(2006-02-21) From Magmas to Semigroups to Monoids
In a monoid, the associative internal operator has a neutral element.
Bourbaki calls magma
a set endowed with some internal well-defined operation.
(Avoid the term groupoid for this, which is now best reserved
for a concept in category theory.)
Multiplicative notations are often used where the binary operator is
understood between consecutive symbols representing elements.
Saying that an operation
is multiplicative merely stresses the use of that convention
usually supplemented by the optional use of a so-called multiplicative,
symbol, (like a dot, an x-shaped cross, a
delta or any special
ad hoc symbol)
whenever a clear separation between elements is deemed typographically appropriate.
Once a particular operator is so singled out, it's called a multiplication
and the qualifier multiplicative can then be used, especially to
distinguish that from co-existing additive concepts.
Occasionally, the multiplicative vocabulary is applied to several co-existing operators,
making distinct multiplicative symbols mandatory.
(Usage may or may not allow one such symbol to be dropped.)
For example, the dot-product and the cross-product
over 3D-vectors are both construed as multiplications
but the cross symbol can never be dropped. In some international texts,
that cross is replaced by a wedge,
which is the general symbol for an
exterior product, of which the ordinary cross-product can be construed to be a special case.
When that particular operation is not denoted by a
cross symbol, it's called a vectorial product
(Bourbaki: produit vectoriel, in French.)
Semigroups
If its operator is associative, a magma is
called a semigroup.
Associativity is the property which makes the use of parentheses optional:
x y z = (x y) z = x (y z)
The order of a
finite semigroup is its number of elements.
We count two semigroups as distinct when there's no
isomorphism between them:
Number of distinct semigroups of order n.
(A027851)
0
1
2
3
4
5
n = 6
n = 7
n = 8
n = 9
1
1
5
24
188
1915
28634
1627672
3684030417
105978177936292
A semigroup operator may or may not be commutative.
(In a commutative semigroup, xy is the same as yx
for any pair of elements x and y.)
Number of distinct commutative semigroups of order n.
(A029851)
0
1
2
3
4
5
n = 6
n = 7
n = 8
n = 9
1
1
3
12
64
405
3312
44370
2209839
623492664
The numbers of distinct noncommutative semigroups are obtained from the above two tables,
by termwise subtractions. Those numbers are always even because such
semigroups come in pairs linked by an anti-isomorphism.
Number of distinct noncommutative semigroups of order n.
0
1
2
3
4
5
n = 6
n = 7
n = 8
n = 9
0
0
2
12
124
1510
25322
1583302
3681820578
105977554443628
Lone operators are often designed to be associative.
In complex structures with several operators, non-associativity may emerge in a natural way:
For example, in the realm of
hypercomplex numbers,
the multiplication of octonions or sedenions is not associative.
Likewise, the cross-product in ordinary three-dimensional vector space is
not associative. Instead, it verifies what's called
Jacobi's identity:
Number of commutative monoids of order n.
(A058132)
0
1
2
3
4
5
n = 6
n = 7
n = 8
n = 9
0
1
2
5
19
84
509
3901
48957
2264764
The number of noncommutative monoids is obtained by subtracting the corresponding
entries of the above tables. It's always an even
number because such monoids come in pairs linked by an anti-isomorphism.
Number of noncommutative monoids of order n.
0
1
2
3
4
5
n = 6
n = 7
n = 8
n = 9
0
0
0
2
16
144
1728
27658
1620040
3683621866
A monoid can also be defined as a category with
a single object (the arrows of that category being the elements of the monoid).
(2006-03-04) Invertible Elements in a Monoid
Two flavors of invertibility, which coincide when both exist.
In a monoid, an element x is said to be
right invertible if there's a right-inverse
x' of x (which is to say that the product
xx' is unity).
It's called left invertible
if there's a left-inverse x'' (such that
x''x is unity).
When both inverses exist, they are necessarily equal
(HINT: Consider x''xx' ).
In this case, x is said to be invertible
and its (unique) inverse is denoted x-1.
Cancellativity
(cancellable element & cancellative operation)
(2006-02-21) The Free Monoid
(strings over a given alphabet)
A particular monoid where only the neutral element is invertible.
All the finite strings (or words) whose characters (letters or symbols)
are taken from a given alphabet form
a monoid under the operation of concatenation
(concatenating two strings means appending the second to [the right of] the first).
The concatenation of two strings called A and B
is best called "A before B".
The empty string
is the neutral element for concatenation.
This monoid is free from any relations equating
distinct strings of basic symbols. Hence the name
(French: monoïde libre ).
Clearly, concatenating two nonempty strings yields something other than the empty
string. The empty string is thus the only string with an inverse...
The free monoid over an alphabet of only one symbol is isomorphic to the natural integers
endowed with addition (0,1,2,3...). In every other case, a free monoid
is clearly not commutative.
(2013-01-02) Exponentiation
(for power-associative "multiplications")
Raising something to the power of an integer.
Whenever some kind of associative multiplication
is defined, something like x3 is
simply shorthand for xxx. It's always legitimate to raise
an element to the power of a positive integer that way.
The n-th power of an element can be well-defined even for non-associative
multiplications.
The weaker property of alternativity
suffices (e.g., the multiplication of octonions
is just alternative).
By definition, the weakest property of multiplication
which allows a well-defined exponentiation
(to the power of a positive integer)
is called power associativity.
Almost all "multiplications" have it.
x0 is always defined as equal to the neutral element,
if there is one (otherwise, it's undefined).
It makes no difference whether x is invertible
or not (with ordinary arithmetic, zero
to the power of zero equals one).
One enlightening example is that of the free monoids,
defined in the previous section:
xn denotes the concatenation of n strings identical to x.
Thus, x0 denotes the empty string (the concatenation of no strings).
x need not be "invertible".
If x is invertible, x-3 simply denotes
(x-1)3.
Only invertible elements can be raised to the power of a negative
integer.
Empty sums, empty products, empty intersections, etc.
Raising something to the power of zero is a special case of an empty product.
The result of not performing at all some well-defined associative operation depends
on that operation alone: It's equal to its neutral element (whenever it has one).
An empty sum is 0 (the neutral element for addition).
An empty product is 1 (the neutral element for multiplication).
A group is a set G on which an internal operation
is defined which verifies the following properties
(using multiplicative notations for the operator).
Closure :
"xÎG,
"yÎG,
x y Î G
(The product is well-defined.)
Associativity :
"xÎG,
"yÎG,
"zÎG,
(x y) z = x (y z)
A unity element (e) exists :
$eÎG,
"xÎG,
e x = x e = x
Universal Invertibility :
"xÎG,
$x'ÎG,
x x' = x' x = e
G is called a commutative group
(or abelian group) when we also have:
Commutativity (optional) :
"xÎG,
"yÎG,
x y = y x
Niels Abel (1802-1829)
considered algebraic equations whose Galois groups are commutative
(and went on to prove that they are solvable).
Leopold Kronecker (1823-1891)
called such equations Abelian and the qualifier was first applied to
all commutative groups by
Camille Jordan (1838-1922)
in 1870. The term is now so common in that capacity that it's usually lowercased (abelian).
An additive group is merely a group
(usually abelian) where additive notations
are used: The plus sign (+) denotes the group operator.
Additive notations are rarely used for a noncommutative operator.
(In a ring, addition is always commutative.
In a near-ring, it doesn't have to be.)
One well-known exception is the addition of transfinite ordinalsà la Cantor (which I dare consider a misguided effort).
Single-sided
group properties imply double-sided ones :
The double-sidedness of two of the above group axioms need not
be postulated; it can be derived from one-sided equivalents of those
axioms :
There's a right-neutral element e :
"x, x e = x
Every element is right-invertible :
"x,
$x', x x' = e
Indeed, we may compute x' x
using just those two single-sided postulates:
x' x
=
x' x e
=
x' x x' (x' )'
=
x'e (x' )'
=
x' (x' )'
=
e
That would show x' to be the inverse of x,
if we knew that e is neutral
on both sides. That fact is easy to prove, using the above as a lemma:
"xÎG,
e x = (x x' ) x = x (x' x) =
x e = x
This double-sided neutrality implies that there's only one unity e .
(HINT: Assume another unity e' and
consider e e' ).
