Você está na página 1de 45

Group

About the Group theory

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Thu, 26 May 2011 01:36:45 UTC
Contents
Articles
Group (mathematics) 1
Orthogonal group 21
Lie group 31

References
Article Sources and Contributors 41
Image Sources, Licenses and Contributors 42

Article Licenses
License 43
Group (mathematics) 1

Group (mathematics)
In mathematics, a group is an algebraic structure consisting of a set
together with an operation that combines any two of its elements to
form a third element. To qualify as a group, the set and the operation
must satisfy a few conditions called group axioms, namely closure,
associativity, identity and invertibility. Many familiar mathematical
structures such as number systems obey these axioms: for example, the
integers endowed with the addition operation form a group. However,
the abstract formalization of the group axioms, detached as it is from
the concrete nature of any particular group and its operation, allows
entities with highly diverse mathematical origins in abstract algebra
and beyond to be handled in a flexible way, while retaining their
essential structural aspects. The ubiquity of groups in numerous areas
within and outside mathematics makes them a central organizing The possible manipulations of this Rubik's Cube
form a group.
principle of contemporary mathematics.[1] [2]

Groups share a fundamental kinship with the notion of symmetry. A symmetry group encodes symmetry features of
a geometrical object: it consists of the set of transformations that leave the object unchanged, and the operation of
combining two such transformations by performing one after the other. Such symmetry groups, particularly the
continuous Lie groups, play an important role in many academic disciplines. Matrix groups, for example, can be
used to understand fundamental physical laws underlying special relativity and symmetry phenomena in molecular
chemistry.

The concept of a group arose from the study of polynomial equations, starting with Évariste Galois in the 1830s.
After contributions from other fields such as number theory and geometry, the group notion was generalized and
firmly established around 1870. Modern group theory—a very active mathematical discipline—studies groups in
their own right.a[›] To explore groups, mathematicians have devised various notions to break groups into smaller,
better-understandable pieces, such as subgroups, quotient groups and simple groups. In addition to their abstract
properties, group theorists also study the different ways in which a group can be expressed concretely (its group
representations), both from a theoretical and a computational point of view. A particularly rich theory has been
developed for finite groups, which culminated with the monumental classification of finite simple groups completed
in 1983. Since the mid-1980s, geometric group theory, which studies finitely generated groups as geometric objects,
has become a particularly active area in group theory.

Definition and illustration

First example: the integers


One of the most familiar groups is the set of integers Z which consists of the numbers
..., −4, −3, −2, −1, 0, 1, 2, 3, 4, ...[3]
The following properties of integer addition serve as a model for the abstract group axioms given in the definition
below.
1. For any two integers a and b, the sum a + b is also an integer. In other words, the process of adding integers two
at a time always yields an integer, not some other type of number such as a fraction. This property is known as
closure under addition.
2. For all integers a, b and c, (a + b) + c = a + (b + c). Expressed in words, adding a to b first, and then adding the
result to c gives the same final result as adding a to the sum of b and c, a property known as associativity.
Group (mathematics) 2

3. If a is any integer, then 0 + a = a + 0 = a. Zero is called the identity element of addition because adding it to any
integer returns the same integer.
4. For every integer a, there is an integer b such that a + b = b + a = 0. The integer b is called the inverse element of
the integer a and is denoted −a.
The integers, together with the operation +, form a mathematical object belonging to a broad class sharing similar
structural aspects. To appropriately understand these structures as a collective, the following abstract definition is
developed.

Definition
A group is a set, G, together with an operation • (called the group law of G) that combines any two elements a and b
to form another element, denoted a • b or ab. To qualify as a group, the set and operation, (G, •), must satisfy four
requirements known as the group axioms:[4]
Closure
For all a, b in G, the result of the operation, a • b, is also in G.b[›]
Associativity
For all a, b and c in G, (a • b) • c = a • (b • c).
Identity element
There exists an element e in G, such that for every element a in G, the equation e • a = a • e = a holds. The
identity element of a group G is often written as 1 or 1G,[5] a notation inherited from the multiplicative
identity.
Inverse element
For each a in G, there exists an element b in G such that a • b = b • a = 1G.
The order in which the group operation is carried out can be significant. In other words, the result of combining
element a with element b need not yield the same result as combining element b with element a; the equation
a•b=b•a
may not always be true. This equation does always hold in the group of integers under addition, because a + b = b +
a for any two integers (commutativity of addition). However, it does not always hold in the symmetry group below.
Groups for which the equation a • b = b • a always holds are called abelian (in honor of Niels Abel). Thus, the
integer addition group is abelian, but the symmetry group in the following section is not.
The set G is called the underlying set of the group (G, •). Often the group's underlying set G is used as a short name
for the group (G, •). Along the same lines, sometimes a shorthand expression such as "a subset of the group G" is
used when what is actually meant is "a subset of the underlying set G of the group (G, •)." Usually, it is clear from
the context whether a symbol like G refers to a group or to an underlying set.

Second example: a symmetry group


The symmetry operations (i.e., rotations and reflections) of a square form the elements of a group called the dihedral
group of degree 4, denoted D4, whose group operation is the composition of symmetry operations.[6] The following
symmetries occur:
Group (mathematics) 3

id (keeping it as is) r1 (rotation by 90° right) r2 (rotation by 180° right) r3 (rotation by 270° right)

fv (vertical flip) fh (horizontal flip) fd (diagonal flip) fc (counter-diagonal flip)


The elements of the symmetry group of the square (D4). The vertices are colored and numbered only to visualize the operations.

• the identity operation leaving everything unchanged, denoted id;


• rotations of the square by 90° right, 180° right, and 270° right, denoted by r1, r2 and r3, respectively;
• reflections about the vertical and horizontal middle line (fh and fv), or through the two diagonals (fd and fc).
The defining operation of this group is function composition: The eight symmetries are functions from the square to
the square, and two symmetries are combined by composing them as functions, that is, applying the first one to the
square, and the second one to the result of the first application. The result of performing first a and then b is written
symbolically from right to left as
b • a ("apply the symmetry b after performing the symmetry a"). The right-to-left notation is the same notation
that is used for composition of functions.
The group table on the right lists the results of all such compositions possible. For example, rotating by 270° right
(r3) and then flipping horizontally (fh) is the same as performing a reflection along the diagonal (fd). Using the above
symbols, highlighted in blue in the group table:
fh • r3 = fd.

Group table of D4
• id r1 r2 r3 fv fh fd fc

id id r1 r2 r3 fv fh fd fc

r1 r1 r2 r3 id fc fd fv fh

r2 r2 r3 id r1 fh fv fc fd

r3 r3 id r1 r2 fd fc fh fv

fv fv fd fh fc id r2 r1 r3

fh fh fc fv fd r2 id r3 r1

fd fd fh fc fv r3 r1 id r2

fc fc fv fd fh r1 r3 r2 id

The elements id, r1, r2, and r3 form a subgroup, highlighted in red (upper left region). A left and right coset of this subgroup is highlighted in green
(in the last row) and yellow (last column), respectively.

Given this set of symmetries and the described operation, the group axioms can be understood as follows:
1. The closure axiom demands that the composition b • a of any two symmetries a and b is also a symmetry.
Another example for the group operation is
r3 • fh = fc,
Group (mathematics) 4

i.e. rotating 270° right after flipping horizontally equals flipping along the counter-diagonal (fc). Indeed every
other combination of two symmetries still gives a symmetry, as can be checked using the group table.
2. The associativity constraint deals with composing more than two symmetries: Starting with three elements a, b
and c of D4, there are two possible ways of using these three symmetries in this order to determine a symmetry of
the square. One of these ways is to first compose a and b into a single symmetry, then to compose that symmetry
with c. The other way is to first compose b and c, then to compose the resulting symmetry with a. The
associativity condition
(a • b) • c = a • (b • c)
means that these two ways are the same, i.e., a product of many group elements can be simplified in any order.
For example, (fd • fv) • r2 = fd • (fv • r2) can be checked using the group table at the right

(fd • fv) • r2 = r3 • r2 = r1, which equals

fd • (fv • r2) = fd • fh = r1.

While associativity is true for the symmetries of the square and addition of numbers, it is not true for all
operations. For instance, subtraction of numbers is not associative: (7 − 3) − 2 = 2 is not the same as 7 − (3 − 2) =
6.
3. The identity element is the symmetry id leaving everything unchanged: for any symmetry a, performing id after a
(or a after id) equals a, in symbolic form,
id • a = a,
a • id = a.
4. An inverse element undoes the transformation of some other element. Every symmetry can be undone: each of
transformations—identity id, the flips fh, fv, fd, fc and the 180° rotation r2—is its own inverse, because performing
each one twice brings the square back to its original orientation. The rotations r3 and r1 are each other's inverse,
because rotating 90° and then rotation 270° (or vice versa) yields a rotation over 360° which leaves the square
unchanged. In symbols,
fh • fh = id,
r3 • r1 = r1 • r3 = id.
In contrast to the group of integers above, where the order of the operation is irrelevant, it does matter in D4: fh • r1 =
fc but r1 • fh = fd. In other words, D4 is not abelian, which makes the group structure more difficult than the integers
introduced first.

