1 Natural numbers


The standard ordering of the natural numbers is a well ordering and has the additional property that every non-zero
natural number has a unique predecessor.
Another well ordering of the natural numbers is given by dening that all even numbers are less than all odd numbers,
and the usual ordering applies within the evens and the odds:

0 2 4 6 8 ... 1 3 5 7 9 ...

This is a well-ordered set of order type + . Every element has a successor (there is no largest element). Two
elements lack a predecessor: 0 and 1.

42.2.2 Integers
Unlike the standard ordering of the natural numbers, the standard ordering of the integers is not a well ordering,
since, for example, the set of negative integers does not contain a least element.
The following relation R is an example of well ordering of the integers: x R y if and only if one of the following
conditions holds:

1. x = 0
2. x is positive, and y is negative
3. x and y are both positive, and x y
4. x and y are both negative, and |x| |y|

This relation R can be visualized as follows:

0 1 2 3 4 ... 1 2 3 ...

R is isomorphic to the ordinal number + .


Another relation for well ordering the integers is the following denition: x y i (|x| < |y| or (|x| = |y| and x y)).
This well order can be visualized as follows:

0 1 1 2 2 3 3 4 4 ...

This has the order type .

42.2.3 Reals
The standard ordering of any real interval is not a well ordering, since, for example, the open interval (0, 1) [0,1]
does not contain a least element. From the ZFC axioms of set theory (including the axiom of choice) one can show
that there is a well order of the reals. Also Wacaw Sierpiski proved that ZF + GCH (the generalized continuum
hypothesis) imply the axiom of choice and hence a well order of the reals. Nonetheless, it is possible to show that
the ZFC+GCH axioms alone are not sucient to prove the existence of a denable (by a formula) well order of
the reals.[1] However it is consistent with ZFC that a denable well ordering of the reals existsfor example, it is
consistent with ZFC that V=L, and it follows from ZFC+V=L that a particular formula well orders the reals, or indeed
any set.
An uncountable subset of the real numbers with the standard ordering cannot be a well order: Suppose X is a subset
of R well ordered by . For each x in X, let s(x) be the successor of x in ordering on X (unless x is the last element
of X). Let A = { (x, s(x)) | x X } whose elements are nonempty and disjoint intervals. Each such interval contains
132 CHAPTER 42. WELL-ORDER

at least one rational number, so there is an injective function from A to Q. There is an injection from X to A (except
possibly for a last element of X which could be mapped to zero later). And it is well known that there is an injection
from Q to the natural numbers (which could be chosen to avoid hitting zero). Thus there is an injection from X to the
natural numbers which means that X is countable. On the other hand, a countably innite subset of the reals may or
may not be a well order with the standard "". For example,

The natural numbers are a well order under the standard ordering .
The set {1/n : n =1,2,3,...} has no least element and is therefore not a well order under standard ordering .

Examples of well orders:

The set of numbers { 2n | 0 n < } has order type .


The set of numbers { 2n 2mn | 0 m,n < } has order type . The previous set is the set of limit
points within the set. Within the set of real numbers, either with the ordinary topology or the order topology,
0 is also a limit point of the set. It is also a limit point of the set of limit points.
The set of numbers { 2n | 0 n < } { 1 } has order type + 1. With the order topology of this set, 1 is
a limit point of the set. With the ordinary topology (or equivalently, the order topology) of the real numbers it
is not.

42.3 Equivalent formulations


If a set is totally ordered, then the following are equivalent to each other:

1. The set is well ordered. That is, every nonempty subset has a least element.
2. Transnite induction works for the entire ordered set.
3. Every strictly decreasing sequence of elements of the set must terminate after only nitely many steps (assuming
the axiom of dependent choice).
4. Every subordering is isomorphic to an initial segment.

42.4 Order topology


Every well-ordered set can be made into a topological space by endowing it with the order topology.
With respect to this topology there can be two kinds of elements:

isolated points - these are the minimum and the elements with a predecessor.
limit points - this type does not occur in nite sets, and may or may not occur in an innite set; the innite sets
without limit point are the sets of order type , for example N.

For subsets we can distinguish:

Subsets with a maximum (that is, subsets which are bounded by themselves); this can be an isolated point or a
limit point of the whole set; in the latter case it may or may not be also a limit point of the subset.
Subsets which are unbounded by themselves but bounded in the whole set; they have no maximum, but a
supremum outside the subset; if the subset is non-empty this supremum is a limit point of the subset and hence
also of the whole set; if the subset is empty this supremum is the minimum of the whole set.
Subsets which are unbounded in the whole set.

A subset is conal in the whole set if and only if it is unbounded in the whole set or it has a maximum which is also
maximum of the whole set.
A well-ordered set as topological space is a rst-countable space if and only if it has order type less than or equal to
1 (omega-one), that is, if and only if the set is countable or has the smallest uncountable order type.
42.5. SEE ALSO 133

42.5 See also


Tree (set theory), generalization

Well-ordering theorem
Ordinal number

Well-founded set

Well partial order


Prewellordering

Directed set

42.6 References
[1] S. Feferman: Some Applications of the Notions of Forcing and Generic Sets, Fundamenta Mathematicae, 56 (1964)
325-345

Folland, Gerald B. (1999). Real Analysis: Modern Techniques and Their Applications. Pure and applied math-
ematics (2nd ed.). John Wiley & Sons. pp. 46, 9. ISBN 978-0-471-31716-6.
Chapter 43

Well-quasi-ordering

In mathematics, specically order theory, a well-quasi-ordering or wqo is a quasi-ordering such that any innite
sequence of elements x0 , x1 , x2 , from X contains an increasing pair xi xj with i < j .

43.1 Motivation
Well-founded induction can be used on any set with a well-founded relation, thus one is interested in when a quasi-
order is well-founded. However the class of well-founded quasiorders is not closed under certain operationsthat
is, when a quasi-order is used to obtain a new quasi-order on a set of structures derived from our original set, this
quasiorder is found to be not well-founded. By placing stronger restrictions on the original well-founded quasiordering
one can hope to ensure that our derived quasiorderings are still well-founded.
An example of this is the power set operation. Given a quasiordering for a set X one can dene a quasiorder +
on X 's power set P (X) by setting A + B if and only if for each element of A one can nd some element of B
that is larger than it under . One can show that this quasiordering on P (X) needn't be well-founded, but if one
takes the original quasi-ordering to be a well-quasi-ordering, then it is.

43.2 Formal denition


A well-quasi-ordering on a set X is a quasi-ordering (i.e., a reexive, transitive binary relation) such that any innite
sequence of elements x0 , x1 , x2 , from X contains an increasing pair xi xj with i < j . The set X is said to
be well-quasi-ordered, or shortly wqo.
A well partial order, or a wpo, is a wqo that is a proper ordering relation, i.e., it is antisymmetric.
Among other ways of dening wqos, one is to say that they are quasi-orderings which do not contain innite strictly
decreasing sequences (of the form x0 > x1 > x2 >) nor innite sequences of pairwise incomparable elements.
Hence a quasi-order (X,) is wqo if and only if (X,<) is well-founded and has no innite antichains.

43.3 Examples
(N, ) , the set of natural numbers with standard ordering, is a well partial order (in fact, a well-order). How-
ever, (Z, ) , the set of positive and negative integers, is not a well-quasi-order, because it is not well-founded.

(N, |) , the set of natural numbers ordered by divisibility, is not a well partial order: the prime numbers are an
innite antichain.

(Nk , ) , the set of vectors of k natural numbers (where k is nite) with component-wise ordering, is a well
partial order (Dicksons lemma). More generally, if (X, ) is well-quasi-order, then (X k , k ) is also a well-
quasi-order for all k .