Similarly, there's only one inverse x'
of x
(HINT: Let x" be another
and consider x' x x" ).
So we may safely talk about the inverse of x.
Note, finally, that (x' )' = x
(HINT:
x' (x' )' = e ).
(2006-02-21) Subgroups
A subgroup is a group contained in another group.
A subgroup H of a group G is a subset H of G which forms
a group under the group operation defined over G.
H is a subgroup of G if and only if it contains the
product of any element of H by the
inverse of any other element of H.
A multiplicative
subgroup is said to be stable by division.
"xÎH,
"yÎH,
x y-1 Î H
When additive nomenclature and notations are used, this translates
into the following statement, which says that a subgroup of an additive group
is merely a subset that's stable by subtraction :
"xÎH,
"yÎH,
x - y Î H
A proper
subgroup of G is a subgroup of G not equal to G itself.
The trivial group {e} has no proper subgroup.
Any intersection of subgroups is a subgroup.
The centralizer in a group G
of a subset E consists of all the elements of G which
commute with every element of E.
It is a subgroup of G.
The centralizer in G of G itself is the
center of G, denoted Z(G)
(it's the intersection of all centralizers in G).
The center is a normal subgroup of G, but
other centralizers may not be.
Elements of Z(G) are called central.
(2020-01-27) Ideals of a multiplicative
semigroupA A multiplicatively absorbent subset of A is an ideal of A.
By definition:
For a left-idealI,
the product ax is in I whenever x is:
"aÎA,
aI Í I
For a right-idealI,
the product xa is in I whenever x is:
"aÎA,
Ia Í I
Unless otherwise specified,
an ideal is both a right-ideal and a left-ideal.
Note that the empty set is an ideal of any semigroup.
This is most often used in the context of
ring theory, where an ideal of
a ring (not necessarily a unital ring)
is defined as being both a multiplicative ideal
(in the above sense) and an additivesubgroup
(thus, the empty set is not an ideal of a ring because it's not a group).
Any intersection of [one-sided] ideals is a [one-sided] ideal.
The intersection of all the ideals of a semigroup is called its minimal ideal.
If it's nonempty, the minimal ideal M of a
commutative semigroup is a group.
This is to say that M has a neutral element, even if the whole semigroup doesn't.
(2006-03-09) Generators of a Group
The smallest subgroup containing E
is said to be generated by E.
For any subset E of a group G, the intersection
of all subgroups of G
containing E is a subgroup of G,
called the subgroup generated by E.
E is said to be a set of generators of whatever subgroup it generates.
A group which is generated by a finite set is said to be finitely generated.
For example, the additive group
(,+) of the integers is generated by the set {1}.
It's also generated by {2,3} or any other pair of coprime
integers (because of Bezout's lemma).
More generally, (,+)
is generated by any set of coprime integers
(not necessarily pairwise coprime) like {6,10,15}.
A finite group
(of order n ) which is generated by a single
element is a cyclic group.
An element of such a group which generates the whole group
is called a primitive element
(or a primitive root, with the vocabulary inherited from representing
the cyclic group of order n as
the "n-th roots of unity" in complex numbers).
There are f(n) different elements
in a cyclic group which are primitive ones
( f being Euler's totient function).
The multiplicative group
(+, ´)
of positive rationals is not finitely generated.
It's generated by the prime numbers
{2,3,5,7,11,13,17,19...}.
(2016-05-29) Presentations of a group:
Generators and relators.
Describing a group using the relations obeyed by its
generators.
A finitely-generated group can be described by naming a set of
generators and
stating the nontrivial relations they obey (the relators).
Those relators are normally given by expressions which are equal
to the neutral element (minimally so)
but explicit equations are also commonly used.
A free group has no relators.
The simplest free group is isomorphic to the additive group of the integers
(,+) and has the
following multiplicative presentation, which names a single generator
and states no relators:
< a | >
Less trivially, the octic groupD4
could be presented as follows:
< r, s | r 4, s 2, srsr >
or
< r, s | r 4 = s 2 = srsr = 1 >
Do not confuse such presentations with
(linear) representations.
(2006-03-02) Cosets, Index and Lagrange's Theorem
The order of a subgroup divides the order of the group.
By definition, the order |G| of a finite group G
is its number of elements.
(The order of an element x is the order of the subgroup generated by {x}.)
Cosets :
In a group G,
the left-coset of an element x, with respect to
the subgroup H,
is the subset x H of G
(consisting of all products
x h where h is an element of H).
Similarly, the right-coset is H x.
Index of a Subgroup :
Two left-cosets with respect to H
are either disjoint or identical and they have
the same cardinality as H
(i.e., the same number of elements if finite).
Whenever it's finite,
the number of left-cosets with respect to H is equal to the number of
right-cosets.
It's denoted [G:H] and is called the index of
H in G.
Lagrange's Theorem :
In the case of a finite group G,
the fact that such left-cosets form a partition of G
shows that the order of the subgroup H divides evenly
the order of G.
This result is known as Lagrange's Theorem.
It's now presented as one of the most basic results of Group Theory,
named in honor of Joseph-Louis Lagrange (1736-1813),
who made a related remark in 1777.
The general result was probably known to Cauchy (1789-1857) but it was only formally proved in 1861,
by Camille Jordan (1838-1922; X1855).
Commensurability :
Two subgroups are said to be commensurable when the
index of their
intersection is finite in each of them.
The qualifier is inherited from ancient Greek mathematics,
where two real numbers are called commensurable when they are proportional to
two integers. The two additive groups generated
by two such numbers are indeed commensurable in the above sense
(their intersection is the additive group generated by the
lowest common multiple
of the two numbers).
(2020-05-19) Cauchy's group theorem
(Cauchy, 1845)
If a prime number p divides |G|,
then some element of G has order p.
The commutative case can be used as a lemma to prove Cauchy's theorem
and also its generalization by Sylow (Sylow's first theorem, 1872).
Lemma :Cauchy's group theorem holds for abelian groups.
Proof :
Let G be an abelian group G whose order is a multiple of the prime p:
|G| = n = m×p.
First, we see that the proposition is true if G is cyclic,
generated by element a (since the element am is then of order p).
Otherwise, we proceed by induction on m,
starting with the case m = 1 which makes p the order of G.
This is trivial because, by Lagrange's theorem,
the order of an element must divide the order of the group and can thus
only be 1 or p. In other words, any element of G besides identity is a satisfactory
element of order p (which establishes also that G is cyclic).
For m ≥ 2, consider an element h of G, besides identity.
Let H be the nontrivial subgroup generated by h.
H is a normal subgroup (as is any subgroup in the abelian case).
Both H and G/H are nontrivial group of order strictly less than n
(because we've already disposed of the cases where G is cyclic).
Since the product of their orders
(respectively |G| and [G:H])
equals n = m×p, at least one of them is divisible by p.
In either case, the induction hypothesis implies that the corresponding group
contains an element of order p. Either way, we can use that to obtain
an element of order p in G, as follows:
If H contains an element x of order p, then x is also in G and we're done.
If G/H contains an element y of order p, then y is the class xH of some
element x of G. For any integer k, the class of xk is yk.
So, the order of x is the same as the order of y, namely p.
This concludes the proof that Cauchy's theorem holds for abelian groups.
(2020-05-19) Sylow's Theorems [pronounced see-lov ] (1872)
On the number of subgroups of given order in a finite group G,
For a prime p,
a p-group is a group where the order
of any element is a power of p.
If it's a subgroup of G, it's called a p-subgroup of G.
In a finite group G, a Sylow p-subgroup (abbreviated p-SSG) is a
maximal p-subgroup of G.
The set of all p-SSG is denoted Sylp (G).
Remarkably, all of those are isomorphic to each other.
To a normal subgroup H corresponds an
equivalence relation among
elements of G defined by calling x and y equivalent when
xy-1 is in H (in other words,
when x and y have the same
left cosets with respect to H).
The equivalence classes so defined
form a group denoted G/H and
called the quotient
of H in G
(or of G by H) also dubbed "G modulo H".
Although the above equivalence relation is defined for any
subgroup H, the equivalence classes form a group
only when H is normal.
Examples of Normal Subgroups :
Any group G is a normal subgroup of itself
(the only non-proper one).