History
The modern concept of an abstract group developed out of several fields of mathematics.[7] [8] [9] The original
motivation for group theory was the quest for solutions of polynomial equations of degree higher than 4. The
19th-century French mathematician Évariste Galois, extending prior work of Paolo Ruffini and Joseph-Louis
Lagrange, gave a criterion for the solvability of a particular polynomial equation in terms of the symmetry group of
its roots (solutions). The elements of such a Galois group correspond to certain permutations of the roots. At first,
Galois' ideas were rejected by his contemporaries, and published only posthumously.[10] [11] More general
permutation groups were investigated in particular by Augustin Louis Cauchy. Arthur Cayley's On the theory of
groups, as depending on the symbolic equation θn = 1 (1854) gives the first abstract definition of a finite group.[12]
Geometry was a second field in which groups were used systematically, especially symmetry groups as part of Felix
Klein's 1872 Erlangen program.[13] After novel geometries such as hyperbolic and projective geometry had emerged,
Klein used group theory to organize them in a more coherent way. Further advancing these ideas, Sophus Lie
founded the study of Lie groups in 1884.[14]
Group (mathematics) 5

The third field contributing to group theory was number theory. Certain abelian group structures had been used
implicitly in Carl Friedrich Gauss' number-theoretical work Disquisitiones Arithmeticae (1798), and more explicitly
by Leopold Kronecker.[15] In 1847, Ernst Kummer led early attempts to prove Fermat's Last Theorem to a climax by
developing groups describing factorization into prime numbers.[16]
The convergence of these various sources into a uniform theory of groups started with Camille Jordan's Traité des
substitutions et des équations algébriques (1870).[17] Walther von Dyck (1882) gave the first statement of the
modern definition of an abstract group.[18] As of the 20th century, groups gained wide recognition by the pioneering
work of Ferdinand Georg Frobenius and William Burnside, who worked on representation theory of finite groups,
Richard Brauer's modular representation theory and Issai Schur's papers.[19] The theory of Lie groups, and more
generally locally compact groups was pushed by Hermann Weyl, Élie Cartan and many others.[20] Its algebraic
counterpart, the theory of algebraic groups, was first shaped by Claude Chevalley (from the late 1930s) and later by
pivotal work of Armand Borel and Jacques Tits.[21]
The University of Chicago's 1960–61 Group Theory Year brought together group theorists such as Daniel
Gorenstein, John G. Thompson and Walter Feit, laying the foundation of a collaboration that, with input from
numerous other mathematicians, classified all finite simple groups in 1982. This project exceeded previous
mathematical endeavours by its sheer size, in both length of proof and number of researchers. Research is ongoing to
simplify the proof of this classification.[22] These days, group theory is still a highly active mathematical branch
crucially impacting many other fields.a[›]

Elementary consequences of the group axioms


Basic facts about all groups that can be obtained directly from the group axioms are commonly subsumed under
elementary group theory.[23] For example, repeated applications of the associativity axiom show that the
unambiguity of
a • b • c = (a • b) • c = a • (b • c)
generalizes to more than three factors. Because this implies that parentheses can be inserted anywhere within such a
series of terms, parentheses are usually omitted.[24]
The axioms may be weakened to assert only the existence of a left identity and left inverses. Both can be shown to be
actually two-sided, so the resulting definition is equivalent to the one given above.[25]

Uniqueness of identity element and inverses


Two important consequences of the group axioms are the uniqueness of the identity element and the uniqueness of
inverse elements. There can be only one identity element in a group, and each element in a group has exactly one
inverse element. Thus, it is customary to speak of the identity, and the inverse of an element.[26]
To prove the uniqueness of an inverse element of a, suppose that a has two inverses, denoted l and r, in a group (G,
•). Then
Group (mathematics) 6

l = l • 1G as 1G is the identity element

= l • (a • r) because r is an inverse of a, so 1G = a • r

= (l • a) • r by associativity, which allows to rearrange the parentheses

= 1G • r since l is an inverse of a, i.e. l • a = 1G

= r for 1G is the identity element

The two extremal terms l and r are equal, since they are connected by a chain of equalities. In other words there is
only one inverse element of a. Similarly, to prove that the identity element of a group is unique, assume G is a group
with two identity elements 1G and e. Then 1G = 1G • e = e, hence 1G and e are equal.

Division
In groups, it is possible to perform division: given elements a and b of the group G, there is exactly one solution x in
G to the equation x • a = b.[26] In fact, right multiplication of the equation by a−1 gives the solution x = x • a • a−1 = b
• a−1. Similarly there is exactly one solution y in G to the equation a • y = b, namely y = a−1 • b. In general, x and y
need not agree.
A consequence of this is that multiplying by a group element g is a bijection. Specifically, if g is an element of the
group G, there is a bijection from G to itself called left translation by g sending h ∈ G to g • h. Similarly, right
translation by g is a bijection from G to itself sending h to h • g. If G is abelian, left and right translation by a group
element are the same.

Basic concepts
To understand groups beyond the level of mere symbolic manipulations as above, more structural concepts have to
be employed.c[›] There is a conceptual principle underlying all of the following notions: to take advantage of the
structure offered by groups (which sets, being "structureless", do not have), constructions related to groups have to
be compatible with the group operation. This compatibility manifests itself in the following notions in various ways.
For example, groups can be related to each other via functions called group homomorphisms. By the mentioned
principle, they are required to respect the group structures in a precise sense. The structure of groups can also be
understood by breaking them into pieces called subgroups and quotient groups. The principle of "preserving
structures"—a recurring topic in mathematics throughout—is an instance of working in a category, in this case the
category of groups.[27]

Group homomorphisms
Group homomorphismsg[›] are functions that preserve group structure. A function a: G → H between two groups
(G,•) and (H,*) is a homomorphism if the equation
a(g • k) = a(g) * a(k)
holds for all elements g, k in G. In other words, the result is the same when performing the group operation after or
before applying the map a. This requirement ensures that a(1G) = 1H, and also a(g)−1 = a(g−1) for all g in G. Thus a
group homomorphism respects all the structure of G provided by the group axioms.[28]
Two groups G and H are called isomorphic if there exist group homomorphisms a: G → H and b: H → G, such that
applying the two functions one after another (in each of the two possible orders) equal the identity function of G and
H, respectively. That is, a(b(h)) = h and b(a(g)) = g for any g in G and h in H. From an abstract point of view,
isomorphic groups carry the same information. For example, proving that g • g = 1G for some element g of G is
equivalent to proving that a(g) • a(g) = 1H, because applying a to the first equality yields the second, and applying b
to the second gives back the first.
Group (mathematics) 7

Subgroups
Informally, a subgroup is a group H contained within a bigger one, G.[29] Concretely, the identity element of G is
contained in H, and whenever h1 and h2 are in H, then so are h1 • h2 and h1−1, so the elements of H, equipped with
the group operation on G restricted to H, indeed form a group.
In the example above, the identity and the rotations constitute a subgroup R = {id, r1, r2, r3}, highlighted in red in the
group table above: any two rotations composed are still a rotation, and a rotation can be undone by (i.e. is inverse to)
the complementary rotations 270° for 90°, 180° for 180°, and 90° for 270° (note that rotation in the opposite
direction is not defined). The subgroup test is a necessary and sufficient condition for a subset H of a group G to be a
subgroup: it is sufficient to check that g−1h ∈ H for all elements g, h ∈ H. Knowing the subgroups is important in
understanding the group as a whole.d[›]
Given any subset S of a group G, the subgroup generated by S consists of products of elements of S and their
inverses. It is the smallest subgroup of G containing S.[30] In the introductory example above, the subgroup generated
by r2 and fv consists of these two elements, the identity element id and fh = fv • r2. Again, this is a subgroup, because
combining any two of these four elements or their inverses (which are, in this particular case, these same elements)
yields an element of this subgroup.

Cosets
In many situations it is desirable to consider two group elements the same if they differ by an element of a given
subgroup. For example, in D4 above, once a flip is performed, the square never gets back to the r2 configuration by
just applying the rotation operations (and no further flips), i.e. the rotation operations are irrelevant to the question
whether a flip has been performed. Cosets are used to formalize this insight: a subgroup H defines left and right
cosets, which can be thought of as translations of H by arbitrary group elements g. In symbolic terms, the left and
right coset of H containing g are
gH = {g • h, h ∈ H} and Hg = {h • g, h ∈ H}, respectively.[31]
The cosets of any subgroup H form a partition of G; that is, the union of all left cosets is equal to G and two left
cosets are either equal or have an empty intersection.[32] The first case g1H = g2H happens precisely when g1−1 • g2
∈ H, i.e. if the two elements differ by an element of H. Similar considerations apply to the right cosets of H. The left
and right cosets of H may or may not be equal. If they are, i.e. for all g in G, gH = Hg, then H is said to be a normal
subgroup. One may then simply refer to N as the set of cosets.
In D4, the introductory symmetry group, the left cosets gR of the subgroup R consisting of the rotations are either
equal to R, if g is an element of R itself, or otherwise equal to U = fcR = {fc, fv, fd, fh} (highlighted in green). The
subgroup R is also normal, because fcR = U = Rfc and similarly for any element other than fc.