134
43.4. WQOS VERSUS WELL PARTIAL ORDERS 135

Let X be an arbitrary nite set with at least two elements. The set X of words over X ordered lexicographically
(as in a dictionary) is not a well-quasi-order because it contains the innite decreasing sequence b, ab, aab, aaab, . . .
. Similarly, X ordered by the prex relation is not a well-quasi-order, because the previous sequence is an in-
nite antichain of this partial order. However, X ordered by the subsequence relation is a well partial order.[1]
(If X has only one element, these three partial orders are identical.)
More generally, (X , ) , the set of nite X -sequences ordered by embedding is a well-quasi-order if and
only if (X, ) is a well-quasi-order (Higmans lemma). Recall that one embeds a sequence u into a sequence v
by nding a subsequence of v that has the same length as u and that dominates it term by term. When (X, =)
is a nite unordered set, u v if and only if u is a subsequence of v .
(X , ) , the set of innite sequences over a well-quasi-order (X, ) , ordered by embedding, is not a well-
quasi-order in general. That is, Higmans lemma does not carry over to innite sequences. Better-quasi-
orderings have been introduced to generalize Higmans lemma to sequences of arbitrary lengths.
Embedding between nite trees with nodes labeled by elements of a wqo (X, ) is a wqo (Kruskals tree
theorem).
Embedding between innite trees with nodes labeled by elements of a wqo (X, ) is a wqo (Nash-Williams'
theorem).
Embedding between countable scattered linear order types is a well-quasi-order (Laver's theorem).
Embedding between countable boolean algebras is a well-quasi-order. This follows from Lavers theorem and
a theorem of Ketonen.
Finite graphs ordered by a notion of embedding called "graph minor" is a well-quasi-order (RobertsonSeymour
theorem).
Graphs of nite tree-depth ordered by the induced subgraph relation form a well-quasi-order,[2] as do the
cographs ordered by induced subgraphs.[3]

43.4 Wqos versus well partial orders


In practice, the wqos one manipulates are quite often not orderings (see examples above), and the theory is technically
smoother if we do not require antisymmetry, so it is built with wqos as the basic notion. On the other hand, according
to Milner 1985, no real gain in generality is obtained by considering quasi-orders rather than partial orders... it is simply
more convenient to do so.
Observe that a wpo is a wqo, and that a wqo gives rise to a wpo between equivalence classes induced by the kernel
of the wqo. For example, if we order Z by divisibility, we end up with n m if and only if n = m , so that
(Z, |) (N, |) .

43.5 Innite increasing subsequences


If ( X , ) is wqo then every innite sequence x0 , x1 , x2 , contains an innite increasing subsequence xn0
xn1 xn2 (with n0 < n1 < n2 <). Such a subsequence is sometimes called perfect. This can be proved by
a Ramsey argument: given some sequence (xi )i , consider the set I of indexes i such that xi has no larger or equal
xj to its right, i.e., with i < j . If I is innite, then the I -extracted subsequence contradicts the assumption that X
is wqo. So I is nite, and any xn with n larger than any index in I can be used as the starting point of an innite
increasing subsequence.
The existence of such innite increasing subsequences is sometimes taken as a denition for well-quasi-ordering,
leading to an equivalent notion.

43.6 Properties of wqos


Given a quasiordering (X, ) the quasiordering (P (X), + ) dened by A + B a Ab B(a
b) is well-founded if and only if (X, ) is a wqo.[4]
136 CHAPTER 43. WELL-QUASI-ORDERING

A quasiordering is a wqo if and only if the corresponding partial order (obtained by quotienting by x y
x y y x ) has no innite descending sequences or antichains. (This can be proved using a Ramsey
argument as above.)
Given a well-quasi-ordering (X, ) , any sequence of subsets S0 S1 ... X such that i N, x, y
X, x yx Si y Si eventually stabilises (meaning there is an index n N such that Sn = Sn+1 = ...
; subsets S X with the property x, y X, x y x S y S are usually called upward-closed):
assuming the contrary i Nj N, j > i, x Sj \ Si , a contradiction is reached by extracting an innite
non-ascending subsequence.
Given a well-quasi-ordering (X, ) , any subset S X which is upward-closed with respect to has a nite
number of minimal elements w.r.t. , for otherwise the minimal elements of S would constitute an innite
antichain.

43.7 Notes
^ Here x < y means: x y and y x.

[1] Gasarch, W. (1998), A survey of recursive combinatorics, Handbook of Recursive Mathematics, Vol. 2, Stud. Logic
Found. Math., 139, Amsterdam: North-Holland, pp. 10411176, MR 1673598, doi:10.1016/S0049-237X(98)80049-9.
See in particular page 1160.
[2] Neetil, Jaroslav; Ossona de Mendez, Patrice (2012), Lemma 6.13, Sparsity: Graphs, Structures, and Algorithms, Algo-
rithms and Combinatorics, 28, Heidelberg: Springer, p. 137, ISBN 978-3-642-27874-7, MR 2920058, doi:10.1007/978-
3-642-27875-4.
[3] Damaschke, Peter (1990), Induced subgraphs and well-quasi-ordering, Journal of Graph Theory, 14 (4): 427435, MR
1067237, doi:10.1002/jgt.3190140406.
[4] Forster, Thomas (2003). Better-quasi-orderings and coinduction. Theoretical Computer Science. 309 (13): 111123.
doi:10.1016/S0304-3975(03)00131-2.

43.8 References
Dickson, L. E. (1913). Finiteness of the odd perfect and primitive abundant numbers with r distinct prime
factors. American Journal of Mathematics. 35 (4): 413422. JSTOR 2370405. doi:10.2307/2370405.
Higman, G. (1952). Ordering by divisibility in abstract algebras. Proceedings of the London Mathematical
Society. 2: 326336. doi:10.1112/plms/s3-2.1.326.
Kruskal, J. B. (1972). The theory of well-quasi-ordering: A frequently discovered concept. Journal of
Combinatorial Theory. Series A. 13 (3): 297305. doi:10.1016/0097-3165(72)90063-5.
Ketonen, Jussi (1978). The structure of countable Boolean algebras. Annals of Mathematics. 108 (1): 4189.
JSTOR 1970929. doi:10.2307/1970929.
Milner, E. C. (1985). Basic WQO- and BQO-theory. In Rival, I. Graphs and Order. The Role of Graphs in
the Theory of Ordered Sets and Its Applications. D. Reidel Publishing Co. pp. 487502. ISBN 90-277-1943-8.
Gallier, Jean H. (1991). Whats so special about Kruskals theorem and the ordinal o? A survey of some re-
sults in proof theory. Annals of Pure and Applied Logic. 53 (3): 199260. doi:10.1016/0168-0072(91)90022-
E.

43.9 See also


Better-quasi-ordering
Prewellordering
Well-order
43.10. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 137