The trivial group {e} is a normal subgroup of
any group G whose neutral element is e.
(It's a proper
subgroup of any such G but itself.)
The center Z(G) of a group G consists of all the elements
which commute with every element G.
A member of Z(G) is called a central element.
A noncentral element is an element which
doesn't commute with at least one other element.
The center is a normal subgroup.
So is any subgroup of the center
(in particular, any subgroup of an abelian group
is normal).
If f is a homomorphism
or an antihomomorphism
from group G, then the kernel
of f (ker f )
is a normal subgroup of G.
More generally,
so is the inverse image (pre-image)
of any normal subgroup of f (G).
For a normal subgroup H of G,
the direct image f (H)
is a normal subgroup of f (G).
For any subset E of the group G, the subgroup generated by
all the conjugates of the elements of E
is called conjugate closure of E.
It's a normal subgroup containing E.
In fact, it's the smallest normal subgroup containing E (i.e, it's the intersection
of all normal subgroups containing E). It's thus also
known as the normal closure of E.
Any Subgroup is a Normal Subgroup of its Normalizer :
The normalizer of a subgroup H
consists of all elements x of the group G
for which x H = H x
(in particular all elements of H
belong to its normalizer). The normalizer of H
is a subgroup of G.
By definition, H is a normal subgroup
of its normalizer
(H need not be a normal subgroup of the whole group G).
(2019-04-13) Wielandt Symbols (Helmut Wielandt, c. 1960)
Notations for [proper] normal subgroups (or ideals).
Two conventions are floating around to distinguish between a standard (reflexive)
ordering relation and its strict (antireflexive) counterpart:
The highlighted entries may be ambiguous.
I don't recommend the grey ones.
The above notations for normal subgroups were introduced by
Helmut Wielandt around 1960.
They are now also used to denote ideals in
ring theory
(since an ideal is to a ring what a normal subgroup is to a group).
(2006-04-05) Homomorphisms and Anti-homomorphisms
Functions for which the image of a product is a product of the images.
An homomorphism is a map
(or function)
which preserves some specific algebraic operation(s).
A group homomorphism is thus a map f
from a [multiplicative]
group G into another group H, which is such that:
"xÎG,
"yÎG,
f (x y) = f (x) f (y)
If f is surjective
("onto" H) it's called an epihomomorphism
(or "homomorphism onto").
If it's injective
("one-to-one") it's called an monomorphism.
If it's bijective
("one-to-one onto") it's an isomorphism.
An homomorphism from G to itself is called an
endomorphism of G.
A bijective endomorphism is called an automorphism.
The automorphisms of a group G form a group, denoted Aut(G).
Anti-homomorphisms :
An anti-homomorphism, with respect to a
multiplicative operator, is a function f
which reverses the order of that multiplication :
"xÎG,
"yÎG,
f (x y) = f (y) f (x)
In any group, inversion is an example of an anti-homomorphism:
( x y ) -1 = y -1 x -1
The concepts defined above for homomorphisms have their counterparts for anti-homomorphisms:
Anti-epihomomorphism, anti-monomomorphism,
anti-isomorphism, anti-endomorphism and anti-automorphism.
Kernel (French: noyau )
For a homomorphism (or an anti-homomorphism) f from
group G to a group of identity e,
the kernel of
f is a normal subgroup of G defined by
ker f =
{ xÎG | f (x) = e }
(The homomorphic pre-image
of any normal subgroup is normal.)
(2006-03-05) Sym(E) = Symmetric Group on E
(Gersonides, 1321)
The group of the permutations of E
(bijections of the set E onto itself).
A permutation of E is a one-to-one correspondence
(bijection) of E onto itself.
The term is most commonly used when E is finite,
but it's also acceptable when E is infinite
(possibly uncountably so).
In the finite case, the symmetric group of degree n
is denoted Sn.
Its order is the number of permutations of n elements,
namely n!
("n factorial").
The largest order of an element in the symmetric
group Sn is traditionally denoted
g(n) where g is called
Landau's function,
in honor of the German mathematician
Edmund Landau (1877-1938)
who proved the following asymptotic
equivalence in 1902:
Log g(n) ~ (n Log n)½
Even permutations
form the alternating groupAn
(whose order is n!/2 ).
It's the derived subgroup of the symmetric group:
An = S'n
An even permutation is obtained by an even number of switches
(swaps of two elements).
The parity, or signature, of
a finite permutation may be determined by counting its
number of inversions.
Notations for small permutations :
One standard way to record computations in the realm of very small
finite groups is to use a string of different characters (digits or letters)
to denote the permutation which transforms the sorted elements in the top row into
the matching elements of the bottom row. Both row are placed between parentheses.
Juxtaposition of two such notation indicates the composition
of the functions so denoted, with the usual convention that the rightmost
function is to be applied first (composition isn't commutative):
f o g
g o f
(
1234 1243
)(
1234 1324
)
=
(
1234 1423
)
(
1234 1324
)(
1234 1243
)
=
(
1234 1342
)
Cycle decomposition of a permutations :
A cyclic permutation of n elements is denoted by a sequence
between parentheses. The image of an element is the element to its right
(the last element is mapped back to the first one). There are n
equivalent ways to denote such a permutation, since there's a free choice of
which element is written first in the list. Equivalent notations are equated:
(1 4 5 3 2) = (4 5 3 2 1)
A cycle is a permutation of n elements which is a
cyclic permutation of m of those elements (m ≤ n)
which leaves the others unchanged. Two or more cycles are said to be disjoint
when operate on different elements (each cycle applies only to elements which are left
unchanged by the others).
Any permutation can be decomposed as a composition of disjoint cycles in a unique way
(up to the order of those cycles, which is irrelevant since disjoint cycles commute).
Our previous example
entail cycles which do not commute because they're not disjoints. namely:
(3 4) (2 3) = (2 4 3)
(2 3) (3 4) = (2 3 4)
A cycle of order 2 (a 2-cycle) is called a switch or
a transposition.
It's useful to know that the signature
of a cycle of order n is (-1)n+1.
Arthur Cayley (1821-1895)
observed that a group G
is always isomorphic to a subgroup of Sym(G).
Proof :
In the multiplicative group G,
we associate to an element a
the bijection T(a) which sends an element x to
ax .
T is an injectivehomomorphism (i.e., a monomorphism)
from G to Sym(G),
which is called the regular representation of G.
T(a) o T(b) = T(a b)
So, any
finite group of order n is isomorphic to a
subgroup of Sn .
(2006-03-02) Inn(G):
The Group of Inner Automorphisms on G
An inner automorphism is a conjugation by a given element of G.
To any element a of G is associated
a special type of automorphism,
called an inner automorphism
(French: automorphisme intérieur )
defined as follows
( fa is called conjugation by a ).
" x,
fa(x) = a x a-1
[ Note that fa o fb
= fab ]
Under function composition, inner automorphisms form
a normal subgroup (see proof later in this section)
denoted Inn(G), of the group of the automorphisms on G, denoted Aut(G)
(itself a subgroup of Sym(G),
the symmetric group on G).
Conjugation by a
is the identity function just if a belongs
to the center of G. Consequently:
Inn(G) is isomorphic to the quotient of
G by its center.
Note that a subgroup H of G which is mapped onto itself by
any inner automorphism
is a normal subgroup (also
called invariant subgroup).
More generally, two subgroups of G are said to be conjugates
of each other when there is an inner isomorphism between them.
The above claim that
Inn(G) is a normal subgroup of Aut(G)
is established by showing that
conjugation by any automorphism g of an inner
automorphism (conjugation by a)
yields another inner automorphism. That can be proved in a single line:
" x,
g o fa o g-1 (x)
=
g ( a g-1(x) a-1 )
=
g(a) x g(a)-1
=
fg(a) (x)
(2006-03-02) Out(G):
The Outer Automorphism Group
The members of Out(G) are classes of automorphisms of G.
A group G is said to be centerless
when its center is trivial, which is to say
that only the identity element commutes with every element.
A complete group is a centerless group whose only automorphisms are
the inner ones. (Equivalently, it's a group whose center and
outer automorphism group are trivial.)
If a group G is complete, it's isomorphic to Aut(G)
(its automorphisms).