Quotient groups
In addition to disregarding the internal structure of a subgroup by considering its cosets, it is desirable to endow this
coarser entity with a group law called quotient group or factor group. For this to be possible, the subgroup has to be
normal. Given any normal subgroup N, the quotient group is defined by
G / N = {gN, g ∈ G}, "G modulo N".[33]
This set inherits a group operation (sometimes called coset multiplication, or coset addition) from the original group
G: (gN) • (hN) = (gh)N for all g and h in G. This definition is motivated by the idea (itself an instance of general
structural considerations outlined above) that the map G → G / N that associates to any element g its coset gN be a
group homomorphism, or by general abstract considerations called universal properties. The coset eN = N serves as
the identity in this group, and the inverse of gN in the quotient group is (gN)−1 = (g−1)N.e[›]
Group (mathematics) 8

• R U

R R U

U U R

Group table of the quotient group D4 / R.

The elements of the quotient group D4 / R are R itself, which represents the identity, and U = fvR. The group
operation on the quotient is shown at the right. For example, U • U = fvR • fvR = (fv • fv)R = R. Both the subgroup R
= {id, r1, r2, r3}, as well as the corresponding quotient are abelian, whereas D4 is not abelian. Building bigger groups
by smaller ones, such as D4 from its subgroup R and the quotient D4 / R is abstracted by a notion called semidirect
product.
Quotient and subgroups together form a way of describing every group by its presentation: any group is the quotient
of the free group over the generators of the group, quotiented by the subgroup of relations. The dihedral group D4,
for example, can be generated by two elements r and f (for example, r = r1, the right rotation and f = fv the vertical
(or any other) flip), which means that every symmetry of the square is a finite composition of these two symmetries
or their inverses. Together with the relations
r 4 = f 2 = (r • f)2 = 1,[34]
the group is completely described. A presentation of a group can also be used to construct the Cayley graph, a device
used to graphically capture discrete groups.
Sub- and quotient groups are related in the following way: a subset H of G can be seen as an injective map H → G,
i.e. any element of the target has at most one element that maps to it. The counterpart to injective maps are surjective
maps (every element of the target is mapped onto), such as the canonical map G → G / N.y[›] Interpreting subgroup
and quotients in light of these homomorphisms emphasizes the structural concept inherent to these definitions
alluded to in the introduction. In general, homomorphisms are neither injective nor surjective. Kernel and image of
group homomorphisms and the first isomorphism theorem address this phenomenon.

Examples and applications

A periodic wallpaper pattern gives rise to a wallpaper group.

The fundamental group of a plane minus a point (bold) consists of loops around the missing point. This group is isomorphic to the integers.
Group (mathematics) 9

Examples and applications of groups abound. A starting point is the group Z of integers with addition as group
operation, introduced above. If instead of addition multiplication is considered, one obtains multiplicative groups.
These groups are predecessors of important constructions in abstract algebra.
Groups are also applied in many other mathematical areas. Mathematical objects are often examined by associating
groups to them and studying the properties of the corresponding groups. For example, Henri Poincaré founded what
is now called algebraic topology by introducing the fundamental group.[35] By means of this connection, topological
properties such as proximity and continuity translate into properties of groups.i[›] For example, elements of the
fundamental group are represented by loops. The second image at the right shows some loops in a plane minus a
point. The blue loop is considered null-homotopic (and thus irrelevant), because it can be continuously shrunk to a
point. The presence of the hole prevents the orange loop from being shrunk to a point. The fundamental group of the
plane with a point deleted turns out to be infinite cyclic, generated by the orange loop (or any other loop winding
once around the hole). This way, the fundamental group detects the hole.
In more recent applications, the influence has also been reversed to motivate geometric constructions by a
group-theoretical background.j[›] In a similar vein, geometric group theory employs geometric concepts, for example
in the study of hyperbolic groups.[36] Further branches crucially applying groups include algebraic geometry and
number theory.[37]
In addition to the above theoretical applications, many practical applications of groups exist. Cryptography relies on
the combination of the abstract group theory approach together with algorithmical knowledge obtained in
computational group theory, in particular when implemented for finite groups.[38] Applications of group theory are
not restricted to mathematics; sciences such as physics, chemistry and computer science benefit from the concept.

Numbers
Many number systems, such as the integers and the rationals enjoy a naturally given group structure. In some cases,
such as with the rationals, both addition and multiplication operations give rise to group structures. Such number
systems are predecessors to more general algebraic structures known as rings and fields. Further abstract algebraic
concepts such as modules, vector spaces and algebras also form groups.

Integers
The group of integers Z under addition, denoted (Z, +), has been described above. The integers, with the operation of
multiplication instead of addition, (Z, ·) do not form a group. The closure, associativity and identity axioms are
satisfied, but inverses do not exist: for example, a = 2 is an integer, but the only solution to the equation a · b = 1 in
this case is b = 1/2, which is a rational number, but not an integer. Hence not every element of Z has a
(multiplicative) inverse.k[›]

Rationals
The desire for the existence of multiplicative inverses suggests considering fractions

Fractions of integers (with b nonzero) are known as rational numbers.l[›] The set of all such fractions is commonly
denoted Q. There is still a minor obstacle for (Q, ·), the rationals with multiplication, being a group: because the
rational number 0 does not have a multiplicative inverse (i.e., there is no x such that x · 0 = 1), (Q, ·) is still not a
group.
However, the set of all nonzero rational numbers Q \ {0} = {q ∈ Q, q ≠ 0} does form an abelian group under
multiplication, denoted (Q \ {0}, ·).m[›] Associativity and identity element axioms follow from the properties of
integers. The closure requirement still holds true after removing zero, because the product of two nonzero rationals is
never zero. Finally, the inverse of a/b is b/a, therefore the axiom of the inverse element is satisfied.
Group (mathematics) 10

The rational numbers (including 0) also form a group under addition. Intertwining addition and multiplication
operations yields more complicated structures called rings and—if division is possible, such as in Q—fields, which
occupy a central position in abstract algebra. Group theoretic arguments therefore underlie parts of the theory of
those entities.n[›]

Nonzero integers modulo a prime


For any prime number p, modular arithmetic furnishes the multiplicative group of integers modulo p.[39] Its elements
are integers not divisible by p, considered modulo p, i.e. two numbers are considered equivalent if their difference is
divisible by p. For example, if p = 5, there are exactly four group elements 1, 2, 3, 4: multiples of 5 are excluded and
6 and −4 are both equivalent to 1 etc. The group operation is given by multiplication. Therefore, 4 · 4 = 1, because
the usual product 16 is equivalent to 1, for 5 divides 16 − 1 = 15, denoted
16 ≡ 1 (mod 5).
The primality of p ensures that the product of two integers neither of which is divisible by p is not divisible by p
either, hence the indicated set of classes is closed under multiplication.o[›] The identity element is 1, as usual for a
multiplicative group, and the associativity follows from the corresponding property of integers. Finally, the inverse
element axiom requires that given an integer a not divisible by p, there exists an integer b such that
a · b ≡ 1 (mod p), i.e. p divides the difference a · b − 1.
The inverse b can be found by using Bézout's identity and the fact that the greatest common divisor gcd(a, p) equals
1.[40] In the case p = 5 above, the inverse of 4 is 4, and the inverse of 3 is 2, as 3 · 2 = 6 ≡ 1 (mod 5). Hence all group
axioms are fulfilled. Actually, this example is similar to (Q\{0}, ·) above, because it turns out to be the multiplicative
group of nonzero elements in the finite field Fp, denoted Fp×.[41] These groups are crucial to public-key
cryptography.p[›]

Cyclic groups
A cyclic group is a group all of whose elements are powers (when the group
operation is written additively, the term 'multiple' can be used) of a particular
element a.[42] In multiplicative notation, the elements of the group are:
..., a−3, a−2, a−1, a0 = e, a, a2, a3, ...,
where a2 means a • a, and a−3 stands for a−1 • a−1 • a−1=(a • a • a)−1 etc.h[›] Such
an element a is called a generator or a primitive element of the group.
A typical example for this class of groups is the group of n-th complex roots of
unity, given by complex numbers z satisfying zn = 1 (and whose operation is
The 6th complex roots of unity form
multiplication).[43] Any cyclic group with n elements is isomorphic to this group.
a cyclic group. z is a primitive
Using some field theory, the group Fp× can be shown to be cyclic: for example, if element, but z2 is not, because the
p = 5, 3 is a generator since 31 = 3, 32 = 9 ≡ 4, 33 ≡ 2, and 34 ≡ 1. odd powers of z are not a power of
z2.
Some cyclic groups have an infinite number of elements. In these groups, for
every non-zero element a, all the powers of a are distinct; despite the name
"cyclic group", the powers of the elements do not cycle. An infinite cyclic group is isomorphic to (Z, +), the group of
integers under addition introduced above.[44] As these two prototypes are both abelian, so is any cyclic group.
The study of abelian groups is quite mature, including the fundamental theorem of finitely generated abelian groups;
and reflecting this state of affairs, many group-related notions, such as center and commutator, describe the extent to
which a given group is not abelian.[45]
Group (mathematics) 11

Symmetry groups
Symmetry groups are groups consisting of symmetries of given mathematical objects—be they of geometric nature,
such as the introductory symmetry group of the square, or of algebraic nature, such as polynomial equations and
their solutions.[46] Conceptually, group theory can be thought of as the study of symmetry.t[›] Symmetries in
mathematics greatly simplify the study of geometrical or analytical objects. A group is said to act on another
mathematical object X if every group element performs some operation on X compatibly to the group law. In the
rightmost example below, an element of order 7 of the (2,3,7) triangle group acts on the tiling by permuting the
highlighted warped triangles (and the other ones, too). By a group action, the group pattern is connected to the
structure of the object being acted on.
In chemical fields, such as crystallography, space groups and point
groups describe molecular symmetries and crystal symmetries.
These symmetries underlie the chemical and physical behavior of
these systems, and group theory enables simplification of quantum
mechanical analysis of these properties.[47] For example, group
theory is used to show that optical transitions between certain
quantum levels cannot occur simply because of the symmetry of
the states involved.