43.10 Text and image sources, contributors, and licenses


43.10.1 Text
Accessibility relation Source: https://en.wikipedia.org/wiki/Accessibility_relation?oldid=798054303 Contributors: Michael Hardy, Wma-
han, Roisterer, Urhixidur, Spug, Firsfron, Nivaca, Bgwhite, KSchutte, Imz, Mhss, Lacatosias, Mets501, Simeon, Gregbard, Egrin,
Classicalecon, Auntof6, Alexbot, Sun Creator, Dthomsen8, Addbot, Frehley, Tempodivalse, AnomieBOT, Sophivorus, John of Reading,
MikeyMouse10, ISpamThisSite3, Widr, BG19bot, Forestfrolic, Jochen Burghardt, Cabbagist, Deacon Vorbis, KolbertBot, Trdc95 and
Anonymous: 32
Ancestral relation Source: https://en.wikipedia.org/wiki/Ancestral_relation?oldid=794455835 Contributors: Zundark, Meloman, CBM,
Palnot, AnomieBOT, Malts y, TricksterWolf, Jochen Burghardt, Magic links bot and Anonymous: 5
Antisymmetric relation Source: https://en.wikipedia.org/wiki/Antisymmetric_relation?oldid=800807645 Contributors: AxelBoldt, Patrick,
Michael Hardy, Charles Matthews, Dcoetzee, MathMartin, Rholton, Tea2min, Giftlite, Sam Hocevar, Lumidek, Nparikh, Ascnder, Paul
August, El C, Spoon!, Rpresser, Jumbuck, EvenT, Adrian.benko, Isnow, Nivaca, Fresheneesz, Chobot, Roboto de Ajvol, Kelovy, Arthur
Rubin, Wasseralm, SmackBot, RDBury, Incnis Mrsi, InverseHypercube, Jcarroll, Bluebot, Jdthood, Lambiam, DabMachine, Mike Fikes,
Gregbard, WillowW, JAnDbot, TAnthony, Catskineater, Mark lee stillwell, Semmelweiss, PaulTanenbaum, Jackfork, Henry Delforn
(old), DuaneLAnderson, P30Carl, Hakuku, Addbot, LaaknorBot, CarsracBot, Verbal, Legobot, Luckas-bot, Yobot, Xqbot, Adavis444,
Erik9bot, De bezige bij, EmausBot, Theophil789, Joel B. Lewis, AvocatoBot, MadGuy7023, Alexjbest, Jochen Burghardt, Nbrader,
Boga159, Suelru, Abhinay15134001 and Anonymous: 23
Asymmetric relation Source: https://en.wikipedia.org/wiki/Asymmetric_relation?oldid=796725284 Contributors: Patrick, Charles Matthews,
Dcoetzee, Tea2min, Brianjd, Paul August, Oliphaunt, Palica, Fresheneesz, Mditto, Arthur Rubin, SmackBot, Jdthood, Tsca.bot, Jklin,
Lakinekaki, CRGreathouse, Gregbard, Alastair Haines, Magioladitis, Robin S, Jy00912345, Taifunbrowser, VolkovBot, Henry Delforn
(old), Bender2k14, Addbot, Yobot, Erik9bot, WikitanvirBot, Theophil789, TheodoreYou, IfYouDoIfYouDon't, MerlIwBot, Helpful
Pixie Bot, Lerutit, Jochen Burghardt, Dualspace, Bender the Bot and Anonymous: 15
Better-quasi-ordering Source: https://en.wikipedia.org/wiki/Better-quasi-ordering?oldid=794455916 Contributors: Xezbeth, Chris Capoc-
cia, Citation bot, LilHelpa, FrescoBot, Gongfarmerzed, BG19bot, Jochen Burghardt and Mzheng086
Binary relation Source: https://en.wikipedia.org/wiki/Binary_relation?oldid=796087488 Contributors: AxelBoldt, Bryan Derksen, Zun-
dark, Tarquin, Jan Hidders, Roadrunner, Mjb, Tomo, Patrick, Xavic69, Michael Hardy, Wshun, Isomorphic, Dominus, Ixfd64, Takuya-
Murata, Charles Matthews, Timwi, Dcoetzee, Jitse Niesen, Robbot, Chocolateboy, MathMartin, Tea2min, Giftlite, Fropu, Dratman,
Jorge Stol, Jlr~enwiki, Andycjp, Quarl, Guanabot, Yuval madar, Slipstream, Paul August, Elwikipedista~enwiki, Shanes, EmilJ, Ran-
dall Holmes, Ardric47, Obradovic Goran, Eje211, Alansohn, Dallashan~enwiki, Keenan Pepper, PAR, Adrian.benko, Oleg Alexan-
drov, Joriki, Linas, Apokrif, MFH, Dpv, Pigcatian, Penumbra2000, Fresheneesz, Chobot, YurikBot, Hairy Dude, Koeyahoo, Trova-
tore, Bota47, Arthur Rubin, Netrapt, SmackBot, Royalguard11, SEIBasaurus, Cybercobra, Jon Awbrey, Turms, Lambiam, Dbtfz, Mr
Stephen, Mets501, Dreftymac, Happy-melon, Petr Matas, CRGreathouse, CBM, Yrodro, WillowW, Xantharius, Thijs!bot, Egrin,
Rlupsa, Marek69, Fayenatic london, JAnDbot, MER-C, TAnthony, Magioladitis, Vanish2, Avicennasis, David Eppstein, Robin S, Akurn,
Adavidb, LajujKej, Owlgorithm, Djjrjr, Policron, DavidCBryant, Quux0r, VolkovBot, Boute, Vipinhari, Anonymous Dissident, PaulTa-
nenbaum, Jackfork, Wykypydya, Dmcq, AlleborgoBot, AHMartin, Ocsenave, Sftd, Paradoctor, Henry Delforn (old), MiNombreDeGuerra,
DuaneLAnderson, Anchor Link Bot, CBM2, Classicalecon, ClueBot, Snigbrook, Rhubbarb, Hans Adler, SounderBruce, SilvonenBot,
BYS2, Plmday, Addbot, LinkFA-Bot, Tide rolls, Jarble, Legobot, Luckas-bot, Yobot, Ht686rg90, Lesliepogi, Pcap, Labus, Nallim-
bot, Reindra, FredrikMeyer, AnomieBOT, Floquenbeam, Royote, Hahahaha4, Materialscientist, Belkovich, Citation bot, Racconish,
Jellystones, Xqbot, Isheden, Geero, GhalyBot, Ernsts, Howard McCay, Constructive editor, Mark Renier, Mfwitten, RandomDSdevel,
NearSetAccount, SpaceFlight89, Yunshui, Miracle Pen, Brambleclawx, RjwilmsiBot, Nomen4Omen, Chharvey, SporkBot, OnePt618,
Sameer143, Socialservice, ResearchRave, ClueBot NG, Wcherowi, Frietjes, Helpful Pixie Bot, Koertefa, BG19bot, ChrisGualtieri,
YFdyh-bot, Dexbot, Makecat-bot, ScitDei, Lerutit, Jochen Burghardt, Jodosma, Karim132, Cosmia Nebula, Monkbot, Pratincola, ,
Neycalazans, Some1Redirects4You, The Quixotic Potato, Luis150902, Magic links bot and Anonymous: 114
Comparability Source: https://en.wikipedia.org/wiki/Comparability?oldid=794455109 Contributors: Patrick, Charles Matthews, Tea2min,
Tsirel, Oleg Alexandrov, Rjwilmsi, Archelon, 16@r, Mets501, CharacterZero, ThreeBlindMice, THF, David Eppstein, PaulTanenbaum,
Burdel britnico, Justin W Smith, SchreiberBike, Yobot, Gamewizard71, ChuispastonBot, Snotbot, Ollieinc, Jochen Burghardt and
Anonymous: 5
Congruence relation Source: https://en.wikipedia.org/wiki/Congruence_relation?oldid=794453149 Contributors: AxelBoldt, Toby~enwiki,
Toby Bartels, Michael Hardy, Charles Matthews, Jitse Niesen, Greenrd, Aleph4, Romanm, MathMartin, Henrygb, Tosha, Giftlite, Arved,
Waltpohl, Mani1, ZeroOne, Boger1, PWilkinson, WojciechSwiderski~enwiki, Bookandcoee, Oleg Alexandrov, Bluemoose, Marudub-
shinki, Pako, Pasky, Kevmitch, DavidHouse~enwiki, Reyk, Netrapt, SmackBot, Imz, Mgreenbe, BiT, Mhss, Bluebot, Mohamed Al-
Dabbagh, Jim.belk, Mets501, Vaughan Pratt, CRGreathouse, CBM, Sam Staton, Goldencako, Thijs!bot, Gdickeson, JAnDbot, Magio-
laditis, JoergenB, VolkovBot, EuTuga, Anchor Link Bot, Amahoney, Sandeepjshenoy, Hans Adler, Palnot, Addbot, Mancini0, Legobot,
Yobot, Calle, DannyAsher, Obersachsebot, GrouchoBot, FrescoBot, Stpasta, TobeBot, Dinamik-bot, Ayamewolfe, EmausBot, ZroBot,
Toshio Yamaguchi, Elaz85, Anita5192, MerlIwBot, SteenthIWbot, Jochen Burghardt, Pietro13, Dizzyzane, Magic links bot and Anony-
mous: 35
Covering relation Source: https://en.wikipedia.org/wiki/Covering_relation?oldid=798042015 Contributors: Rp, Aleph4, Rjwilmsi, Michael
Slone, Nbarth, Mr Stephen, Gregbard, David Eppstein, R'n'B, PaulTanenbaum, Arcfrk, Addbot, Jamiguet, Yobot, LucienBOT, BG19bot,
Solomon7968, Jochen Burghardt, , Alabarre, KolbertBot and Anonymous: 1
Dense order Source: https://en.wikipedia.org/wiki/Dense_order?oldid=794453275 Contributors: EmilJ, Physicistjedi, MarSch, Michael
Slone, SmackBot, Imz, Melchoir, Turms, JAnDbot, David Eppstein, VolkovBot, TXiKiBoT, Palnot, Addbot, ., Pcap, Erik9bot,
Tom.Reding, ZroBot, Helpful Pixie Bot, Qetuth, Jochen Burghardt, Brirush, GeoreyT2000, tale.cohomology, Magic links bot and
Anonymous: 6
Dependence relation Source: https://en.wikipedia.org/wiki/Dependence_relation?oldid=794455154 Contributors: Michael Hardy, Charles
Matthews, Jitse Niesen, Josh Parris, Wavelength, Robbjedi, Keegan, Geometry guy, 777sms and Jochen Burghardt
138 CHAPTER 43. WELL-QUASI-ORDERING

Dependency relation Source: https://en.wikipedia.org/wiki/Dependency_relation?oldid=794455330 Contributors: Michael Hardy, William