However, the converse need not be true (one counterexample is D4 ).
(2006-03-20) Conjugates and the
Conjugacy Class Formula The conjugacy classes of a group G form a partition of G.
Two elements x and y of a group G are said to be
conjugates
when there's an inner automorphism from one to the other,
that is, when there's an element a of G
such that ax = ya.
So defined, conjugacy is an
equivalence relation (it's reflexive, symmetric and transitive).
The conjugacy class of an element x is the set of
all elements of G which are conjugate to it.
Every element is in one and only one of those classes
(equivalence classes always form such a partition).
If x is in the center of G,
denoted Z(G), then the conjugacy class of x is
simply {x} (a set of only one element).
More generally, we would establish that the number of elements that are conjugate
to x is equal to the index in G
of the centralizer C of {x}.
That number is usually denoted [ G : C ].
Tallying the conjugacy classes with more than one element by assigning each
a different index i, we obtain the so-called
conjugacy class formula :
| G | = | Z(G) | +
å i
[ G : Ci ]
The second term is an empty sum (equal to zero) when G is commutative.
(2006-03-05) Simple Groups
A group is simple
when it has just twonormal subgroups.
{e} and G are trivially always normal
subgroups of G.
The group G is said to be simple when its
only normal subgroups are those two.
Simple groups (8:52)
by Liliana De Castro (Socratica, 2018-01-10).
(2006-03-06) Derived Subgroup G'
(or "Commutator Subgroup")
G', G(1) or [G,G]
is the subgroup of G generated by its commutators.
The commutator [x,y] of two elements of the
multiplicative group G is:
[x,y] = x y x-1 y-1
= x y (y x)-1
The set of all commutators isn't necessarily a subgroup.
What's called the derived subgroup
(or commutator subgroup)
is the subgroup they generate
(i.e., the smallest subgroup which includes all commutators).
A group is said to be perfect
when it's equal to its derived subgroup.
In particular, a group which contains only commutators is perfect.
That's so for all finite non-abelian simple groups,
as was first conjectured by
Øystein Ore (1899-1968) in 1951.
Ore's conjecture
was proved in 2010, using the
classification of finite simple groups.
The derived subgroup of a group is a normal subgroup,
as the following identity demonstrates (since the set of commutators is thus shown
to be stable under anyinner automorphism,
so is the subgroup they generate).
a [x,y] a-1 =
[ axa-1, aya-1 ]
G' is also the smallest normal subgroup
of G whose quotient group in G
is abelian (i.e., commutative). The group
G/G' is known as the abelianization of G
(it's the largest abelian quotient in G).
Examples of Derived Subgroups :
The derived subgroup of any abelian group is the
trivial subgroup.
The derived subgroup of the symmetric groupSn is the alternating groupAn.
The derived subgroup of the alternating group is itself:
A'n = An.
(2006-03-21) Direct Product (or Direct Sum)
The group made from the independent juxtaposition of several groups.
The direct product of two groups G and H is the
group obtained by using for the cartesian product
G ´ H
independent operations on the components:
(g,h) (g',h') = ( g h , g'h' )
The term direct sum is used for the same concept with
additive notations:
(g,h) + (g',h') = ( g+h , g'+h' )
Similar rules can be used for cartesian products of any number of monoids.
Extensions to infinitely many components :
The concept extends naturally to direct sums (or
direct products ) of infinitely many monoids.
Such direct sums are usually understood to be finitely restricted
(by considering just the elements having only a finite number of components
that differ from the relevant neutral element).
This assumption is always made in the case of vector spaces
(only finitely many components are nonzero in the resulting structure)
and it's prudent to clearly distinguish between the two possibilities for infinite
cartesian products endowed with component-wise operations.
For example, the fundamental theorem of arithmetic
provides a standard isomorphism between the multiplicative monoid of the positive integers
and the finitely restricted direct sum of infinitely many copies of the nonnegative
integers (each such copy being associated with a prime number).
Using standard notations,
this can be expressed as:
( *
, ´ ) =
(
()
, + )
Note that the set appearing in the right-hand-side of the above is countable,
because of the parenthesized exponent which indicates a finite restriction in the
above sense. A lack of parentheses around the exponent would denote
an uncountable set
which is rarely investigated, if ever
(that beast includes elements idenfified with products of infinitely many
coprime integers).
(2016-01-10) Fundamental Theorem for Finite Abelian Groups
Finite abelian groups are either cyclic or direct sums.
Thus, if n is the k-th power of a prime,
the number of non-isomorphic abelian groups is equal to
the number p(k) of partitions of k.
More generally,
if the prime factorization of n is
q1k1 q2k2... qmkm
then the number of non-isomorphic abelian groups of order n is equal to:
For what n are there one million abelian groups of order n ?
By trying only the first 61,
we see that the only partition numbers which divide 1000000 are
p(1) = 1, p(2) = 2 and p(4) = 5.
Therefore, there are exactly 1000000 distinct abelian groups of order n
if and only if the factorization of n consists of:
6 primes of multiplicitity 4.
6 primes of multiplicitity 2.
Any number of primes of multiplicitity 1 (possibly none).
(2014-12-21) Hol(G) : Holomorph group of the group G.
A semi-direct product of G and
its group of automorphisms Aut(G).
If f is a homomorphism from a group
H to Aut(G), the semi-direct product
of G and H with respect to f is the group denoted
G ´f H consisting of
the cartesian product
G ´ H with the
multiplication :
(x,a) (y,b) =
( x f (a) (y) , ab )
When f is the trivial homomorphism (i.e.,
f (a) is the identity of G for any a)
this semi-direct product is just the
direct product of G and H.
Holomorph :
When H is equal to Aut(G) we may use the identity of Aut(G)
as the homorphism f appearing in the above definition and define the
holomorph Hol(G) as the semi-direct product
of G and Aut(G) in which :
(2006-03-05) Some Finite Groups
Groups of small orders and their families...
Additive notations
(using the symbol "+" for the operator)
are often used for commutative groups (abelian groups).
Groups isomorphic to the group
Cn = (/n, +)
of residues modulo n are
called cyclic groups.
The above is a special case of the notation A/I which denotes
a ring obtained as a quotient of a ring by one of its ideals.
As such it's a structure endowed with two operators.
The single-operator additive group of that structure is properly
denoted (A/I,+).
The expression (A/I,+,x) is pleonastic.
The notation
Zp is for something else.
The smallest noncyclic groups are thus of order 4 and 6.
The Klein group is the noncyclic group of order 4.
The smallest noncommutative group is the following group
S3 = D3
(the 6 symmetries of an equilateral triangle).
Klein Group
+
0
1
2
3
0
0
1
2
3
1
1
0
3
2
2
2
3
0
1
3
3
2
1
0
Dihedral Group 
D3
A
B
C
D
E
F
A
A
B
C
D
E
F
B
B
C
A
E
F
D
C
C
A
B
F
D
E
D
D
F
E
A
C
B
E
E
D
F
B
A
C
F
F
E
D
C
B
A
The Klein Group (V)
is isomorphic to the
direct sumC2 ´ C2 Felix Klein
called it Vierergruppe
in 1884.
The dihedral groupDn consists of the 2n symmetries of a regular n-gon
(n rotations, n flips).
When he proposed the cyclic structure of benzene in 1865,
August Kekulé (1829-1896)
thought that the C6H6 molecule had trigonal
symmetry (expressed by the order-6 group D3 tabulated above)
because of his vision that single and double bonds were alternating along the
carbon ring. The currently accepted symmetry for the benzene molecule
is the hexagonal group D6 (of order 12)
with 3 of the binding electrons in a delocalized orbital covering the whole ring.
There are 5 groups of order 8. Three are abelian :
C8 and the two direct sumsC2+C4 and
C2+C2+C2
(the additive group of the field of order 8).
The other two groups of order 8 are noncommutative,
namely the dihedral group D4 (the symmetries of a square)
and the quaternion groupQ8 :
(2006-03-05) Quaternion Group Q8
& Quaternions (Hamilton, 1843)
On October 16, 1843, the fundamental equations below
(which imply the given multiplication table)
occurred at once to Hamilton
as he was crossing Brougham Bridge
(Broom Bridge)
in Dublin.