Not only are groups useful to assess the implications of


symmetries in molecules, but surprisingly they also predict that
molecules sometimes can change symmetry. The Jahn-Teller
effect is a distortion of a molecule of high symmetry when it
Rotations and flips form the symmetry group of a great
adopts a particular ground state of lower symmetry from a set of
icosahedron.
possible ground states that are related to each other by the
symmetry operations of the molecule.[48] [49]

Likewise, group theory helps predict the changes in physical properties that occur when a material undergoes a phase
transition, for example, from a cubic to a tetrahedral crystalline form. An example is ferroelectric materials, where
the change from a paraelectric to a ferroelectric state occurs at the Curie temperature and is related to a change from
the high-symmetry paraelectric state to the lower symmetry ferroelectic state, accompanied by a so-called soft
phonon mode, a vibrational lattice mode that goes to zero frequency at the transition.[50]
Such spontaneous symmetry breaking has found further application in elementary particle physics, where its
occurrence is related to the appearance of Goldstone bosons.

Buckminsterfullerene Ammonia, NH3. Its Cubane C8H8 The (2,3,7) triangle group, a
Hexaaquacopper(II) complex ion,
displays symmetry group is of order 6, features hyperbolic group, acts on
[Cu(OH2)6]2+. Compared to a perfectly
icosahedral symmetry. generated by a 120° rotation octahedral this tiling of the hyperbolic
symmetrical shape, the molecule is
and a reflection. symmetry. plane.
vertically dilated by about 22%
(Jahn-Teller effect).

Finite symmetry groups such as the Mathieu groups are used in coding theory, which is in turn applied in error
correction of transmitted data, and in CD players.[51] Another application is differential Galois theory, which
Group (mathematics) 12

characterizes functions having antiderivatives of a prescribed form, giving group-theoretic criteria for when solutions
of certain differential equations are well-behaved.u[›] Geometric properties that remain stable under group actions are
investigated in (geometric) invariant theory.[52]

General linear group and representation theory


Matrix groups consist of matrices together with matrix
multiplication. The general linear group GL(n, R)
consists of all invertible n-by-n matrices with real
entries.[53] Its subgroups are referred to as matrix
groups or linear groups. The dihedral group example
mentioned above can be viewed as a (very small)
matrix group. Another important matrix group is the Two vectors (the left illustration) multiplied by matrices (the middle
special orthogonal group SO(n). It describes all and right illustrations). The middle illustration represents a clockwise
rotation by 90°, while the right-most one stretches the x-coordinate
possible rotations in n dimensions. Via Euler angles,
by factor 2.
rotation matrices are used in computer graphics.[54]

Representation theory is both an application of the group concept and important for a deeper understanding of
groups.[55] [56] It studies the group by its group actions on other spaces. A broad class of group representations are
linear representations, i.e. the group is acting on a vector space, such as the three-dimensional Euclidean space R3. A
representation of G on an n-dimensional real vector space is simply a group homomorphism
ρ: G → GL(n, R)
from the group to the general linear group. This way, the group operation, which may be abstractly given, translates
to the multiplication of matrices making it accessible to explicit computations.w[›]
Given a group action, this gives further means to study the object being acted on.x[›] On the other hand, it also yields
information about the group. Group representations are an organizing principle in the theory of finite groups, Lie
groups, algebraic groups and topological groups, especially (locally) compact groups.[55] [57]

Galois groups
Galois groups have been developed to help solve polynomial equations by capturing their symmetry features.[58] [59]
For example, the solutions of the quadratic equation ax2 + bx + c = 0 are given by

Exchanging "+" and "−" in the expression, i.e. permuting the two solutions of the equation can be viewed as a (very
simple) group operation. Similar formulae are known for cubic and quartic equations, but do not exist in general for
degree 5 and higher.[60] Abstract properties of Galois groups associated with polynomials (in particular their
solvability) give a criterion for polynomials that have all their solutions expressible by radicals, i.e. solutions
expressible using solely addition, multiplication, and roots similar to the formula above.[61]
The problem can be dealt with by shifting to field theory and considering the splitting field of a polynomial. Modern
Galois theory generalizes the above type of Galois groups to field extensions and establishes—via the fundamental
theorem of Galois theory—a precise relationship between fields and groups, underlining once again the ubiquity of
groups in mathematics.
Group (mathematics) 13

Finite groups
A group is called finite if it has a finite number of elements. The number of elements is called the order of the group
G.[62] An important class is the symmetric groups SN, the groups of permutations of N letters. For example, the
symmetric group on 3 letters S3 is the group consisting of all possible swaps of the three letters ABC, i.e. contains the
elements ABC, ACB, ..., up to CBA, in total 6 (or 3 factorial) elements. This class is fundamental insofar as any finite
group can be expressed as a subgroup of a symmetric group SN for a suitable integer N (Cayley's theorem). Parallel
to the group of symmetries of the square above, S3 can also be interpreted as the group of symmetries of an
equilateral triangle.
The order of an element a in a group G is the least positive integer n such that a n = e, where a n represents

i.e. application of the operation • to n copies of a. (If • represents multiplication, then an corresponds to the nth power
of a.) In infinite groups, such an n may not exist, in which case the order of a is said to be infinity. The order of an
element equals the order of the cyclic subgroup generated by this element.
More sophisticated counting techniques, for example counting cosets, yield more precise statements about finite
groups: Lagrange's Theorem states that for a finite group G the order of any finite subgroup H divides the order of G.
The Sylow theorems give a partial converse.
The dihedral group (discussed above) is a finite group of order 8. The order of r1 is 4, as is the order of the subgroup
R it generates (see above). The order of the reflection elements fv etc. is 2. Both orders divide 8, as predicted by
Lagrange's Theorem. The groups Fp× above have order p − 1.

Classification of finite simple groups


Mathematicians often strive for a complete classification (or list) of a mathematical notion. In the context of finite
groups, this aim quickly leads to difficult and profound mathematics. According to Lagrange's theorem, finite groups
of order p, a prime number, are necessarily cyclic (abelian) groups Zp. Groups of order p2 can also be shown to be
abelian, a statement which does not generalize to order p3, as the non-abelian group D4 of order 8 = 23 above
shows.[63] Computer algebra systems can be used to list small groups, but there is no classification of all finite
groups.q[›] An intermediate step is the classification of finite simple groups.r[›] A nontrivial group is called simple if
its only normal subgroups are the trivial group and the group itself.s[›] The Jordan–Hölder theorem exhibits finite
simple groups as the building blocks for all finite groups.[64] Listing all finite simple groups was a major
achievement in contemporary group theory. 1998 Fields Medal winner Richard Borcherds succeeded to prove the
monstrous moonshine conjectures, a surprising and deep relation of the largest finite simple sporadic group—the
"monster group"—with certain modular functions, a piece of classical complex analysis, and string theory, a theory
supposed to unify the description of many physical phenomena.[65]
Group (mathematics) 14

Groups with additional structure


Many groups are simultaneously groups and examples of other mathematical structures. In the language of category
theory, they are group objects in a category, meaning that they are objects (that is, examples of another mathematical
structure) which come with transformations (called morphisms) that mimic the group axioms. For example, every
group (as defined above) is also a set, so a group is a group object in the category of sets.