M. Connolley, GPHemsley, Robbot, Wizzy, Goochelaar, Linas, Nihiltres, Jsnx, SmackBot, Chris the speller, NickPenguin, Mukake, David
Eppstein, Homei, Classicalecon, Addbot, Yobot, WikitanvirBot, Jochen Burghardt, Mark viking, W. P. Uzer, Christian Nassif-Haynes,
JMP EAX and Anonymous: 4
Directed set Source: https://en.wikipedia.org/wiki/Directed_set?oldid=794455379 Contributors: AxelBoldt, The Anome, SimonP, Patrick,
Michael Hardy, AugPi, Nikai, Revolver, Dfeuer, Dysprosia, Tea2min, Giftlite, Markus Krtzsch, Smimram, Paul August, Varuna,
Msh210, Eric Kvaalen, SteinbDJ, Linas, Dionyziz, Salix alba, Margosbot~enwiki, Hairy Dude, Zwobot, Futanari, Reedy, Fitch, Mhss,
Vaughan Pratt, CBM, Blaisorblade, Larvy, JoergenB, CommonsDelinker, LokiClock, Don4of4, AlleborgoBot, SieBot, Tom Leinster,
He7d3r, Beroal, Palnot, Plmday, Legobot, Luckas-bot, Obersachsebot, Yaddie, ZroBot, Haraldbre, Helpful Pixie Bot, BG19bot, Freeze
S, Jochen Burghardt, Zoydb, Austrartsua, Some1Redirects4You, Magic links bot and Anonymous: 30
Equality (mathematics) Source: https://en.wikipedia.org/wiki/Equality_(mathematics)?oldid=801814534 Contributors: Toby Bartels,
Patrick, Michael Hardy, TakuyaMurata, Looxix~enwiki, Pizza Puzzle, Charles Matthews, Dysprosia, WhisperToMe, Banno, Robbot,
RedWolf, Lowellian, Tea2min, Alan Liefting, Giftlite, Christopher Parham, Recentchanges, Michael Devore, Jabowery, DefLog~enwiki,
Chowbok, Smiller933, Shahab, AlexG, Wrp103, Plugwash, Rgdboer, Spoon!, Iltseng, PWilkinson, MPerel, Jumbuck, Msh210, Hu,
Japanese Searobin, Simetrical, Linas, MattGiuca, Isnow, Wbeek, Qwertyus, Island, Scottkeir, Jshadias, Pasky, FlaBot, VKokielov, Mar-
gosbot~enwiki, Fresheneesz, Chobot, DVdm, Gwernol, Laurentius, Hairy Dude, Pi Delport, Pnrj, TransUtopian, Reyk, Tinlv7, SmackBot,
RDBury, Incnis Mrsi, Melchoir, Blue520, Josephprymak, BiT, Gilliam, Bluebot, Nbarth, Jdthood, Jon Awbrey, Lambiam, Attys, Load-
master, Mets501, Tauolunga, CBM, Sdorrance, Simeon, Gregbard, Cydebot, Benzi455, Blaisorblade, Xantharius, Uv~enwiki, Marek69,
Cj67, Dugwiki, AntiVandalBot, Widefox, Malcolm, JAnDbot, Thenub314, Edward321, R'n'B, Policron, Alan U. Kennington, Anony-
mous Dissident, PaulTanenbaum, UnitedStatesian, Enigmaman, Vikrant42, Tachikomas All Memory, Flyer22 Reborn, Ctxppc, Clue-
Bot, BodhisattvaBot, SilvonenBot, Addbot, Debresser, Numbo3-bot, Apteva, Legobot, Luckas-bot, Yobot, TaBOT-zerem, Amirobot,
Pcap, KamikazeBot, Ningauble, Bryan.burgers, MassimoAr, AnomieBOT, Kingpin13, Citation bot, Capricorn42, Kevfest08, NOrbeck,
VladimirReshetnikov, Der Falke, FrescoBot, Tkuvho, AmphBot, RedBot, Jauhienij, TobeBot, Belovedeagle, Vrenator, CobraBot, Duo-
duoduo, Ebe123, ZroBot, Sungzungkim, D.Lazard, ClueBot NG, Iiii I I I, Wcherowi, Faus, Titodutta, ChrisGualtieri, Jochen Burghardt,
Brirush, DialaceStarvy, Monkbot, Sunmist, Loraof, Samf4u, Lizard Pancakes123456789012345678901234567890, Gmalaven, This is
a mobile phone, Deacon Vorbis and Anonymous: 85
Equipollence (geometry) Source: https://en.wikipedia.org/wiki/Equipollence_(geometry)?oldid=801931499 Contributors: Michael Hardy,
Mdob, Rgdboer, Siddhant, Sadads, Cydebot, Addbot, Omnipaedista, Erik9bot, J.Victor, Specs112, Makhokh, Brad7777, Jochen Burghardt,
JJMC89, InternetArchiveBot, Cyrus the Penner, KolbertBot and Anonymous: 1
Equivalence class Source: https://en.wikipedia.org/wiki/Equivalence_class?oldid=794453377 Contributors: AxelBoldt, Zundark, The
Anome, Patrick, Michael Hardy, Wshun, Salsa Shark, Revolver, Charles Matthews, Dysprosia, Wolfgang Kufner, Greenrd, Hyacinth,
Psychonaut, Naddy, GreatWhiteNortherner, Tea2min, Giftlite, WiseWoman, Lethe, Fropu, Fuzzy Logic, Noisy, Tibbetts, Liuyao,
Rgdboer, Msh210, MattGiuca, Graham87, Salix alba, Mike Segal, Jameshsher, Laurentius, Hede2000, Arthur Rubin, Paul D. Anderson,
Lunch, SmackBot, Mhss, Nbarth, Javalenok, Jennica, Lhf, Frentos, Mets501, Andrew Delong, Egrin, Magioladitis, David Eppstein,
VolkovBot, LokiClock, Rjgodoy, Quietbritishjim, Dogah, Henry Delforn (old), Sjn28, Classicalecon, Watchduck, Kausikghatak, Addbot,
Jasper Deng, WikiDreamer Bot, Yobot, Calle, Rinke 80, Omnipaedista, Erik9bot, HJ Mitchell, WillNess, Igor Yalovecky, Quondum,