He carved them into the stone of the bridge (the original carving is gone
but a plaque
celebrates this act of "mathematical vandalism").
i 2 =
j 2 =
k 2 =
i j k = -1
Quaternion GroupQ8
1
i
j
k
-1
-i
-j
-k
1
1
i
j
k
-1
-i
-j
-k
i
i
-1
k
-j
-i
1
-k
j
j
j
-k
-1
i
-j
k
1
-i
k
k
j
-i
-1
-k
-j
i
1
-1
-1
-i
-j
-k
1
i
j
k
-i
-i
1
-k
j
i
-1
k
-j
-j
-j
k
1
-i
j
-k
-1
i
-k
-k
-j
i
1
k
j
-i
-1
Red (i) and Blue (j) generators ofQ8
The real line combined with an oriented
3-dimensional Euclidean space of orthonormal basis (i,j,k)
forms the quaternions, a 4-dimensional
normed division algebra
similar to 2-dimensional
complex numbers, except multiplication
is not commutative:
(a,A) + (b,B)
=
( a+b , A+B )
(a,A) (b,B)
=
( ab - A.B , aB + bA +
A´B )
This is how the 3-dimensional "dot product" and "cross product"
were invented,
well before the generalized idea of a vector
became commonplace.
The above quaternionic units can be used to build a
Dirac operator D
(yielding the opposite of the Laplacian
D when applied twice):
D =
i ¶ / ¶x +
j ¶ / ¶y +
k ¶ / ¶z
The Laplacian remains the same in two systems of coordinates (a.ka. reference frames)
obtained from each other by rigid rotation.
(2023-03-10) Gamma group (order 32) and spin group (order 16).
The law for multiplication in the algebra generated
by the Dirac matrices.
The multiplicative group generated by the four gamma matrices a,b,c,d is of order 32.
It consists of 16 disjoint pairs of elements which are (additive) opposites of each other.
With one element of each such pairs, we form a basis for the
algebra of dimension 16 generated by the 4 gamma matrices.
The group clearly contains the identity matrix of dimension four (I) and the product e = abcd.
Introducing e allows every element of the group to be uniquely represented
by a sign (+ 1 or -1) along with a signed product of at most two factors among the five elementary elements
a,b,c,d,e. With zero such factors, we have 2 elements (+I and -I), 
with one factor we have 10 elements (a,b,c,d,e and their opposites) and with 2 distinct factors,
we obtain 20 = 2×C (5,2) elements. The grand total is indeed 32.
Among the 32 elements of the gamma group, we find:
1 element of order 1, namely I.
15 elements of order 2: -I, a, -a, ab, -ab, ac, -ac, ad, -ad, ae, -ae.
20 elements of order 4, whose square is -I.
Products of gamma matrices :
a = g0
b = g1
c = g2
d = g3 (with e = abcd)
0
Grade 1
Grade 2 (bivectors)
Grade 3
4
I
a
b
c
d
ab
ac
ad
bc
bd
cd
ae
be
ce
de
e
a
I
ab
ac
ad
b
c
d
de
-ce
be
e
cd
-bd
bc
ae
b
-ab
-I
bc
bd
a
-de
ce
-c
-d
ae
-cd
-e
-ad
ac
be
c
-ac
-bc
-I
cd
de
a
-be
b
-ae
-d
bd
ad
-e
-ab
ce
d
-ad
-bd
-cd
-I
-ce
be
a
ae
b
c
-bc
-ac
ab
-e
de
ab
-b
-a
de
-ce
I
-bc
-bd
-ac
-ad
e
-be
-ae
-d
c
cd
ac
-c
-de
-a
be
bc
I
-cd
ab
-e
-ad
-ce
d
-ae
-b
-bd
ad
-d
ce
-be
-a
bd
cd
I
e
ab
ac
-de
-c
b
-ae
bc
bc
de
c
-b
ae
ac
-ab
e
-I
cd
-bd
-d
ce
-be
-a
-ad
bd
-ce
d
-ae
-b
ad
-e
-ab
-cd
-I
bc
c
de
a
-be
ac
cd
be
ae
d
-c
e
ad
-ac
bd
-bc
-I
-b
-a
de
-ce
-ab
ae
-e
-cd
bd
-bc
be
ce
de
-d
c
-b
I
ab
ac
ad
-a
be
cd
e
ad
-ac
ae
d
-c
-ce
-de
-a
-ab
-I
bc
bd
-b
ce
-bd
-ad
e
ab
-d
ae
b
be
a
-de
-ac
-bc
-I
cd
-c
de
bc
ac
-ab
e
c
-b
ae
-a
be
ce
-ad
-bd
-cd
-I
-d
e
-ae
-be
-ce
-de
cd
-bd
bc
-ad
ac
-ab
a
b
c
d
-I
This algebra of dimension 16 is known as the spacetime algebra Cl (1,3)
which is just the
Clifford algebra of dimension 4 with Minkowski metric.
Pseudoscalars (Grade 4) commute only with scalars or bivectors (Grade 2).
The bivectors and the scalars form the centralizer of the pseudoscalars.
The above table doesn't depend on Dirac's representation of a,b,c,d in terms of 4×4 matrices.
It can be entirely constructed from the pairwise anticommutativity of a,b,c,d and the following relations.
Therefore, an isomorphic group is entirely specified by a choice of a basis of four mutually anticommutative elements verifying these:
a2 = I ,
b2 = c2 = d2 = -I
and the definition e = abcd
The last relation implies that e2 = -I and also that e anticommutes with the other four.
Thus, a is special (it's the only single-letter element which squares to unity) but
b,c,d,e are placed on an equal footing.
Here's one remarkable identity:
det ( ta + xb + yc + zd + ue) = ( t 2 - x 2 -
y 2 - z 2 - u 2) 2
Enumeratimg the automorphisms of Dirac's gamma group :
Let f be an automorphism. f (a) must be of order 2 and is therefore,
up to a change of sign, an element of {a,ab,ac,ad,ae}.
For the three distinct images of b,c,d to anticommute with f (a) and with each other,
they must belong to {b,c,d,e} up to sign. Conversely, if those conditions are met,
the images of a,b,c,d generate the whole group, as the above table can be constructed using
only the rules for combining single letters. Thus, we have 5 choices
for f (a) and 4×3×2 choices for the other three letters,
knowing that we may then pick any choice of four signs among 16 possibilities.
Therefore:
Dirac's Gamma group has 5! 24 = 120 × 16 = 1920 automorphisms.
(2014-12-17) D4 : The fourth dihedral group (8 elements)
The octic group is represented by the eight symmetries of a square.
This is a centerless
group G isomorphic to Aut(G) but not to Inn(G).
A nice example of an incomplete group isomorphic to its automorphisms.
The dihedral group D4
can be represented as the group of the 8 symmetries of a square,
with vertices numbered clockwise 1,2,3,4.
It's generated by :
A = e
B = r
C = r2
D = r3
E = s
F = s r = r3 s
G = s r2 = r2 s
H = s r3 = r s
A is the identity.
C is the half-turn.
B and D are quarter-turns.
E and G are diagonal flips.
F and H are side flips.
Swapping an even number of the above pairs
yields one of the inner automorphisms
tabulated at right.
The 4 inner automorphisms 
are allevenpermutations :
A
B
C
D
E
F
G
H
fA = fC
A
B
C
D
E
F
G
H
fB = fD
A
B
C
D
G
H
E
F
fE = fG
A
D
C
B
E
H
G
F
fF = fH
A
D
C
B
G
F
E
H
If there was an automorphism swapping an odd number of the three
pairs (B,D), (E,G) and (F,H) then
we could combine it with one of the four inner automorphisms to
obtain some automorphism f leaving (A,B,C,D) invariant and
swapping either (E,G) or (F,H). Neither is possible, since:
If f only swaps E and G, then
f (B) f (F) = B F = E
¹ f (B F) = G
If f only swaps F and H, then
f (B) f (E) = B E = H
¹ f (B E) = F
Therefore, any other automorphism must involve sending at least
one element of the three aforementioned pairs to an element of another.
Any automorphism must leave invariant A (the identity) and C
(the only other element with a square root).
Likewise, the order-4 elements, B and D, must be invariant or transform into each other.