Topological groups
Some topological spaces may be endowed with a group law. In order
for the group law and the topology to interweave well, the group
operations must be continuous functions, that is, g • h, and g−1 must
not vary wildly if g and h vary only little. Such groups are called
topological groups, and they are the group objects in the category of
topological spaces.[66] The most basic examples are the reals R under
addition, (R \ {0}, ·), and similarly with any other topological field
such as the complex numbers or p-adic numbers. All of these groups
are locally compact, so they have Haar measures and can be studied via
harmonic analysis. The former offer an abstract formalism of invariant
integrals. Invariance means, in the case of real numbers for example:

The unit circle in the complex plane under


complex multiplication is a Lie group and,
therefore, a topological group. It is topological
since complex multiplication and division are
continuous. It is a manifold and thus a Lie group,
because every small piece, such as the red arc in
the figure, looks like a part of the real line (shown
at the bottom).

for any constant c. Matrix groups over these fields fall under this regime, as do adele rings and adelic algebraic
groups, which are basic to number theory.[67] Galois groups of infinite field extensions such as the absolute Galois
group can also be equipped with a topology, the so-called Krull topology, which in turn is central to generalize the
above sketched connection of fields and groups to infinite field extensions.[68] An advanced generalization of this
idea, adapted to the needs of algebraic geometry, is the étale fundamental group.[69]

Lie groups
Lie groups (in honor of Sophus Lie) are groups which also have a manifold structure, i.e. they are spaces looking
locally like some Euclidean space of the appropriate dimension.[70] Again, the additional structure, here the manifold
structure, has to be compatible, i.e. the maps corresponding to multiplication and the inverse have to be smooth.
A standard example is the general linear group introduced above: it is an open subset of the space of all n-by-n
matrices, because it is given by the inequality
det (A) ≠ 0,
where A denotes an n-by-n matrix.[71]
Lie groups are of fundamental importance in physics: Noether's theorem links continuous symmetries to conserved
quantities.[72] Rotation, as well as translations in space and time are basic symmetries of the laws of mechanics.
Group (mathematics) 15

They can, for instance, be used to construct simple models—imposing, say, axial symmetry on a situation will
typically lead to significant simplification in the equations one needs to solve to provide a physical description.v[›]
Another example are the Lorentz transformations, which relate measurements of time and velocity of two observers
in motion relative to each other. They can be deduced in a purely group-theoretical way, by expressing the
transformations as a rotational symmetry of Minkowski space. The latter serves—in the absence of significant
gravitation—as a model of space time in special relativity.[73] The full symmetry group of Minkowski space, i.e.
including translations, is known as the Poincaré group. By the above, it plays a pivotal role in special relativity and,
by implication, for quantum field theories.[74] Symmetries that vary with location are central to the modern
description of physical interactions with the help of gauge theory.[75]

Generalizations
Group-like structures

Totality Associativity Identity Inverses

Group Yes Yes Yes Yes

Monoid Yes Yes Yes No

Semigroup Yes Yes No No

Loop Yes No Yes Yes

Quasigroup Yes No No No

Magma Yes No No No

Groupoid No Yes Yes Yes

Category No Yes Yes No

In abstract algebra, more general structures are defined by relaxing some of the axioms defining a group.[27] [76] [77]
For example, if the requirement that every element has an inverse is eliminated, the resulting algebraic structure is
called a monoid. The natural numbers N (including 0) under addition form a monoid, as do the nonzero integers
under multiplication (Z \ {0}, ·), see above. There is a general method to formally add inverses to elements to any
(abelian) monoid, much the same way as (Q \ {0}, ·) is derived from (Z \ {0}, ·), known as the Grothendieck group.
Groupoids are similar to groups except that the composition a • b need not be defined for all a and b. They arise in
the study of more complicated forms of symmetry, often in topological and analytical structures, such as the
fundamental groupoid or stacks. Finally, it is possible to generalize any of these concepts by replacing the binary
operation with an arbitrary n-ary one (i.e. an operation taking n arguments). With the proper generalization of the
group axioms this gives rise to an n-ary group.[78] The table gives a list of several structures generalizing groups.

Notes
^ a: Mathematical Reviews lists 3,224 research papers on group theory and its generalizations written in 2005.
^ b: The closure axiom is already implied by the condition that • be a binary operation. Some authors therefore omit
this axiom. Lang 2002
^ c: See, for example, the books of Lang (2002, 2005) and Herstein (1996, 1975).
^ d: However, a group is not determined by its lattice of subgroups. See Suzuki 1951.
^ e: The fact that the group operation extends this canonically is an instance of a universal property.
^ f: For example, if G is finite, then the size of any subgroup and any quotient group divides the size of G, according
to Lagrange's theorem.
^ g: The word homomorphism derives from Greek ὁμός—the same and μορφή—structure.
^ h: The additive notation for elements of a cyclic group would be t • a, t in Z.
Group (mathematics) 16

^ i: See the Seifert–van Kampen theorem for an example.


^ j: An example is group cohomology of a group which equals the singular homology of its classifying space.
^ k: Elements which do have multiplicative inverses are called units, see Lang 2002, §II.1, p. 84.
^ l: The transition from the integers to the rationals by adding fractions is generalized by the quotient field.
^ m: The same is true for any field F instead of Q. See Lang 2005, §III.1, p. 86.
^ n: For example, a finite subgroup of the multiplicative group of a field is necessarily cyclic. See Lang 2002,
Theorem IV.1.9. The notions of torsion of a module and simple algebras are other instances of this principle.
^ o: The stated property is a possible definition of prime numbers. See prime element.
^ p: For example, the Diffie-Hellman protocol uses the discrete logarithm.
^ q: The groups of order at most 2000 are known. Up to isomorphism, there are about 49 billion. See Besche, Eick
& O'Brien 2001.
^ r: The gap between the classification of simple groups and the one of all groups lies in the extension problem, a
problem too hard to be solved in general. See Aschbacher 2004, p. 737.
^ s: Equivalently, a nontrivial group is simple if its only quotient groups are the trivial group and the group itself.
See Michler 2006, Carter 1989.
^ t: More rigorously, every group is the symmetry group of some graph; see Frucht's theorem, Frucht 1939.
^ u: More precisely, the monodromy action on the vector space of solutions of the differential equations is
considered. See Kuga 1993, pp. 105–113.
^ v: See Schwarzschild metric for an example where symmetry greatly reduces the complexity of physical systems.
^ w: This was crucial to the classification of finite simple groups, for example. See Aschbacher 2004.
^ x: See, for example, Schur's Lemma for the impact of a group action on simple modules. A more involved
example is the action of an absolute Galois group on étale cohomology.
^ y: Injective and surjective maps correspond to mono- and epimorphisms, respectively. They are interchanged when
passing to the dual category.

Citations
[1] Herstein 1975, §2, p. 26
[2] Hall 1967, §1.1, p. 1: "The idea of a group is one which pervades the whole of mathematics both pure and applied."
[3] Lang 2005, App. 2, p. 360
[4] Herstein 1975, §2.1, p. 27
[5] Weisstein, Eric W., " Identity Element (http:/ / mathworld. wolfram. com/ IdentityElement. html)" from MathWorld.
[6] Herstein 1975, §2.6, p. 54
[7] Wussing 2007
[8] Kleiner 1986
[9] Smith 1906
[10] Galois 1908
[11] Kleiner 1986, p. 202
[12] Cayley 1889
[13] Wussing 2007, §III.2
[14] Lie 1973
[15] Kleiner 1986, p. 204
[16] Wussing 2007, §I.3.4
[17] Jordan 1870
[18] von Dyck 1882
[19] Curtis 2003
[20] Mackey 1976
[21] Borel 2001
[22] Aschbacher 2004
[23] Ledermann 1953, §1.2, pp. 4–5
[24] Ledermann 1973, §I.1, p. 3
[25] Lang 2002, §I.2, p. 7
[26] Lang 2005, §II.1, p. 17
Group (mathematics) 17

[27] Mac Lane 1998


[28] Lang 2005, §II.3, p. 34
[29] Lang 2005, §II.1, p. 19
[30] Ledermann 1973, §II.12, p. 39
[31] Lang 2005, §II.4, p. 41
[32] Lang 2002, §I.2, p. 12
[33] Lang 2005, §II.4, p. 45
[34] Lang 2002, §I.2, p. 9
[35] Hatcher 2002, Chapter I, p. 30
[36] Coornaert, Delzant & Papadopoulos 1990
[37] for example, class groups and Picard groups; see Neukirch 1999, in particular §§I.12 and I.13
[38] Seress 1997
[39] Lang 2005, Chapter VII
[40] Rosen 2000, p. 54 (Theorem 2.1)
[41] Lang 2005, §VIII.1, p. 292
[42] Lang 2005, §II.1, p. 22
[43] Lang 2005, §II.2, p. 26
[44] Lang 2005, §II.1, p. 22 (example 11)
[45] Lang 2002, §I.5, p. 26, 29
[46] Weyl 1952
[47] Conway, Delgado Friedrichs & Huson et al. 2001. See also Bishop 1993
[48] Bersuker, Isaac (2006), The Jahn-Teller Effect, Cambridge University Press, p. 2, ISBN 0-521-82212-2
[49] Jahn & Teller 1937
[50] Dove, Martin T (2003), Structure and Dynamics: an atomic view of materials, Oxford University Press, p. 265, ISBN 0-19-850678-3
[51] Welsh 1989
[52] Mumford, Fogarty & Kirwan 1994
[53] Lay 2003
[54] Kuipers 1999
[55] Fulton & Harris 1991
[56] Serre 1977
[57] Rudin 1990
[58] Robinson 1996, p. viii
[59] Artin 1998
[60] Lang 2002, Chapter VI (see in particular p. 273 for concrete examples)
[61] Lang 2002, p. 292 (Theorem VI.7.2)
[62] Kurzweil & Stellmacher 2004
[63] Artin 1991, Theorem 6.1.14. See also Lang 2002, p. 77 for similar results.
[64] Lang 2002, §I. 3, p. 22
[65] Ronan 2007
[66] Husain 1966
[67] Neukirch 1999
[68] Shatz 1972
[69] Milne 1980
[70] Warner 1983
[71] Borel 1991
[72] Goldstein 1980
[73] Weinberg 1972
[74] Naber 2003
[75] Becchi 1997
[76] Denecke & Wismath 2002
[77] Romanowska & Smith 2002
[78] Dudek 2001
Group (mathematics) 18

