Right group

From Wikipedia, the free encyclopedia

In mathematics, a right group[1][2] is an algebraic structure consisting of a set together with a binary operation that combines two elements into a third element while obeying the right group axioms. The right group axioms are similar to the group axioms, but while groups can have only one identity and any element can have only one inverse, right groups allow for multiple one-sided identity elements and multiple one-sided inverse elements.

It can be proven (theorem 1.27 in [2]) that a right group is isomorphic to the direct product of a right zero semigroup and a group, while a right abelian group[1] is the direct product of a right zero semigroup and an abelian group. Left group[1][2] and left abelian group[1] are defined in analogous way, by substituting right for left in the definitions. The rest of this article will be mostly concerned about right groups, but everything applies to left groups by doing the appropriate right/left substitutions.

Definition[edit]

A right group, originally called multiple group,[3][4] is a set with a binary operation ⋅, satisfying the following axioms:[4]

Closure
For all and in , there is an element c in such that .
Associativity
For all in , .
Left identity element
There is at least one left identity in . That is, there exists an element such that for all in . Such an element does not need to be unique.
Right inverse elements
For every in and every identity element , also in , there is at least one element in , such that . Such element is said to be the right inverse of with respect to .

Examples[edit]

Direct product of finite sets[edit]

The following example is provided by.[4] Take the group , the right zero semigroup and construct a right group as the direct product of and .

is simply the cyclic group of order 3, with as its identity, and and as the inverses of each other.

table
e a b
e e a b
a a b e
b b e a

is the right zero semigroup of order 2. Notice the each element repeats along its column, since by definition , for any and in .

table
1 2
1 1 2
2 1 2

The direct product of these two structures is defined as follows:

  • The elements of are ordered pairs such that is in and is in .
  • The operation is defined element-wise:
    Formula 1:

The elements of will look like and so on. For brevity, let's rename these as , and so on. The Cayley table of is as follows:

table
e1 a1 b1 e2 a2 b2
e1 e1 a1 b1 e2 a2 b2
a1 a1 b1 e1 a2 b2 e2
b1 b1 e1 a1 b2 e2 a2
e2 e1 a1 b1 e2 a2 b2
a2 a1 b1 e1 a2 b2 e2
b2 b1 e1 a1 b2 e2 a2

Here are some facts about :

  • has two left identities: and . Some examples:
  • Each element has two right inverses. For example, the right inverses of with regards to and are and , respectively.

Complex numbers in polar coordinates[edit]

Clifford gives a second example[4] involving complex numbers. Given two non-zero complex numbers a and b, the following operation forms a right group:

All complex numbers with modulus equal to 1 are left identities, and all complex numbers will have a right inverse with respect to any left identity.

The inner structure of this right group becomes clear when we use polar coordinates: let and , where A and B are the magnitudes and and are the arguments (angles) of a and b, respectively. (this is not the regular multiplication of complex numbers) then becomes . If we represent the magnitudes and arguments as ordered pairs, we can write this as:

Formula 2:

This right group is the direct product of a group (positive real numbers under multiplication) and a right zero semigroup induced by the real numbers. Structurally, this is identical to formula 1 above. In fact, this is how all right group operations look like when written as ordered pairs of the direct product of their factors.

Complex numbers in cartesian coordinates[edit]

If we take the and complex numbers and define an operation similar to example 2 but use cartesian instead of polar coordinates and addition instead of multiplication, we get another right group, with operation defined as follows:

, or equivalently:
Formula 3:

A practical example from computer science[edit]

Consider the following example from computer science, where a set would be implemented as a programming language type.

  • Let be the set of date times in an arbitrary programming language.
  • Let be the set of transformations equivalent to adding a duration to an element of .
  • Let be the set of time zone transformations on elements of .

Both and are subsets of , the full transformation semigroup on . behaves like a group, where there is a zero duration and every duration has an inverse duration. If we treat these transformations as right semigroup actions, behaves like a right zero semigroup, such that a time zone transformation always cancels any previous time zone transformation on a given date time.

Given any two arbitrary date times and (ignore issues regarding representation boundaries), one can find a pair of a duration and a time zone that will transform into . This composite transformation of time zone conversion and duration adding is isomorphic to the right group .

Taking the java.time package as an example,[5] the sets and would correspond to the class ZonedDateTime, the function plus and the function withZoneSameInstant, respectively. More concretely, for any ZonedDateTime t1 and t2, there is a Duration d and a ZoneId z, such that:

t2 = t1.plus(d).withZoneSomeInstant(z)

The expression above can be written more concisely using right action notation borrowed from group theory as:

It can also be verified that durations and time zones, when viewed as transformations on date/times, in addition to obeying the axioms of groups and right zero semigroups, respectively, they commute with each other. That is, for any date/time t, any duration d and any timezone z:

This is the same as saying:

References[edit]

  1. ^ a b c d Nagy, Attila (2001). Special classes of semigroups. Dordrecht: Kluwer Academic Publishers. ISBN 0-7923-6890-8. OCLC 46240335.
  2. ^ a b c Clifford, A. H. (29 June 2014). The algebraic theory of semigroups. Preston, G. B. (Reprinted with corrections ed.). Providence, Rhode Island. ISBN 978-1-4704-1234-0. OCLC 882503487.{{cite book}}: CS1 maint: location missing publisher (link)
  3. ^ Hollings, Christopher D. (2017-09-01). "'Nobody could possibly misunderstand what a group is': a study in early twentieth-century group axiomatics". Archive for History of Exact Sciences. 71 (5): 409–481. doi:10.1007/s00407-017-0193-8. ISSN 1432-0657. PMC 5573778. PMID 28912607.
  4. ^ a b c d Clifford, A. H. (1933). "A System Arising from a Weakened Set of Group Postulates". Annals of Mathematics. 34 (4): 865–871. doi:10.2307/1968703. ISSN 0003-486X. JSTOR 1968703.
  5. ^ "java.time (Java Platform SE 8 )". docs.oracle.com. Retrieved 2021-06-03.