Aut (D4 ) ,
the group of automorphisms of D4 , is isomorphic to
D4.
One of the 8 isomorphisms betweenD4and Aut (D4 )
D4
Aut (D4 )
Inn (D4 )
A
1234
e
ABCDEFGH
a
fA = fC
B
2341
r
ABCDFGHE
b
C
3412
r2
ABCDGHEF
c
fB = fD
D
4123
r3
ABCDHEFG
d
1
E
1432
s
ADCBEHGF
e
fE = fG
2
F
2143
sr = r3s
ADCBHGFE
f
3
G
3214
sr2 = r2s
ADCBGFEH
g
fF = fH
4
H
4321
sr3 = rs
ADCBFEHG
h
Column 1 gives the correspondence between the square's vertices and the order-2 automorphisms,
which directly sends column 3 to column 5 (adjusting for even parity by swapping B and D if needed).
(2006-05-09) Enumeration of Groups of Small Order
The number g(n) of different groups of order n (up to isomorphism).
If the integer n is coprime with its
Euler totient
f(n), then there's only one group of order n
(the cyclic group). This applies to the following values of n:
1, 2, 3, 5, 7, 11, 13, 15, 17, 19, 23, 29, 31, 33, 35, 37, 41, 43, 47, 51...
(A003277).
This result is attributed to
William
Burnside (1852-1927) and those numbers are known as cyclic numbers.
A divisor of a
Carmichael number
is necessarily odd and cyclic.
Around 1980, I made the conjecture that the converse is true; any odd cyclic
number seems to divide at least one Carmichael number
(if the conjecture is true, it divides infinitely many of them,
since a cyclic number has infinitely many cyclic multiples).
In 2007, Joe Crump and myself proved this to hold for cyclic numbers
below 10000.
We're not attempting to gather more numerical evidence at this time...
For noncyclic orders
(A060679)
here's the number of distinct groups:
g(n) = 2 if n is either
the square of a prime or a squarefree number with
only one of its prime factors congruent to 1
modulo another
(A054395).
The following table gives, for each m,
the numbers n for which g(n) = m.
Numbers n for which there are
precisely m groups of order
n
The classification of noncommutative
finite simple groups is much tougher...
Arguably, the final classification effort started with the 1963 publication of
a 255-page proof of the Odd Order Theorem
(or Feit-Thompson theorem)
which implies that all noncommutative simple finite groups are of even order:
Solvability of Groups of Odd Order by
John
G. Thompson (1932-) and
Walter Feit
(1930-2004). Pacific Journal of Mathematics13
(1963) 775-1029.
The classification was declared complete in 1982,
despite pending gaps...
This was the result of a tremendous collective effort, spanning decades.
A key figure in this accomplishment was
Daniel
Gorenstein (1923-1992).
The Classification Theorem :
Unless it's one of the 27
sporadic groups presented below
(including the Tits Group,
often dubiously tallied with twisted Chevalley groups)
a finite simple group
must belong to one of the following 18 countable families:
16 types of Chevalley groups, listed below, each
uniformly described in terms of a finite field
of order q (q being the power of a prime).
For example, the first such type consists of the projective
group of square matrices of dimension n+1 with coefficients in Fq :
An(q) = PSL (n+1, Fq )
Simple Chevalley Groups ( u Ù v
denotes the GCD of u and v)
Chevalley groups are named after Claude
Chevalley (1909-1984) who was the youngest founder of the
Bourbaki group in 1935.
In 1955, Chevalley found a uniform way to describe
Lie groups over arbitrary fields.
With finite fields,
this led to what J.H. Conway (1937-2020)
and others have called untwisted Chevalley groups
(they're listed first in the above table, with unsuperscripted symbols).
The twisted Chevalley groups (denoted by superscripted symbols) result
from two modifications of Chevalley's approach. One was proposed in 1959 by
Robert Steinberg (1922-2014).
The other (1960-1961)
is due to
Michio Suzuki (1926-1998) and
Rimhak Ree (1922-2005).
The above highlighted entry 2F4(2 2m+1 )
is simple only for positive values of m.
For m=0, this group is not simple but it has a simple normal subgroup
of index 2 and order 17971200
(its derived subgroup) which is known as the Tits Group,
and is best classified among sporadic groups.
Simple Groups (8:52)
by Liliana De Castro (Socratica, 2017-01-10).
(2006-03-06) The 26 or 27 Sporadic Groups
Noncommutative non-alternating finite simple groups not of Lie type.
20 of these are related to the largest and most famous of them all, the Fischer-Griess Monster.
Six other sporadic groups ( highlighted )
unrelated to the Monster are known as oddments or pariahs.
The 27th sporadic group is, arguably, the
aforementionedTits Group.
The Mathieu group M21 doesn't belong to the above list.
It's simple but can't be considered sporadic
because it's isomorphic to PSL(3,4):
M21 = PSL(3,4) = PSL(3,F4 ) = A2 (4)
The Fischer-Griess Monster Group
is also known as Fischer's Monster or
the Monster Group.
It was predicted independently by Bernd Fischer and Robert L. Griess in 1973.
At first, Griess dubbed it the Friendly Giant
and constructed it explicitely in 1981,
as the automorphism group of a 196883-dimensional commutative nonassociative
algebra over the rational numbers.
The Leech Lattice is
the densest packing of 24-dimensional hyperspheres
(each touches 196560 others). Its automorphisms feature
a center of order two.
Modulo that center, they form
the Conway Group (Co1).
Simon P. Norton gave a construction of the group proposed by
Koichiro
Harada (now called the Harada-Norton group).
Norton also proposed the monstruous moonshine
conjecture with his advisor,
John H. Conway.
The Higman-Sims Group (HS) is named after Donald G. Higman and Charles C. Sims,
who described it jointly in 1968. It's a subgroup of index 2
in the group of automorphisms of the Higman-Sims graph
(the strongly-regular graph with 100 nodes of degree 22, where adjacent nodes
have no common neighbors and nonadjacent nodes have 6 common neighbors).
The Hall-Janko Group (HJ)
is named after Marshall Hall,
Jr. (1910-1990) and
Zvonimir Janko (1932-).
It's a subgroup of index 2
in the automorphisms of the Hall-Wales graph
constructed in 1968 by
David Wales,
as the strongly-regular graph with 100 nodes of degree 36, where adjacent nodes have 14 common neighbors
and nonadjacent nodes have 12 (also called Hall-Janko graph).
The modern quest for a complete list of sporadic groups was launched by
the discovery of the first of the
Janko Groups
(J1) by Zvonimir Janko, in 1965.
The first sporadic groups
(M11 , M12 , M22 ,
M23 , M24 )
are subgroups of M24 discovered between 1860 and 1873
by Emile Mathieu (1835-1890; X1854).
Georg Frobenius (1849-1917)
proved M12 to be simple in 1904.
(2017-08-02) Torsion T = Tor(G) of an infinite group G
T is the set of all elements of G which have a finite order.
An element of finite order is called a torsion element.
If the identity is the only such element, the group G
is said to be torsion-free.
A torsion element whose order divides k is called a k-torsion.
On the other hand, a torsion group (also called a periodic group)
is a group consisting only of torsion elements (which is to say that all elements have finite orders).
All finite groups are periodic (i.e., Tor(G) = G).
If the orders of the elements in a periodic group are bounded,
then they have a
least common multiple n and the group
is said to be of exponent n.
One example of an infinite finitely-generated torsion group was given in 1964,
by Evgeny Golod (1935-2018)
and Igor Shafarevich (1923-2017).
(2015-05-03) Linear representations of a group G :
Homomorphisms from G into a group of matrices.
GL(n,K) is the group of invertible n by n matrices with entries
in a fieldK.
All finite groups are linear.
Compact groups...
Lie groups...
Faithful representations (isomorphisms).
Irreducible representations do not allow any nontrivial
proper invariant subspace.
(2023-04-04) Lie Groups.
Groups which are also smooth manifolds (locally Euclidean).
The tangent space to a Lie group is a Lie algebra.
The converse is true in finitely many dimensions but there are Lie algebras with infinitely
many dimensions which cannot be realized as the tangent space to a Lie group.
The earliest counterexample
is due to the bourbakistAdrien Douady (1935-2006).