References

General references
• Artin, Michael (1991), Algebra, Prentice Hall, ISBN 978-0-89871-510-1, Chapter 2 contains an
undergraduate-level exposition of the notions covered in this article.
• Devlin, Keith (2000), The Language of Mathematics: Making the Invisible Visible, Owl Books,
ISBN 978-0-8050-7254-9, Chapter 5 provides a layman-accessible explanation of groups.
• Fulton, William; Harris, Joe (1991), Representation theory. A first course, Graduate Texts in Mathematics,
Readings in Mathematics, 129, New York: Springer-Verlag, ISBN 978-0-387-97495-8, MR1153249, ISBN
978-0-387-97527-6.
• Hall, G. G. (1967), Applied group theory, American Elsevier Publishing Co., Inc., New York, MR0219593, an
elementary introduction.
• Herstein, Israel Nathan (1996), Abstract algebra (3rd ed.), Upper Saddle River, NJ: Prentice Hall Inc.,
ISBN 978-0-13-374562-7, MR1375019.
• Herstein, Israel Nathan (1975), Topics in algebra (2nd ed.), Lexington, Mass.: Xerox College Publishing,
MR0356988.
• Lang, Serge (2002), Algebra, Graduate Texts in Mathematics, 211 (Revised third ed.), New York:
Springer-Verlag, ISBN 978-0-387-95385-4, MR1878556
• Lang, Serge (2005), Undergraduate Algebra (3rd ed.), Berlin, New York: Springer-Verlag,
ISBN 978-0-387-22025-3.
• Ledermann, Walter (1953), Introduction to the theory of finite groups, Oliver and Boyd, Edinburgh and London,
MR0054593.
• Ledermann, Walter (1973), Introduction to group theory, New York: Barnes and Noble, OCLC 795613.
• Robinson, Derek John Scott (1996), A course in the theory of groups, Berlin, New York: Springer-Verlag,
ISBN 978-0-387-94461-6.

Special references
• Artin, Emil (1998), Galois Theory, New York: Dover Publications, ISBN 978-0-486-62342-9.
• Aschbacher, Michael (2004), "The Status of the Classification of the Finite Simple Groups" (http://www.ams.
org/notices/200407/fea-aschbacher.pdf) (PDF), Notices of the American Mathematical Society 51 (7): 736–740.
• Becchi, C. (1997), Introduction to Gauge Theories, arXiv:hep-ph/9705211.
• Besche, Hans Ulrich; Eick, Bettina; O'Brien, E. A. (2001), "The groups of order at most 2000" (http://www.
ams.org/era/2001-07-01/S1079-6762-01-00087-7/home.html), Electronic Research Announcements of the
American Mathematical Society 7: 1–4, doi:10.1090/S1079-6762-01-00087-7, MR1826989.
• Bishop, David H. L. (1993), Group theory and chemistry, New York: Dover Publications,
ISBN 978-0-486-67355-4.
• Borel, Armand (1991), Linear algebraic groups, Graduate Texts in Mathematics, 126 (2nd ed.), Berlin, New
York: Springer-Verlag, ISBN 978-0-387-97370-8, MR1102012.
• Carter, Roger W. (1989), Simple groups of Lie type, New York: John Wiley & Sons, ISBN 978-0-471-50683-6.
• Conway, John Horton; Delgado Friedrichs, Olaf; Huson, Daniel H.; Thurston, William P. (2001), "On
three-dimensional space groups", Beiträge zur Algebra und Geometrie 42 (2): 475–507,
arXiv:math.MG/9911185, MR1865535.
• (French) Coornaert, M.; Delzant, T.; Papadopoulos, A. (1990), Géométrie et théorie des groupes [Geometry and
Group Theory], Lecture Notes in Mathematics, 1441, Berlin, New York: Springer-Verlag,
ISBN 978-3-540-52977-4, MR1075994.
• Denecke, Klaus; Wismath, Shelly L. (2002), Universal algebra and applications in theoretical computer science,
London: CRC Press, ISBN 978-1-58488-254-1.
Group (mathematics) 19

• Dudek, W.A. (2001), "On some old problems in n-ary groups" (http://www.quasigroups.eu/contents/
contents8.php?m=trzeci), Quasigroups and Related Systems 8: 15–36.
• (German) Frucht, R. (1939), "Herstellung von Graphen mit vorgegebener abstrakter Gruppe [Construction of
Graphs with Prescribed Group (http://www.numdam.org/numdam-bin/fitem?id=CM_1939__6__239_0)"],
Compositio Mathematica 6: 239–50.
• Goldstein, Herbert (1980), Classical Mechanics (2nd ed.), Reading, MA: Addison-Wesley Publishing,
pp. 588–596, ISBN 0-201-02918-9.
• Hatcher, Allen (2002), Algebraic topology (http://www.math.cornell.edu/~hatcher/AT/ATpage.html),
Cambridge University Press, ISBN 978-0-521-79540-1.
• Husain, Taqdir (1966), Introduction to Topological Groups, Philadelphia: W.B. Saunders Company,
ISBN 978-0-89874-193-3
• Jahn, H.; Teller, E. (1937), "Stability of Polyatomic Molecules in Degenerate Electronic States. I. Orbital
Degeneracy", Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences
(1934–1990) 161 (905): 220–235, doi:10.1098/rspa.1937.0142.
• Kuipers, Jack B. (1999), Quaternions and rotation sequences—A primer with applications to orbits, aerospace,
and virtual reality, Princeton University Press, ISBN 978-0-691-05872-6, MR1670862.
• Kuga, Michio (1993), Galois' dream: group theory and differential equations, Boston, MA: Birkhäuser Boston,
ISBN 978-0-8176-3688-3, MR1199112.
• Kurzweil, Hans; Stellmacher, Bernd (2004), The theory of finite groups, Universitext, Berlin, New York:
Springer-Verlag, ISBN 978-0-387-40510-0, MR2014408.
• Lay, David (2003), Linear Algebra and Its Applications, Addison-Wesley, ISBN 978-0-201-70970-4.
• Mac Lane, Saunders (1998), Categories for the Working Mathematician (2nd ed.), Berlin, New York:
Springer-Verlag, ISBN 978-0-387-98403-2.
• Michler, Gerhard (2006), Theory of finite simple groups, Cambridge University Press, ISBN 978-0-521-86625-5.
• Milne, James S. (1980), Étale cohomology, Princeton University Press, ISBN 978-0-691-08238-7
• Mumford, David; Fogarty, J.; Kirwan, F. (1994), Geometric invariant theory, 34 (3rd ed.), Berlin, New York:
Springer-Verlag, ISBN 978-3-540-56963-3, MR1304906.
• Naber, Gregory L. (2003), The geometry of Minkowski spacetime, New York: Dover Publications,
ISBN 978-0-486-43235-9, MR2044239.
• Neukirch, Jürgen (1999), Algebraic Number Theory, Grundlehren der mathematischen Wissenschaften, 322,
Berlin: Springer-Verlag, ISBN 978-3-540-65399-8, MR1697859.
• Romanowska, A.B.; Smith, J.D.H. (2002), Modes, World Scientific, ISBN 978-981-02-4942-7.
• Ronan, Mark (2007), Symmetry and the Monster: The Story of One of the Greatest Quests of Mathematics,
Oxford University Press, ISBN 978-0-19-280723-6.
• Rosen, Kenneth H. (2000), Elementary number theory and its applications (4th ed.), Addison-Wesley,
ISBN 978-0-201-87073-2, MR1739433.
• Rudin, Walter (1990), Fourier Analysis on Groups, Wiley Classics, Wiley-Blackwell, ISBN 0-471-52364-X.
• Seress, Ákos (1997), "An introduction to computational group theory" (http://www.math.ohio-state.edu/
~akos/notices.ps), Notices of the American Mathematical Society 44 (6): 671–679, MR1452069.
• Serre, Jean-Pierre (1977), Linear representations of finite groups, Berlin, New York: Springer-Verlag,
ISBN 978-0-387-90190-9, MR0450380.
• Shatz, Stephen S. (1972), Profinite groups, arithmetic, and geometry, Princeton University Press,
ISBN 978-0-691-08017-8, MR0347778
• Suzuki, Michio (1951), "On the lattice of subgroups of finite groups", Transactions of the American
Mathematical Society 70 (2): 345–371, doi:10.2307/1990375, JSTOR 1990375.
• Warner, Frank (1983), Foundations of Differentiable Manifolds and Lie Groups, Berlin, New York:
Springer-Verlag, ISBN 978-0-387-90894-6.
Group (mathematics) 20

• Weinberg, Steven (1972), Gravitation and Cosmology, New York: John Wiley & Sons, ISBN 0-471-92567-5.
• Welsh, Dominic (1989), Codes and cryptography, Oxford: Clarendon Press, ISBN 978-0-19-853287-3.
• Weyl, Hermann (1952), Symmetry, Princeton University Press, ISBN 978-0-691-02374-8.