D.Lazard, Herebo, Wcherowi, Rpglover64, ChrisGualtieri, Jochen Burghardt, Brirush, Mark viking, A4b3c2d1e0f, Riddleh, Verdana
Bold, Norbornene, Addoergosum, Anareth, Ehatan, Tt8612399 and Anonymous: 49
Equivalence relation Source: https://en.wikipedia.org/wiki/Equivalence_relation?oldid=801840162 Contributors: AxelBoldt, Zundark,
Toby Bartels, PierreAbbat, Ryguasu, Stevertigo, Patrick, Michael Hardy, Wshun, Dominus, TakuyaMurata, William M. Connolley, AugPi,
Silversh, Ideyal, Revolver, Charles Matthews, Dysprosia, Hyacinth, Fibonacci, Phys, McKay, GPHemsley, Robbot, Fredrik, Romanm,
COGDEN, Ashley Y, Bkell, Tea2min, Tosha, Giftlite, Arved, ShaunMacPherson, Lethe, Herbee, Fropu, LiDaobing, AlexG, Paul Au-
gust, Elwikipedista~enwiki, FirstPrinciples, Rgdboer, Spearhead, Smalljim, SpeedyGonsales, Obradovic Goran, Haham hanuka, Kier-
ano, Msh210, Keenan Pepper, PAR, Jopxton, Oleg Alexandrov, Linas, Apokrif, MFH, BD2412, Salix alba, Alexb@cut-the-knot.com,
Mark J, Epitome83, Chobot, Algebraist, Roboto de Ajvol, YurikBot, Wavelength, RussBot, Nils Grimsmo, BOT-Superzerocool, Googl,
LarryLACa, Arthur Rubin, Pred, Cjfsyntropy, Draicone, RonnieBrown, SmackBot, Adam majewski, Melchoir, Stie, Srnec, Gilliam,
Gelingvistoj, Kurykh, Concerned cynic, Foxjwill, Vanished User 0001, Michael Ross, Jon Awbrey, Jim.belk, Feraudyh, CredoFromStart,
Michael Kinyon, JHunterJ, Vanished user 8ij3r8jwe, Mets501, Rschwieb, Captain Wacky, JForget, CRGreathouse, CBM, 345Kai,
Gregbard, Doctormatt, PepijnvdG, Tawkerbot4, Xantharius, Hanche, BetacommandBot, Thijs!bot, Egrin, Rlupsa, WilliamH, Rnealh,
Salgueiro~enwiki, JAnDbot, Thenub314, Magioladitis, VoABot II, JamesBWatson, MetsBot, Robin S, Philippe.beaudoin, Pekaje, Pomte,
Interwal, Cpiral, GaborLajos, Policron, Taifunbrowser, Idioma-bot, Station1, Davehi1, Billinghurst, Geanixx, AlleborgoBot, SieBot, Bot-
Multichill, This, that and the other, Henry Delforn (old), Aspects, OKBot, Bulkroosh, C1wang, Classicalecon, Wmli, Kclchan, Watch-
duck, Hans Adler, Qwfp, Cdegremo, Palnot, XLinkBot, Gerhardvalentin, Libcub, LaaknorBot, CarsracBot, Dyaa, Legobot, Luckas-bot,
Yobot, Ht686rg90, Gyro Copter, TheresNoTime, Andy.melnikov, ArthurBot, Xqbot, GrouchoBot, Lenore, RibotBOT, Antares5245,
FrescoBot, Sokbot3000, Anthonystevens2, ARandomNicole, Lost-n-translation, Tkuvho, SpaceFlight89, TobeBot, Miracle Pen, Emaus-
Bot, ReneGMata, AvicBot, General Rommel, TyA, Donner60, Gottlob Gdel, ClueBot NG, Bethre, Helpful Pixie Bot, BG19bot, Mark
Arsten, ChrisGualtieri, Rectipaedia, YFdyh-bot, Jochen Burghardt, Rushikeshjogdand1, David9550, Noix07, Adammwagner, Damon-
amc, Superegz, Razvan.belet, InternetArchiveBot, Anareth, Quiddital, Magic links bot and Anonymous: 113
Euclidean relation Source: https://en.wikipedia.org/wiki/Euclidean_relation?oldid=794389209 Contributors: Toby Bartels, Giftlite,
EmilJ, PAR, Apokrif, Salix alba, Hairy Dude, Otto ter Haar, Lhf, Turms, Gregbard, Egrin, David Eppstein, Robertgreer, Cdegremo,
Yobot, Yangtseyangtse, Helpful Pixie Bot, Bender the Bot and Anonymous: 7
Foundational relation Source: https://en.wikipedia.org/wiki/Foundational_relation?oldid=794455597 Contributors: Michael Hardy,
BiH, IkamusumeFan and Jochen Burghardt
Idempotent relation Source: https://en.wikipedia.org/wiki/Idempotent_relation?oldid=794453466 Contributors: Michael Hardy, Bearcat,
CBM, Deltahedron, Jochen Burghardt and Flokam
Intransitivity Source: https://en.wikipedia.org/wiki/Intransitivity?oldid=794453539 Contributors: Michael Hardy, Chris-martin, Rp,
6birc, Radicalsubversiv, Andres, Charles Matthews, Ruakh, Giftlite, Andris, Quickwik, D6, Xezbeth, Paul August, Jnestorius, Spoon!,
Polluks, SmackBot, Melchoir, Mauls, MisterHand, Lambiam, Iamagloworm, Dinkumator, CRGreathouse, CmdrObot, CBM, Thomas-
meeks, Ael 2, Thijs!bot, VoABot II, Cnilep, Ddxc, Anchor Link Bot, PixelBot, DumZiBoT, Pa68, Addbot, Forich, WissensDrster,
Undsoweiter, HRoestBot, ChronoKinetic, RjwilmsiBot, Uanfala, Chharvey, Diakov, Darcourse, Jochen Burghardt, GeoreyZanders,
Deacon Vorbis and Anonymous: 9
43.10. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 139