(2006-03-01) Classical Groups
(multiplicative subgroups of matrices)
Groups of transformations depending on parameters in a field.
The classical groups tabulated below are subgroups of the group GL(n,K)
of invertible n by n matrices with entries in the fieldK.
When K isn't specifed, the field
of real numbers (R) is understood, except that
the field of complex numbers (C) underlies the groups denoted
U(n) and SU(n) (note, however, that the "dimension"
listed is always the real dimension, which is twice the complex
dimension whenever applicable).
A subgroup of GL(n,K) is called a linear representation
(or simply a representation) of any group it happens to be
isomorphic to.
A* denotes the adjoint of the square
matrix A (namely, the "conjugate transpose"
of a complex matrix, or simply the transpose of a real matrix).
A matrix is said to be unimodular if its
determinant is 1.
In the symbol of a group, the letter "S" (for special)
says that its elements are unimodular.
General linear group of Cn .
Nonsingular complex matrices ( det(A) ¹ 0 ).
SL(n,C)
2n2-2
Special linear group
of Cn .
Unimodular complex matrices ( det(A) = 1 ).
U(n) O(n,C)
n2
Unitary group (of Cn ).
Unitary matrices ( A A* = 1 )
SU(n) SO(n,C)
n2-1
Special unitary group
(of Cn ).
Unitary unimodular matrices ( A A* = 1 , det(A) = 1 )
Sp(2n,C)
2n(4n+1)
Symplectic complex group
(of Cn ).
??? ( A W A* = W )
Z(n) Z(n,C)
1 2
Scalar group.
Nonzero scalar multiples of the identity matrix
( A = a 1 )
SZ(n,C) SZ(n,K)
0
Unimodular scalar group.
The finite group formed by all the
"nth roots of unity".
PGL(n) PGL(n,C)
n2-1 2n2-2
Projective
linear group. PGL(n,C)
= GL(n,C) / Z(n,C)
PSL(n,C)
2n2-2
Projective special linear group. PSL(n,C) = SL(n,C) / SZ(n,C)
Alternate Notations :
A notation like GL(Kn) may also be used instead of
GL(n,K). This has the great advantage of being consistent
with more general symbols like GL(V) which apply to a
vector spaceV
whose dimension may be infinite.
On the other hand, when a finite field
is used, GL(n,GF(q)) may be denoted GL(n,q).
A similar convention holds for all the symbols tabulated above.
For example,
the first type of Chevalley groups is
PSL(n,q) = An(q).
There's no risk of confusion with notations like
O(3,1) as used below, which refer to a real
vector space metrically endowed with 3 spacelike dimensions
and 1 timelike dimension,
since we've yet to conceive several dimensions of time and
rarely consider a field of one element.
Some Special Cases :
The simplest unitary group is the "unit circle" or circle group
(denoted T) which is isomorphic to
U(1), SO(2) and
/ .
SZ(n,C) is the cyclic group of order n
(it does "look" cyclic).
The Möbius Group
is isomorphic to PGL(2,C) and/or PSL(2,C).
(2016-05-21) Projective Group
Linear group modulo the scalar group or any group modulo its center.
Traditionally,
the projective group
is the quotient of the general linear group
(i.e., the group of all square matrices
of a given dimension over a given field)
modulo the scalar group (i.e., the diagonal matrices).
The term is also used as a qualifier to denote the quotients
nodulo the scalar group of some subgroups of the general linear group.
By extension, the qualifier projective can even be
used to denote the quotient of any group modulo its own
center.
(See modular group.)
The qualifier "projective" is inherited from the name given to the rules of
geometrical
perspective, first devised by Renaissance artists.
In their drawings, they mapped every point (P) of three-dimensional
Euclidean space to the unique point (M) of a planar canvas intersecting
the straight line (OP) drawn from that point to the eye of the
observer (O). In such a mapping,
a horizontal plane is mapped onto a half-plane of the canvas
which ends on a straight line representing the
horizon
(supposedly "at infinity").
In one of the greatest leaps of imagination ever made by the human mind,
the geometers of the nineteen century realized that
this artistic rendering was just a special case of the
above and they would eventually turn
projective geometry
into a very fruitful independent field of study.
Once called higher geometry,
that became a revered part or higher learning before
being all but forgotten...
(2006-04-12) The Möbius Group
(homographic transformations)
The automorphisms of the Riemann Sphere
(the projective line).
An homographic transformation f
(also called a Möbius transformation or a
fractional linear transformation) sends a
complex number z to:
f (z) =
a z + b
c z + d
It's a [bijective]
transformation of the projective line
(the complex plane plus a single "infinity"
point ¥ beyond its horizon, so to speak).
The image of ¥ is a/c
(or ¥ if c = 0 ).
The image of -d/c
(or ¥ if c = 0 )
is ¥.
The Stereographic Projection
Projective Line
Riemann Sphere
È {¥}
(a,b,c) Î
3 |
a 2 +
b 2 +
c 2 = 1
¥
(0,0,1)
z =
a + i b
1 - c
(a,b,c)
c ¹ 1
z = u + iv
(
2 u
,
2 v
,
| z | 2- 1
)
| z | 2 + 1
| z | 2 + 1
| z | 2 + 1
Automorphic functions
(originally dubbed "Fuchsian functions" by Poincaré,
around 1884) are meromorphic functions
(i.e., ratios of two holomorphic functions;
analytic functions of a complex variable) which
are invariant under a countable infinity of
Möbius transformations).
(2016-05-22) The Modular Group
G The common name of the
projective special linear group
PSL(2,).
The locution gamma group is best reserved for something else.
The modular group consists of all 2 by 2
square matrices
with integer elements
(in )
and unitdeterminant
(that's what special means)
when considered modulo the center
{I,-I} (that's what projective means).
That last specification merely states that a matrix and its opposite are
equivalent representations of the same element of the modular group.
The modular groupG
has the following presentation:
The modular group was first studied in detail, for its own sake,
by Richard Dedekind and
Felix Klein as part of the
Erlangen program (1872).
The closely related elliptic functions (introduced by
Lagrange in 1785)
had already been studied quite extensively by
Abel
(1827-1828) and
Jacobi
(1829)
who shared the grand prix of the French
Academy of Sciences for that work, in 1830
(after Abel's death).
An interesting source of examples in the modular group is provided by
the successive convergents obtained by
truncating the continued fraction expansion of a number, because the
following relation is naturally satisfied:
(2017-07-29) Group Structure of an Elliptic Curve Group operator defined on a cubic planar curve without singular points.
In the Euclidean plane, a cubic curve without singular points is called an
elliptic curve. That same term is also commonly used to denote the cartesian
equation of such a curve or the wonderful group structure its points can be endowed with,
as described below. Elliptic curves can be considered over various fields
(complex numbers, rationals, p-adic numbers, finite fields).
Elliptic curves over finite fields allow
elliptic-curve cryptography (ECC) which was invented in 1985 and has been widely used since 2004.
Mordell's Theorem (1922) :
In 1901, Poincaré had asked
whether the rational points of a curve of genus 1 are finitely generated.
21 years later, Mordell settled that
for elliptic curves:
An elliptic-curve's rational points form a finitely-generated abelian group.
(2017-08-03) Group Law on a Degenerate Cubic Curve
Combining a circle and a straight line so the latter is a subgroup.
In the Euclidean plane,
let's apply the geometric definition of sums on an elliptic curve
to the degenerate cubic consisting of a circle of unit diameter and a straight line at a distance d from its center.
When at least one point is on the circle, the geometric construction of the
sum of two points presents no difficulty. On the other hand,
if both of the points A and B are on the line, their sum C = A+B
is not immediately clear.
To construct it, we may consider any auxiliary point V on the circle and use the following
identity, involving three sums of the previous kind:
A + B = ( (A+V) + B) - V
For convenience, we choose V on the axis of symmetry of the figure,
so that V = -V, in which case we have a symmetrical defining relation:
A + B = (A+V) + (B+V)
If A' is the mirror-image of A+V (with respect to the horizontal axis of symmetry)
then the law introduced in the non-degenerate case says that
A' is at the intersection of the circle and the AV line. Likewise,
the image B' of B+V is the intersection of BV with the circle.