Historical references
• Borel, Armand (2001), Essays in the History of Lie Groups and Algebraic Groups, Providence, R.I.: American
Mathematical Society, ISBN 978-0-8218-0288-5
• Cayley, Arthur (1889), The collected mathematical papers of Arthur Cayley (http://www.hti.umich.edu/cgi/t/
text/pageviewer-idx?c=umhistmath;cc=umhistmath;rgn=full text;idno=ABS3153.0001.001;didno=ABS3153.
0001.001;view=image;seq=00000140), II (1851–1860), Cambridge University Press.
• O'Connor, J.J; Robertson, E.F. (1996), The development of group theory (http://www-groups.dcs.st-and.ac.uk/
~history/HistTopics/Development_group_theory.html).
• Curtis, Charles W. (2003), Pioneers of Representation Theory: Frobenius, Burnside, Schur, and Brauer, History
of Mathematics, Providence, R.I.: American Mathematical Society, ISBN 978-0-8218-2677-5.
• (German) von Dyck, Walther (1882), "Gruppentheoretische Studien (Group-theoretical Studies)", Mathematische
Annalen 20 (1): 1–44, doi:10.1007/BF01443322.
• (French) Galois, Évariste (1908), Tannery, Jules, ed., Manuscrits de Évariste Galois [Évariste Galois'
Manuscripts (http://quod.lib.umich.edu/cgi/t/text/text-idx?c=umhistmath;idno=AAN9280)], Paris:
Gauthier-Villars (Galois work was first published by Joseph Liouville in 1843).
• (French) Jordan, Camille (1870), Traité des substitutions et des équations algébriques [Study of Substitutions and
Algebraic Equations (http://gallica.bnf.fr/notice?N=FRBNF35001297)], Paris: Gauthier-Villars .
• Kleiner, Israel (1986), "The evolution of group theory: a brief survey", Mathematics Magazine 59 (4): 195–215,
doi:10.2307/2690312, MR863090.
• (German) Lie, Sophus (1973), Gesammelte Abhandlungen. Band 1 [Collected papers. Volume 1], New York:
Johnson Reprint Corp., MR0392459.
• Mackey, George Whitelaw (1976), The theory of unitary group representations, University of Chicago Press,
MR0396826
• Smith, David Eugene (1906), History of Modern Mathematics (http://www.gutenberg.org/etext/8746),
Mathematical Monographs, No. 1.
• Wussing, Hans (2007), The Genesis of the Abstract Group Concept: A Contribution to the History of the Origin of
Abstract Group Theory, New York: Dover Publications, ISBN 978-0-486-45868-7.
Orthogonal group 21

Orthogonal group
In mathematics, the orthogonal group of degree n over a field F (written as O(n,F)) is the group of n-by-n
orthogonal matrices with entries from F, with the group operation of matrix multiplication. This is a subgroup of the
general linear group GL(n,F) given by

where QT is the transpose of Q. The classical orthogonal group over the real numbers is usually just written O(n).
More generally the orthogonal group of a non-singular quadratic form over F is the group of linear operators
preserving the form – the above group O(n, F) is then the orthogonal group of the sum-of-n-squares quadratic
form.[1] The Cartan–Dieudonné theorem describes the structure of the orthogonal group for non-singular form. This
article only discusses definite forms – the orthogonal group of the positive definite form (equivalent to sum of n
squares) and negative definite forms (equivalent to the negative sum of n squares) are identical –
– though the associated Pin groups differ; for other non-singular forms O(p,q), see indefinite
orthogonal group.
Every orthogonal matrix has determinant either 1 or −1. The orthogonal n-by-n matrices with determinant 1 form a
normal subgroup of O(n,F) known as the special orthogonal group, SO(n,F). (More precisely, SO(n,F) is the kernel
of the Dickson invariant, discussed below.) By analogy with GL/SL (general linear group, special linear group), the
orthogonal group is sometimes called the general orthogonal group and denoted GO, though this term is also
sometimes used for indefinite orthogonal groups O(p,q).
The derived subgroup Ω(n,F) of O(n,F) is an often studied object because when F is a finite field Ω(n,F) is often a
central extension of a finite simple group.
Both O(n,F) and SO(n,F) are algebraic groups, because the condition that a matrix be orthogonal, i.e. have its own
transpose as inverse, can be expressed as a set of polynomial equations in the entries of the matrix.

Over the real number field


Over the field R of real numbers, the orthogonal group O(n,R) and the special orthogonal group SO(n,R) are often
simply denoted by O(n) and SO(n) if no confusion is possible. They form real compact Lie groups of dimension
n(n − 1)/2. O(n,R) has two connected components, with SO(n,R) being the identity component, i.e., the connected
component containing the identity matrix.
The real orthogonal and real special orthogonal groups have the following geometric interpretations:
O(n,R) is a subgroup of the Euclidean group E(n), the group of isometries of Rn; it contains those that leave the
origin fixed – O(n,R) = E(n) ∩ GL(n,R). It is the symmetry group of the sphere (n = 3) or hypersphere and all
objects with spherical symmetry, if the origin is chosen at the center.
SO(n,R) is a subgroup of E+(n), which consists of direct isometries, i.e., isometries preserving orientation; it contains
those that leave the origin fixed – It is the
rotation group of the sphere and all objects with spherical symmetry, if the origin is chosen at the center.
{ I, −I } is a normal subgroup and even a characteristic subgroup of O(n,R), and, if n is even, also of SO(n,R). If n is
odd, O(n,R) is the direct product of SO(n,R) and { I, −I }. The cyclic group of k-fold rotations Ck is for every
positive integer k a normal subgroup of O(2,R) and SO(2,R).
Relative to suitable orthogonal bases, the isometries are of the form:
Orthogonal group 22

where the matrices R1,...,Rk are 2-by-2 rotation matrices in orthogonal planes of rotation. As a special case, known as
Euler's rotation theorem, any (non-identity) element of SO(3,R) is rotation about a uniquely defined axis.
The orthogonal group is generated by reflections (two reflections give a rotation), as in a Coxeter group,[2] and
elements have length at most n (require at most n reflections to generate; this follows from the above classification,
noting that a rotation is generated by 2 reflections, and is true more generally for indefinite orthogonal groups, by the
Cartan–Dieudonné theorem). A longest element (element needing the most reflections) is reflection through the
origin (the map ), though so are other maximal combinations of rotations (and a reflection, in odd
dimension).
The symmetry group of a circle is O(2,R), also called Dih (S1), where S1 denotes the multiplicative group of complex
numbers of absolute value 1.
SO(2,R) is isomorphic (as a Lie group) to the circle S1 (circle group). This isomorphism sends the complex number
exp(φi) = cos(φ) + i sin(φ) to the orthogonal matrix

The group SO(3,R), understood as the set of rotations of 3-dimensional space, is of major importance in the sciences
and engineering. See rotation group and the general formula for a 3 × 3 rotation matrix in terms of the axis and the
angle.
In terms of algebraic topology, for n > 2 the fundamental group of SO(n,R) is cyclic of order 2, and the spinor group
Spin(n) is its universal cover. For n = 2 the fundamental group is infinite cyclic and the universal cover corresponds
to the real line (the spinor group Spin(2) is the unique 2-fold cover).

Even and odd dimension


The structure of the orthogonal group differs in certain respects between even and odd dimensions – for example,
(reflection through the origin) is orientation-preserving in even dimension, but orientation-reversing in odd
dimension. When this distinction wishes to be emphasized, the groups are generally denoted O(2k) and O(2k+1),
reserving n for the dimension of the space ( or ). The letters p or r are also used, indicating
the rank of the corresponding Lie algebra; in odd dimension the corresponding Lie algebra is while
in even dimension the Lie algebra is
Orthogonal group 23

Lie algebra
The Lie algebra associated to the Lie groups O(n,R) and SO(n,R) consists of the skew-symmetric real n-by-n
matrices, with the Lie bracket given by the commutator. This Lie algebra is often denoted by o(n,R) or by so(n,R),
and called the orthogonal Lie algebra or special orthogonal Lie algebra. These Lie algebras are the compact real
forms of two of the four families of semisimple Lie algebras: in odd dimension while in even
dimension
More intrinsically, given a vector space with an inner product, the special orthogonal Lie algebra is given by the
bivectors on the space, which are sums of simple bivectors (2-blades) . The correspondence is given by the
map where is the covector dual to the vector v; in coordinates these are exactly
the elementary skew-symmetric matrices.
This characterization is used in interpreting the curl of a vector field (naturally a 2-vector) as an infinitesimal rotation
or "curl", hence the name. Generalizing the inner product with a nondegenerate form yields the indefinite orthogonal
Lie algebras
The representation theory of the orthogonal Lie algebras includes both representations corresponding to linear
representations of the orthogonal groups, and representations corresponding to projective representations of the
orthogonal groups (linear representations of spin groups), the so-called spin representation, which are important in
physics.

3D isometries that leave the origin fixed


Isometries of R3 that leave the origin fixed, forming the group O(3,R), can be categorized as:
• SO(3,R):
• identity
• rotation about an axis through the origin by an angle not equal to 180°
• rotation about an axis through the origin by an angle of 180°
• the same with inversion in the origin (x is mapped to −x), i.e. respectively:
• inversion in the origin
• rotation about an axis by an angle not equal to 180°, combined with reflection in the plane through the origin
perpendicular to the axis
• reflection in a plane through the origin.
The 4th and 5th in particular, and in a wider sense the 6th also, are called improper rotations.
See also the similar overview including translations.