Inverse relation Source: https://en.wikipedia.org/wiki/Inverse_relation?oldid=794453831 Contributors: Patrick, Charles Matthews, Al-


tenmann, Tea2min, Giftlite, Kaldari, Spiy sperry, El C, Versageek, Jerey O. Gustafson, MFH, YurikBot, Cedar101, Paul Erik, Knowl-
edgeOfSelf, C.Fred, Nbarth, Faaaa, Nixeagle, Jon Awbrey, Mearnhardtfan, Coredesat, Slakr, DBooth, CBM, Thomasmeeks, Gogo Dodo,
Mojo Hand, Headbomb, Hut 8.5, MetsBot, David Eppstein, Real World Apple, Ars Tottle, VolkovBot, Iamthedeus, Maelgwnbot, Clas-
sicalecon, Rjd0060, Overstay, Trainshift, Pluto Car, Unco Guid, Viva La Information Revolution!, Autocratic Uzbek, Poke Salat Annie,
Flower Mound Belle, Navy Pierre, Mrs. Lovetts Meat Puppets, Chester County Dude, Southeast Penna Poppa, Delaware Valley Girl,
MystBot, Omphaloskeptor, Addbot, Quercus solaris, Luckas-bot, Yobot, Pcap, Materialscientist, MathHisSci, , John of Read-
ing, SporkBot, ClueBot NG, Wcherowi, Jochen Burghardt, Mark viking, JMP EAX, Hdjensofjfnen and Anonymous: 17
Partially ordered set Source: https://en.wikipedia.org/wiki/Partially_ordered_set?oldid=795341342 Contributors: Bryan Derksen, Zun-
dark, Tomo, Patrick, Bcrowell, Chinju, TakuyaMurata, GTBacchus, AugPi, Charles Matthews, Timwi, Dcoetzee, Dysprosia, Doradus,
Maximus Rex, Fibonacci, Tea2min, Giftlite, Markus Krtzsch, Fropu, Peruvianllama, Jason Quinn, Neilc, Gubbubu, DefLog~enwiki,
MarkSweep, Urhixidur, TheJames, Paul August, Bender235, Zaslav, Spoon!, Porton, Haham hanuka, DougOrleans, Msh210, Oleg
Alexandrov, Daira Hopwood, MFH, Salix alba, FlaBot, Vonkje, Chobot, Laurentius, Dmharvey, Vecter, JosephSilverman, Sanguinity,
Cedar101, Modify, RDBury, Incnis Mrsi, Brick Thrower, Cesine, Zanetu, Jcarroll, Nbarth, Jdthood, Javalenok, Kjetil1001, Dreadstar,
Esoth~enwiki, Mike Fikes, A. Pichler, Vaughan Pratt, CRGreathouse, L'uf, CBM, Werratal, Rlupsa, CZeke, Ill logic, JAnDbot, MER-
C, BrotherE, Tbleher, A3nm, David Eppstein, SlamDiego, Bissinger, Haseldon, Daniel5Ko, GaborLajos, NewEnglandYankee, Orphic,
RobertDanielEmerson, TXiKiBoT, Digby Tantrum, PaulTanenbaum, Arcfrk, SieBot, Mochan Shrestha, TheGhostOfAdrianMineha, The-
hotelambush, CodeTalker, Megaloxantha, Peiresc~enwiki, Cheesefondue, Jludwig, ClueBot, Morinus, Justin W Smith, Methossant, Pi
zero, Jonathanrcoxhead, Watchduck, ComputerGeezer, He7d3r, Hans Adler, Jtle515, Palnot, Marc van Leeuwen, Ankan babee, Addbot,
Download, Luckyz, Legobot, Kilom691, AnomieBOT, Erel Segal, Citation bot, SteveWoolf, Undsoweiter, FrescoBot, Nicolas Perrault
III, Conuente, Ricardo Ferreira de Oliveira, Throw it in the Fire, Darren Strash, Gnathan87, Setitup, EmausBot, John of Reading,
Febuiles, Dcirovic, Thecheesykid, ZroBot, Chharvey, The man who was Friday, SporkBot, Zfeinst, Rathgemz, Wcherowi, CocuBot,
Vdamanafshan, Mesoderm, MerlIwBot, Wbm1058, BG19bot, Jakshap, Paolo Lipparini, ElphiBot, Larion Garaczi, Lerutit, Aabhis,
Jochen Burghardt, Mark viking, Eamonford, Sgbmyr, K401sTL3, Tudor987, Victor Lesyk, Esquivalience, Deniz Stiegemann, Nbro,
Some1Redirects4You and Anonymous: 87
Preorder Source: https://en.wikipedia.org/wiki/Preorder?oldid=794454040 Contributors: AxelBoldt, Patrick, Repton, Delirium, An-
dres, Dysprosia, Greenrd, Big Bob the Finder, StevePowell, BenRG, Tea2min, Giftlite, Markus Krtzsch, Lethe, Fropu, Vadmium,
DefLog~enwiki, Zzo38, Jh51681, Barnaby dawson, 4pq1injbok, Paul August, EmilJ, Msh210, Melaen, Joriki, Linas, Dionyziz, Man-
darax, Salix alba, Cjoev, VKokielov, Mathbot, Jrtayloriv, YurikBot, Laurentius, Hairy Dude, WikidSmaht, Trovatore, Modify, Ne-
trapt, Wasseralm, SmackBot, XudongGuan~enwiki, DCary, Jdthood, Mets501, PaulGS, Stotr~enwiki, Zero sharp, CRGreathouse, CBM,
Michael A. White, Magioladitis, David Eppstein, Jwuthe2, Largoplazo, PaulTanenbaum, SieBot, Thehotelambush, MenoBot, Functor
salad, He7d3r, Sun Creator, Cenarium, Wikidsp, 1ForTheMoney, Palnot, , Legobot, Luckas-bot, AnomieBOT, DannyAsher,
Xqbot, VladimirReshetnikov, BercherP, ComputScientist, BrideOfKripkenstein, Notedgrant, WikitanvirBot, Lclem, Dfabera, SporkBot,
PatrickR2, Paolo Lipparini, RichardMills65, Khazar2, Lerutit, Jochen Burghardt, Reatlas, Damonamc, Cryptopocalypse and Anonymous:
27
Prewellordering Source: https://en.wikipedia.org/wiki/Prewellordering?oldid=794454087 Contributors: Zundark, Patrick, Charles Matthews,
EmilJ, Oleg Alexandrov, NickBush24, Trovatore, Tetracube, That Guy, From That Show!, SmackBot, Nbarth, Henry Delforn (old), Pal-
not, Citation bot, Jochen Burghardt, Deacon Vorbis and Anonymous: 3
Quasitransitive relation Source: https://en.wikipedia.org/wiki/Quasitransitive_relation?oldid=794454154 Contributors: Patrick, Charles
Matthews, Giftlite, Btyner, Salix alba, SmackBot, Zahid Abdassabur, CRGreathouse, CBM, Gregbard, Sam Staton, David Eppstein,
Erik9bot, I dream of horses, Deltahedron, Jochen Burghardt and Anonymous: 2
Quotient by an equivalence relation Source: https://en.wikipedia.org/wiki/Quotient_by_an_equivalence_relation?oldid=794454227
Contributors: Michael Hardy, TakuyaMurata, Bearcat, David Eppstein, D.Lazard, Jochen Burghardt, K9re11, Iaritmioawp and Fimatic
Rational consequence relation Source: https://en.wikipedia.org/wiki/Rational_consequence_relation?oldid=794458331 Contributors:
Michael Hardy, Oleg Alexandrov, Reetep, SmackBot, CmdrObot, Gregbard, Enlil2, DavidCBryant, Carriearchdale, AnomieBOT, Ti-
jfo098 and Jochen Burghardt
Reexive closure Source: https://en.wikipedia.org/wiki/Reflexive_closure?oldid=794454287 Contributors: Timwi, Arthur Rubin, Girl-
withglasses, Henry Delforn (old), Addbot, Dawynn, Pcap, Renato sr, GrouchoBot, EmausBot, Jochen Burghardt, Noremacskich, Vaizar
and Anonymous: 2
Reexive relation Source: https://en.wikipedia.org/wiki/Reflexive_relation?oldid=794388617 Contributors: AxelBoldt, DavidSJ, Patrick,
Wshun, TakuyaMurata, Looxix~enwiki, William M. Connolley, Charles Matthews, Josh Cherry, MathMartin, Henrygb, Tea2min, Giftlite,
Jason Quinn, Gubbubu, Urhixidur, Ascnder, Paul August, BenjBot, Spayrard, Spoon!, Jet57, LavosBacons, Wtmitchell, Bookandcof-
fee, Oleg Alexandrov, Joriki, Mel Etitis, LOL, Apokrif, MFH, Isnow, Audiovideo, Margosbot~enwiki, Fresheneesz, Chobot, YurikBot,
Laurentius, Maelin, Mathlaura, KarlHeg, Arthur Rubin, Izzynn, Jdthood, Mhym, Ceosion, Mike Fikes, Fjbex, CRGreathouse, CBM,
Gregbard, Farzaneh, Wikid77, JAnDbot, Policron, Joshua Issac, VolkovBot, Jackfork, Jamelan, Ocsenave, SieBot, Davidellerman, Henry
Delforn (old), Hello71, Cuyaken, ClueBot, Ywanne, Da rulz07, SoxBot III, WikHead, Addbot, Download, Favonian, Luckas-bot, Yobot,
Renato sr, Pkukiss, Galoubet, ArthurBot, Xqbot, Z0973, I dream of horses, RedBot, MastiBot, Gamewizard71, EmausBot, Dcirovic,
Dmayank, DimitriC, ClueBot NG, Kasirbot, Joel B. Lewis, BG19bot, Solomon7968, The1337gamer, ChrisGualtieri, Darcourse, Epti-
ed, Lerutit, Jochen Burghardt, Seanhalle, Cosmia Nebula, Dualspace and Anonymous: 38
Semiorder Source: https://en.wikipedia.org/wiki/Semiorder?oldid=794454381 Contributors: Rjwilmsi, Headbomb, David Eppstein,
Undsoweiter, Joel B. Lewis, Lerutit, Jochen Burghardt and Anonymous: 2
Separoid Source: https://en.wikipedia.org/wiki/Separoid?oldid=794456664 Contributors: Michael Hardy, BD2412, Salix alba, Smack-
Bot, Mhym, Gregbard, Magioladitis, Qworty, Hans Adler, Strausz~enwiki and Jochen Burghardt
Series-parallel partial order Source: https://en.wikipedia.org/wiki/Series-parallel_partial_order?oldid=794454451 Contributors: Za-
slav, A3nm, David Eppstein, Trappist the monk, Helpful Pixie Bot, Deltahedron, Jochen Burghardt and Anonymous: 1
Symmetric closure Source: https://en.wikipedia.org/wiki/Symmetric_closure?oldid=798154122 Contributors: Michael Hardy, Timwi,
Addbot, Pcap, ZroBot, YFdyh-bot, Jochen Burghardt, Deacon Vorbis and Anonymous: 3
Tolerance relation Source: https://en.wikipedia.org/wiki/Tolerance_relation?oldid=794389477 Contributors: Michael Hardy, Apokrif,
Rjwilmsi, SmackBot, NickPenguin, Floridi~enwiki, Classicalecon, Gabrno, Rektroth and Anonymous: 1
140 CHAPTER 43. WELL-QUASI-ORDERING

Total order Source: https://en.wikipedia.org/wiki/Total_order?oldid=794454583 Contributors: Damian Yerrick, AxelBoldt, Zundark,