A+B is on the mirror-image of the line joining A+V and B+V,
which is the line A'B'. So, A+B is at the intersection of A'B'
with our basic vertical line, as shown in the figure at left.
w =
u + v
1 + k uv
If we're concerned with number theory, we choose any rational value for k.
Otherwise, we remark that the above equation encodes a group structure on the real line
in one of three different ways, modulo some rescaling:
k = 0. Ordinary addition.
k = -1. Addition of trigonometric tangents.
k = 1 / c2. Addition of hyperbolic tangents (relativistic rapidities).
Moreover, the limiting case when k tends to infinity can be construed
as ordinary multiplication of the reciprocals of nonzero numbers.
Of course, the (nonzero) rational numbers are not finitely generated under this law,
because there are infinitely many prime numbers.
More generally, we may consider any continuous monotonous function f
from negative infinity to positive infinity and define an abelian group law over the real numbers by:
x o y =
f ( f -1 (x) + f -1 (y) )
Our previous discussion is a special case of that if we choose f
to be either the trigonometric tangent or the hyperbolic tangent.
The former for a line which doesn't intersect the basic circle, the latter for a line which does.
(2017-07-30) Amenable Groups (French: groupes moyennables)
Introduced by
Von Neumann to discuss the Banach-Tarski paradox.
An amenable group is a locally compact
topological group whose elements leave invariant some kind of averaging on bounded functions.
The English word was coined in 1949 by
Mahlon M. Day as a pun ("a-mean-able")
to translate the German term originally used by Von Neumann in 1929 (messbar = measurable).
The French use either the English term or the (better) word moyennable.
(2017-07-28) Richard J. Thompson's Groups (1965)
F is the smallest of the three nested groups F, T and V.
The three Thompson groups F, T and V are also called
vagabond groups, chameleon groups or just
chameleons (the latter term was coined by
Matt Brin
in 1994).
They have unusual properties which have made them counterexamples to several conjectures in group theory.
(2020-01-27) Abelian sandpiles and sandpile groups
Let M be a sandpile for which there's a sandpile Z such that M+Z = M.
Then, Z is a zero (i.e. it's a neutral element for addition) over
the set of all sandpiles of the form X+M, since:
(2006-03-01) The Lorentz Group O(3,1)
has 4 connected components.
Each is isomorphic to the
Restricted Lorentz Group SO+(3,1).
h =
é ê ê ë
-1 0 0 0
0 1 0 0
0 0 1 0
0 0 0 1
ù ú ú û
The Lorentz GroupO(3,1)
is isomorphic to SL(2,C) and
consists of all 4 by 4 real matrices A such that
A* h A = 1,
where h is the metric matrix for three dimensions
of space and one dimension of time.
SO+(3,1) is the (6-dimensional)
Restricted Lorentz Group
consisting of the elements of the Lorentz GroupO(3,1)
which preserve the direction of time and the orientation of space
(boosts and 3D rotations).
In the above, T and P denote a reversal of time and an inversion
of space (the latter could be either a mirror symmetry about a plane
or a symmetry about a point).
The symbol SO(3,1) would denote
the "Special Lorentz Group", the subgroup of the matrices
of O(3,1) with determinant one
(which is a disconnected "half" of O(3,1),
not a connected "quarter" of it).
Poincaré Group :
The Poincaré GroupISO+(3,1) is the
10-dimensional inhomogeneous group of noninverting isometries
for 3 dimensions of space and one dimension of time.
It consists of transformations mapping x to
Lx+a , where
L belongs to the above
Restricted Lorentz GroupSO+(3,1)
and a is some
4-vector.
In spite of their respective successes, General Relativity and the
Standard Model are known to be imperfect theories,
incompatible with each other.
The ultimate laws of physics (if they exist) could only
incorporate those two as approximations applicable to specific experimental
domains (like Newtonian mechanics approximates
Special Relativity for low speeds).
Nobody knows (yet) exactly what symmetries the
ultimate laws of nature should have,
but we may ponder the groups of local symmetries underlying modern mathematical
theories of the 4 known physical interactions:
Maxwell's unification
of electricity and magnetism into
electromagnetism has been ultimately
construed as the discovery
that electrodynamics is invariant under local phase transformations,
with the simple structure of U(1).
The classical quantity associated with that symmetry (by
Noether's theorem ) is simply electric charge.
Quantum electrodynamics (QED) describes electromagnetism as a quantum field.
It became the basic paradigm for all subsequent quantum theories of fundamental physical interactions.
QED describes how photons "mediate" the force
between electrons (or any other charged particles).
The electroweak theory is a satisfying unification
of electromagnetism and weak interactions under the symmetries of the
direct productSU(2)´U(1).
It was devised in 1967 by Steven Weinberg (1933-)
and Abdus Salam (1926-1996) building on earlier work
of Sheldon Glashow (1932-). The three men shared the
1979 Nobel prize
for this. The group SU(2) is isomorphic to 3-dimensional rotations.
The broken electroweak symmetry translates
into 4 vector bosons:
g (the photon)
Z0, W+ and W-.
Broken:
In mathematical physics, a symmetry is said to be
broken when symmetrical equations have an asymmetrical solution.
The theory of strong interactions is known as
quantum chromodynamics (QCD).
It's based on an unbrokenSU(3)
local symmetry, dubbed
color symmetry because of a superficial similarity with
the rules of color vision (whereby
3 primary colors may combine to create colorlessness).
QCD describes how gluons mediate the strong force between
quarks (or anything else with color charge,
including gluons themselves). There are 8 different
types of gluons, corresponding to the 8 dimensions of SU(3).
In this context, SU(3) is often denoted
SUc(3). "C" stands for color.
As described by Albert Einstein's General Theory of Relativity,
gravity's local symmetry is that of the Poincaré
group, which preserves spacetime intervals, as well as the direction of time
and the orientation of space.
The Poincaré group is 10-dimensional.
However, a gauge field
(the graviton) is associated only with the 4 dimensions of spacetime
translations. Suspiciously, no such particle or field is associated with
the 6 dimensions corresponding to Lorentz symmetries
(3 dimensions for spatial rotations
and 3 dimensions for Lorentz boosts).
The so-called Standard Model of particle physicists
describes both strong and electroweak
interactions in a theoretical framework whose symmetries are those of the group
SU(2)´U(1)´SUc(3), which has 12 dimensions.
The model depends on several parameters, adjusted to
fit experimental data but otherwise unexplained.
Different local symmetries would impose different restrictions,
for better or for worse.
One classical group
possessing more dimensions of symmetry (24) than the
Standard Model is SU(5).
The correct local symmetry of
"strong-electroweak" interactions would still not
determine the masses of the vector bosons involved
(particles of spin 1) unless
more is known about the way such a symmetry is broken.
A key aspect of particle physics which is based on a broken symmetry is the
classification of elementary particles into three generations of
flavors.
A mind-boggling supersymmetry across different spins
(SUSY)
seems required of any quantum theory designed
to include gravity in a fully unified quantum theory "of everything":
Supergravity, Superstrings, etc.
In 2010,
Sir Michael Atiyah (1929-2019)
remarked that the known physical symmetries occur naturally in the Tits-Freudenthal
magic square pertaining to associative
hypercomplex numbers obtained through the Cayley-Dickson
construct. He speculated that the introduction of the
fourth (nonassociative) hypercomplex division algebra (the octonions) is somehow
related to gravity. He admonished younger investigators to consider
this possibility, which would give a beautiful role to
allexceptional Lie groups.
(2019-03-12) Renormalization Group and Cosmic Galois Group
QFT's renormalization group is a subgroup
of the cosmic Galois group.
The name cosmic Galois group was coined by
Pierre Cartier around 1998,
as he shared the optimism of Fields medalist
Maxim Kontsevich (1964-)
in the following words:
La parenté de plus en plus manifeste entre le groupe de Grothendieck-Teichmüller (GT) d'une part,
et le groupe de renormalisation de la Théorie Quantique des Champs n'est sans doute que
la première manifestation d'un groupe de symétrie des constantes fondamentales de la physique,
une espèce de groupe de Galois cosmique !
The subject caught the attention of several other people connected with the
IHES.
One comprehensive introduction appears in the work (2004) of
Alain Connes (1947-) and
Matilde Marcolli
(1969-).