XJaM, Fritzlein, Patrick, Michael Hardy, Dori, AugPi, Dysprosia, Jitse Niesen, Greenrd, Zoicon5, Hyacinth, VeryVerily, Fibonacci,
McKay, Aleph4, Gandalf61, MathMartin, Rursus, Tea2min, Giftlite, Mshonle~enwiki, Markus Krtzsch, Lethe, Waltpohl, DefLog~enwiki,
Alberto da Calvairate~enwiki, Quarl, Elroch, Paul August, Susvolans, Army1987, Func, Cmdrjameson, Tsirel, Msh210, Pion, Joriki,
MattGiuca, Yurik, OneWeirdDude, Salix alba, VKokielov, Mathbot, Margosbot~enwiki, Wastingmytime, Chobot, YurikBot, Hede2000,
Archelon, Tetracube, Rdore, Melchoir, Gelingvistoj, Mhss, Chris the speller, Bazonka, Jdthood, Javalenok, Michael Kinyon, Loadmaster,
Mets501, Iridescent, JRSpriggs, George100, CRGreathouse, CBM, Thomasmeeks, Oryanw~enwiki, VectorPosse, JAnDbot, TAnthony,
A3nm, David Eppstein, Infovarius, Osquar F, Drewmutt, PaulTanenbaum, SieBot, Ceroklis, Anchor Link Bot, Heinzi.at, WurmWoode,
Universityuser, Palnot, Marc van Leeuwen, Addbot, Tanhabot, AsphyxiateDrake, Luckas-bot, Yobot, Charlatino, White gecko, 1exec1,
Infvwl, GrouchoBot, Jsjunkie, Quondum, D.Lazard, SporkBot, Wcherowi, CocuBot, BG19bot, YumOooze, YFdyh-bot, Austinfeller,
Jochen Burghardt, Mark viking, , Graboy, Davyker, MrLJ and Anonymous: 58
Total relation Source: https://en.wikipedia.org/wiki/Total_relation?oldid=794454648 Contributors: Patrick, Charles Matthews, Dcoet-
zee, Jitse Niesen, Tea2min, Lethe, Alberto da Calvairate~enwiki, Paul August, Ntmatter, Oleg Alexandrov, Joriki, Salix alba, Nneonneo,
Mathbot, Fresheneesz, Bota47, Jdthood, Stotr~enwiki, JAnDbot, TXiKiBoT, Jamelan, Chstdu, Hans Adler, Erodium, Addbot, FrescoBot,
SporkBot, Helpful Pixie Bot, Deltahedron, Kephir, Jochen Burghardt, Dualspace and Anonymous: 11
Transitive closure Source: https://en.wikipedia.org/wiki/Transitive_closure?oldid=797080733 Contributors: Awaterl, Vkuncak, Patrick,
Michael Hardy, Charles Matthews, Timwi, Dcoetzee, Populus, Borislav, Tea2min, Giftlite, Fropu, Matt Crypto, Alexf, Quickwik,
Creidieki, Obradovic Goran, Oleg Alexandrov, Joriki, Neonfreon, Salix alba, GnniX, AL SAM, Bgwhite, Ott2, Arthur Rubin, Plas-
ticphilosopher, KnightRider~enwiki, Mhss, Plustgarten, Dreadstar, NeilFraser, Lyonsam, Loadmaster, JRSpriggs, CRGreathouse, CBM,
ShelfSkewed, Gregbard, Girlwithglasses, Kirtag Hratiba, Thijs!bot, JAnDbot, A3nm, David Eppstein, Yavoh, Cometstyles, VolkovBot,
Sdrucker, PaulTanenbaum, Jamelan, Tomaxer, LungZeno, Henry Delforn (old), DuaneLAnderson, CBM2, Classicalecon, ClueBot, Ben-
der2k14, PixelBot, AmirOnWiki, MountainGoat8, Tayste, Addbot, Luckas-bot, AnomieBOT, BenzolBot, RedBot, MastiBot, Trappist
the monk, Wizeguytristram, Quondum, Tijfo098, ClueBot NG, Wcherowi, BG19bot, Solomon7968, Danwizard208, Dmitri L. Slabk.,
Dexbot, Hungaricus, Jochen Burghardt, Me, Myself, and I are Here, Vpieterse~enwiki, Seahen, Artdadamo, Edgy4, Deacon Vorbis and
Anonymous: 30
Trichotomy (mathematics) Source: https://en.wikipedia.org/wiki/Trichotomy_(mathematics)?oldid=794454777 Contributors: Zundark,
Patrick, Michael Hardy, Arthur Frayn, Casu Marzu, Henrygb, UtherSRG, Tea2min, Macrakis, Paul August, Rgdboer, Oleg Alexandrov,
VKokielov, Michael Slone, Bota47, JJL, SmackBot, Mhss, Colonies Chris, Jdthood, Cybercobra, Courcelles, JRSpriggs, Meng.benjamin,
Gregbard, Diuoroethene, David Cherney, AntiVandalBot, Indeed123, Minnnnng, YohanN7, Mild Bill Hiccup, MilesAgain, Addbot,
Luckas-bot, AnomieBOT, Gtz, Xqbot, Omnipaedista, Constructive editor, ComputScientist, jlfr, Tkuvho, SporkBot, Paulmiko,
Helpful Pixie Bot, Zorglub x, Jochen Burghardt, Miket25 and Anonymous: 25
Weak ordering Source: https://en.wikipedia.org/wiki/Weak_ordering?oldid=794454860 Contributors: Patrick, Michael Hardy, Chinju,
Dcoetzee, MathMartin, Ruakh, Tea2min, Pretzelpaws, AlphaEtaPi, Zaslav, Aisaac, Rjwilmsi, GnniX, YurikBot, Gadget850, Modify,
Oli Filth, Jdthood, Chlewbot, Radiant chains, Jafet, CRGreathouse, Sdorrance, Gregbard, Widefox, Medinoc, Zeitlupe, David Eppstein,
Jonathanrcoxhead, Watchduck, Addbot, Kne1p, Forich, Citation bot, ArthurBot, Howard McCay, Citation bot 1, Andyx96, SporkBot,
Seabuoy, Richard1962, Joel B. Lewis, Helpful Pixie Bot, Jochen Burghardt, JustBerry, Bender the Bot and Anonymous: 13
Well-founded relation Source: https://en.wikipedia.org/wiki/Well-founded_relation?oldid=799555919 Contributors: The Anome, Ap,
Michael Hardy, Dominus, TakuyaMurata, Cyp, Charles Matthews, VeryVerily, Aleph4, Mountain, Tea2min, Filemon, Marekpetrik,
Lethe, Lupin, Waltpohl, Mani1, Paul August, EmilJ, Nahabedere, MZMcBride, R.e.b., FlaBot, Trovatore, Mikeblas, Crasshopper, Mar-
lasdad, That Guy, From That Show!, SmackBot, Ron.garcia, Nbarth, Mmehdi.g, Mets501, CBM, WillowW, Pgagge, JAnDbot, Albmont,
Leyo, Reedy Bot, Alexsmail, Thehotelambush, Mutilin, Addbot, KamikazeBot, RibotBOT, FrescoBot, Involutive-revolution, MerlIwBot,
Wohlfundi, Jochen Burghardt, YiFeiBot, , Fblanqui, Some1Redirects4You and Anonymous: 26
Well-order Source: https://en.wikipedia.org/wiki/Well-order?oldid=794454971 Contributors: AxelBoldt, Mav, Zundark, Josh Grosse,
Patrick, Michael Hardy, David Martland, Dominus, TakuyaMurata, Andres, Vargenau, Revolver, Charles Matthews, Timwi, Populus,
Aleph4, R3m0t, MathMartin, Tea2min, Tosha, Giftlite, Dbenbenn, Ian Maxwell, Lethe, Arturus~enwiki, Jorend, Karl-Henner, Rich
Farmbrough, Luqui, Paul August, Sligocki, RJFJR, Eyu100, Salix alba, FlaBot, Margosbot~enwiki, Chobot, YurikBot, Hairy Dude,
Trovatore, Obey, Bota47, Arthur Rubin, MullerHolk, Ghazer~enwiki, GrinBot~enwiki, KnightRider~enwiki, Alan McBeth, Gelingvistoj,
Mhss, Nbarth, Loodog, Jim.belk, Loadmaster, JRSpriggs, CBM, WeggeBot, Myasuda, Thijs!bot, Nadav1, Escarbot, Albmont, Odexios,
VolkovBot, Don4of4, SieBot, Rumping, Fyyer, Bender2k14, His Wikiness, Palnot, Addbot, Luckas-bot, Yobot, Ptbotgourou, Jarmiz,
Xqbot, GrouchoBot, Miyagawa, Adrionwells, Mitizhi, RjwilmsiBot, Honestrosewater, Hunterbd, SporkBot, Misshamid, ChrisGualtieri,
Khazar2, Jochen Burghardt, Jose Brox, Wicklet, Leegrc and Anonymous: 32
Well-quasi-ordering Source: https://en.wikipedia.org/wiki/Well-quasi-ordering?oldid=801742167 Contributors: Patrick, Chinju, Charles
Matthews, Tea2min, Peter Kwok, Rich Farmbrough, Paul August, EmilJ, R.e.b., Open2universe, PhS, That Guy, From That Show!,
Mets501, Pierre de Lyon, David Eppstein, Kope, R'n'B, Alexwright, Fcarreiro, Niceguyedc, Palnot, Addbot, DOI bot, AnomieBOT, Ci-
tation bot, FrescoBot, Citation bot 1, Gongfarmerzed, John of Reading, ZroBot, , Mastergreg82, Paolo Lipparini, CitationCleanerBot,
Je Erickson, Jochen Burghardt, Mark viking, Quenhitran, Anrnusna, , Gasarch and Anonymous: 7

43.10.2 Images
File:13-Weak-Orders.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/3c/13-Weak-Orders.svg License: Public do-
main Contributors: Transferred from en.wikipedia to Commons. Original artist: SVG version created by Jafet.vixle at en.wikipedia,
based on a public-domain PNG version created 2006-0 9-14 by David Eppstein
File:Ambox_important.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b4/Ambox_important.svg License: Public do-
main Contributors: Own work based on: Ambox scales.svg Original artist: Dsmurat, penubag
File:Birkhoff120.svg Source: https://upload.wikimedia.org/wikipedia/commons/7/7c/Birkhoff120.svg License: Public domain Contrib-
utors: Own work Original artist: David Eppstein
File:Commons-logo.svg Source: https://upload.wikimedia.org/wikipedia/en/4/4a/Commons-logo.svg License: PD Contributors: ? Orig-
inal artist: ?
43.10. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 141

File:Congruent_non-congruent_triangles.svg Source: https://upload.wikimedia.org/wikipedia/commons/1/12/Congruent_non-congruent_


triangles.svg License: CC BY-SA 3.0 Contributors: Own work Original artist: Lfahlberg
File:Directed_set,_but_no_join_semi-lattice.png Source: https://upload.wikimedia.org/wikipedia/commons/a/a2/Directed_set%2C_
but_no_join_semi-lattice.png License: CC BY-SA 3.0 Contributors: Own work Original artist: Jochen Burghardt
File:E-to-the-i-pi.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/35/E-to-the-i-pi.svg License: CC BY 2.5 Contribu-
tors: No machine-readable source provided. Own work assumed (based on copyright claims). Original artist: No machine-readable author
provided. Dermeister assumed (based on copyright claims).
File:Edit-clear.svg Source: https://upload.wikimedia.org/wikipedia/en/f/f2/Edit-clear.svg License: Public domain Contributors: The
Tango! Desktop Project. Original artist:
The people from the Tango! project. And according to the meta-data in the le, specically: Andreas Nilsson, and Jakub Steiner (although
minimally).
File:Equivalentie.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/38/Equivalentie.svg License: Public domain Con-
tributors: Own work Original artist: Rinke 80
File:Even_and_odd_antisymmetric_relation.png Source: https://upload.wikimedia.org/wikipedia/commons/c/cf/Even_and_odd_antisymmetric_
relation.png License: Public domain Contributors: I made this myself Original artist: Fresheneesz at English Wikipedia
File:Graph_of_non-injective,_non-surjective_function_(red)_and_of_bijective_function_(green).gif Source: https://upload.wikimedia.
org/wikipedia/commons/b/b0/Graph_of_non-injective%2C_non-surjective_function_%28red%29_and_of_bijective_function_%28green%
29.gif License: CC BY-SA 3.0 Contributors: Own work Original artist: Jochen Burghardt
File:GreaterThan.png Source: https://upload.wikimedia.org/wikipedia/commons/5/59/GreaterThan.png License: Public domain Con-
tributors: Transferred from en.wikipedia to Commons. Original artist: The original uploader was Fresheneesz at English Wikipedia
File:GreaterThanOrEqualTo.png Source: https://upload.wikimedia.org/wikipedia/commons/0/0b/GreaterThanOrEqualTo.png License:
Public domain Contributors: Transferred from en.wikipedia to Commons. Original artist: The original uploader was Fresheneesz at English
Wikipedia
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:Hasse_diagram_of_powerset_of_3_no_greatest_or_least.svg Source: https://upload.wikimedia.org/wikipedia/commons/9/9e/
Hasse_diagram_of_powerset_of_3_no_greatest_or_least.svg License: CC-BY-SA-3.0 Contributors: Based on File:Hasse diagram of
powerset of 3.svg Original artist: User:Fibonacci, User:KSmrq
File:Infinite_lattice_of_divisors.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/e6/Infinite_lattice_of_divisors.svg
License: Public domain Contributors: w:de:Datei:Verband TeilerN.png Original artist: <a href='//commons.wikimedia.org/wiki/File:
Watchduck.svg' class='image'><img alt='Watchduck.svg' src='https://upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.
svg/40px-Watchduck.svg.png' width='40' height='46' srcset='https://upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.
svg/60px-Watchduck.svg.png 1.5x, https://upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.svg/80px-Watchduck.svg.
png 2x' data-le-width='703' data-le-height='806' /></a> Watchduck (a.k.a. Tilman Piesk)
File:LampFlowchart.svg Source: https://upload.wikimedia.org/wikipedia/commons/9/91/LampFlowchart.svg License: CC-BY-SA-
3.0 Contributors: vector version of Image:LampFlowchart.png Original artist: svg by Booyabazooka

File:Lexicographic_order_on_pairs_of_natural_numbers.svg Source: https://upload.wikimedia.org/wikipedia/commons/8/8a/Lexicographic_


order_on_pairs_of_natural_numbers.svg License: CC BY-SA 3.0 Contributors: Own work Original artist: Jochen Burghardt
File:Monotonic_but_nonhomomorphic_map_between_lattices.gif Source: https://upload.wikimedia.org/wikipedia/commons/8/8c/
Monotonic_but_nonhomomorphic_map_between_lattices.gif License: CC BY-SA 3.0 Contributors: Own work Original artist: Jochen
Burghardt
File:N-Quadrat,_gedreht.svg Source: https://upload.wikimedia.org/wikipedia/commons/1/13/N-Quadrat%2C_gedreht.svg License: CC0
Contributors: Own work Original artist: Mini-oh
File:Permutohedron.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/3e/Permutohedron.svg License: Public domain
Contributors: Own work Original artist: David Eppstein
File:Question_book-new.svg Source: https://upload.wikimedia.org/wikipedia/en/9/99/Question_book-new.svg License: Cc-by-sa-3.0
Contributors:
Created from scratch in Adobe Illustrator. Based on Image:Question book.png created by User:Equazcion Original artist:
Tkgd2007
File:Relacin_de_dependencia.svg Source: https://upload.wikimedia.org/wikipedia/commons/9/99/Relaci%C3%B3n_de_dependencia.
svg License: GFDL Contributors: Own work Original artist: Dnu72
File:Rubik{}s_cube_v3.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b6/Rubik%27s_cube_v3.svg License: CC-
BY-SA-3.0 Contributors: Image:Rubik{}s cube v2.svg Original artist: User:Booyabazooka, User:Meph666 modied by User:Niabot
File:Semiorder.svg Source: https://upload.wikimedia.org/wikipedia/commons/3/33/Semiorder.svg License: Public domain Contribu-
tors: Own work Original artist: David Eppstein
File:Series-parallel_partial_order.svg Source: https://upload.wikimedia.org/wikipedia/commons/c/c4/Series-parallel_partial_order.
svg License: Public domain Contributors: Own work Original artist: David Eppstein
File:Set_partitions_5;_matrices.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b4/Set_partitions_5%3B_matrices.
svg License: CC BY 3.0 Contributors: Own work Original artist: <a href='//commons.wikimedia.org/wiki/File:Watchduck.svg' class='image'><img
alt='Watchduck.svg' src='https://upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.svg/40px-Watchduck.svg.png' width='40'
height='46' srcset='https://upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.svg/60px-Watchduck.svg.png 1.5x, https:
//upload.wikimedia.org/wikipedia/commons/thumb/d/d8/Watchduck.svg/80px-Watchduck.svg.png 2x' data-le-width='703' data-le-
height='806' /></a> Watchduck (a.k.a. Tilman Piesk)
142 CHAPTER 43. WELL-QUASI-ORDERING

File:Strict_product_order_on_pairs_of_natural_numbers.svg Source: https://upload.wikimedia.org/wikipedia/commons/8/88/Strict_


product_order_on_pairs_of_natural_numbers.svg License: CC BY-SA 3.0 Contributors: Own work Original artist: Jochen Burghardt
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)
File:Transitive-closure.svg Source: https://upload.wikimedia.org/wikipedia/commons/6/60/Transitive-closure.svg License: GFDL Con-
tributors: Own work Original artist: Anish Bramhandkar
File:Wiki_letter_w.svg Source: https://upload.wikimedia.org/wikipedia/en/6/6c/Wiki_letter_w.svg License: Cc-by-sa-3.0 Contributors:
? Original artist: ?
File:Wiki_letter_w_cropped.svg Source: https://upload.wikimedia.org/wikipedia/commons/1/1c/Wiki_letter_w_cropped.svg License:
CC-BY-SA-3.0 Contributors: This le was derived from Wiki letter w.svg: <a href='//commons.wikimedia.org/wiki/File:Wiki_letter_w.
svg' class='image'><img alt='Wiki letter w.svg' src='https://upload.wikimedia.org/wikipedia/commons/thumb/6/6c/Wiki_letter_w.svg/
50px-Wiki_letter_w.svg.png' width='50' height='50' srcset='https://upload.wikimedia.org/wikipedia/commons/thumb/6/6c/Wiki_letter_
w.svg/75px-Wiki_letter_w.svg.png 1.5x, https://upload.wikimedia.org/wikipedia/commons/thumb/6/6c/Wiki_letter_w.svg/100px-Wiki_
letter_w.svg.png 2x' data-le-width='44' data-le-height='44' /></a>
Original artist: Derivative work by Thumperward
File:Wiktionary-logo-en-v2.svg Source: https://upload.wikimedia.org/wikipedia/commons/9/99/Wiktionary-logo-en-v2.svg License:
CC-BY-SA-3.0 Contributors: ? Original artist: ?

43.10.3 Content license


Creative Commons Attribution-Share Alike 3.0

Вам также может понравиться