Conformal group
Being isometries (preserving distances), orthogonal transforms also preserve angles, and are thus conformal maps,
though not all conformal linear transforms are orthogonal. In classical terms this is the difference between
congruence and similarity, as exemplified by SSS (Side-Side-Side) congruence of triangles and AAA
(Angle-Angle-Angle) similarity of triangles. The group of conformal linear maps of Rn is denoted CO(n) for the
conformal orthogonal group, and consists of the product of the orthogonal group with the group of dilations. If n is
odd, these two subgroups do not intersect, and they are a direct product: ,
while if n is even, these subgroups intersect in , so this is not a direct product, but it is a direct product with the
subgroup of dilation by a positive scalar: .
Similarly one can define CSO(n); note that this is always : .
Orthogonal group 24

Over the complex number field


Over the field C of complex numbers, O(n,C) and SO(n,C) are complex Lie groups of dimension n(n − 1)/2 over C
(which means the dimension over R is twice that). O(n,C) has two connected components, and SO(n,C) is the
connected component containing the identity matrix. For n ≥ 2 these groups are noncompact.
Just as in the real case SO(n,C) is not simply connected. For n > 2 the fundamental group of SO(n,C) is cyclic of
order 2 whereas the fundamental group of SO(2,C) is infinite cyclic.
The complex Lie algebra associated to O(n,C) and SO(n,C) consists of the skew-symmetric complex n-by-n
matrices, with the Lie bracket given by the commutator.

Topology

Low dimensional
The low dimensional (real) orthogonal groups are familiar spaces:

The group is double covered by .


There are numerous charts on SO(3), due to the importance of 3-dimensional rotations in engineering applications.
Here Sn denotes the n-dimensional sphere, RPn the n-dimensional real projective space, and SU(n) the special unitary
group of degree n.

Homotopy groups
The homotopy groups of the orthogonal group are related to homotopy groups of spheres, and thus are in general
hard to compute.
However, one can compute the homotopy groups of the stable orthogonal group (aka the infinite orthogonal group),
defined as the direct limit of the sequence of inclusions

(as the inclusions are all closed inclusions, hence cofibrations, this can also be interpreted as a union).
is a homogeneous space for , and one has the following fiber bundle:

which can be understood as "The orthogonal group acts transitively on the unit sphere , and the
stabilizer of a point (thought of as a unit vector) is the orthogonal group of the perpendicular complement, which is
an orthogonal group one dimension lower". The map is the natural inclusion.
Thus the inclusion is (n − 1)-connected, so the homotopy groups stabilize, and
for : thus the homotopy groups of the stable space equal the lower homotopy
groups of the unstable spaces.
Via Bott periodicity, , thus the homotopy groups of O are 8-fold periodic, meaning ,
and one needs only to compute the lower 8 homotopy groups to compute them all.
Orthogonal group 25

Relation to KO-theory
Via the clutching construction, homotopy groups of the stable space O are identified with stable vector bundles on
spheres (up to isomorphism), with a dimension shift of 1: .
Setting (to make fit into the periodicity), one obtains:

Computation and Interpretation of homotopy groups

Low-dimensional groups
The first few homotopy groups can be calculated by using the concrete descriptions of low-dimensional groups.
• from orientation-preserving/reversing (this class survives to and hence
stably)
yields
• , which is spin
• , which surjects onto ; this latter thus vanishes.

Lie groups
From general facts about Lie groups, always vanishes, and is free (free abelian).

Vector bundles
From the vector bundle point of view, is vector bundles over , which is two points. Thus over each
point, the bundle is trivial, and the non-triviality of the bundle is the difference between the dimensions of the vector
spaces over the two points, so
is dimension.
Orthogonal group 26

Loop spaces
Using concrete descriptions of the loop spaces in Bott periodicity, one can interpret higher homotopy of O as lower
homotopy of simple to analyze spaces. Using , O and O/U have two components, and
have components, and the rest are connected.

Interpretation of homotopy groups


In a nutshell:[3]
• is dimension
• is orientation
• is spin
• is topological quantum field theory.
Let , and let be the tautological line bundle over the projective line , and its
class in K-theory. Noting that , these yield vector bundles
over the corresponding spheres, and
• is generated by
• is generated by
• is generated by
• is generated by
From the point of view of symplectic geometry, can be interpreted as the Maslov
index, thinking of it as the fundamental group of the stable Lagrangian Grassmannian as
so

Over finite fields


Orthogonal groups can also be defined over finite fields , where is a power of a prime . When defined over
such fields, they come in two types in even dimension: and ; and one type in odd
dimension: .
If is the vector space on which the orthogonal group acts, it can be written as a direct orthogonal sum as
follows:

where are hyperbolic lines and contains no singular vectors. If , then is of plus type. If
then has odd dimension. If has dimension 2, is of minus type.
In the special case where n = 1, is a dihedral group of order .
We have the following formulas for the order of these groups, O(n,q) = { A in GL(n,q) : A·At=I }, when the
characteristic is greater than two

If −1 is a square in

If −1 is a nonsquare in
Orthogonal group 27

The Dickson invariant


For orthogonal groups, the Dickson invariant is a homomorphism from the orthogonal group to Z/2Z, and is 0 or 1
depending on whether an element is the product of an even or odd number of reflections. More concretely, the
Dickson invariant can be defined as where I is the identity (Taylor 1992,
Theorem 11.43). Over fields that are not of characteristic 2 it is equivalent to the determinant: the determinant is −1
to the power of the Dickson invariant. Over fields of characteristic 2, the determinant is always 1, so the Dickson
invariant gives extra information.
The special orthogonal group is the kernel of the Dickson invariant and usually has index 2 in O(n,F).[4] When the
characteristic of F is not 2, the Dickson Invariant is 0 whenever the determinant is 1. Thus when the characteristic is
not 2, SO(n,F) is commonly defined to be the elements of O(n,F) with determinant 1. Each element in O(n,F) has
determinant −1 or 1. Thus in characteristic 2, the determinant is always 1.
The Dickson invariant can also be defined for Clifford groups and Pin groups in a similar way (in all dimensions).

Orthogonal groups of characteristic 2


Over fields of characteristic 2 orthogonal groups often behave differently. This section lists some of the differences.
Traditionally these groups are known as the hypoabelian groups but this term is no longer used for these groups.
• Any orthogonal group over any field is generated by reflections, except for a unique example where the vector
space is 4 dimensional over the field with 2 elements and the Witt index is 2.[5] Note that a reflection in
characteristic two has a slightly different definition. In characteristic two, the reflection orthogonal to a vector u
takes a vector v to v+B(v,u)/Q(u)·u where B is the bilinear form and Q is the quadratic form associated to the
orthogonal geometry. Compare this to the Householder reflection of odd characteristic or characteristic zero,
which takes v to v − 2·B(v,u)/Q(u)·u.
• The center of the orthogonal group usually has order 1 in characteristic 2, rather than 2, since
• In odd dimensions 2n+1 in characteristic 2, orthogonal groups over perfect fields are the same as symplectic
groups in dimension 2n. In fact the symmetric form is alternating in characteristic 2, and as the dimension is odd
it must have a kernel of dimension 1, and the quotient by this kernel is a symplectic space of dimension 2n, acted
upon by the orthogonal group.
• In even dimensions in characteristic 2 the orthogonal group is a subgroup of the symplectic group, because the
symmetric bilinear form of the quadratic form is also an alternating form.

The spinor norm


The spinor norm is a homomorphism from an orthogonal group over a field F to
F*/F*2,
the multiplicative group of the field F up to square elements, that takes reflection in a vector of norm n to the image
of n in F*/F*2.
For the usual orthogonal group over the reals it is trivial, but it is often non-trivial over other fields, or for the
orthogonal group of a quadratic form over the reals that is not positive definite.
Orthogonal group 28

Galois cohomology and orthogonal groups


In the theory of Galois cohomology of algebraic groups, some further points of view are introduced. They have
explanatory value, in particular in relation with the theory of quadratic forms; but were for the most part post hoc, as
far as the discovery of the phenomena is concerned. The first point is that quadratic forms over a field can be
identified as a Galois H1, or twisted forms (torsors) of an orthogonal group. As an algebraic group, an orthogonal
group is in general neither connected nor simply-connected; the latter point brings in the spin phenomena, while the
former is related to the discriminant.
The 'spin' name of the spinor norm can be explained by a connection to the spin group (more accurately a pin group).
This may now be explained quickly by Galois cohomology (which however postdates the introduction of the term by
more direct use of Clifford algebras). The spin covering of the orthogonal group provides a short exact sequence of
algebraic groups.

Here μ2 is the algebraic group of square roots of 1; over a field of characteristic not 2 it is roughly the same as a
two-element group with trivial Galois action. The connecting homomorphism from H0(OV), which is simply the
group OV(F) of F-valued points, to H1(μ2) is essentially the spinor norm, because H1(μ2) is isomorphic to the
multiplicative group of the field modulo squares.
There is also the connecting homomorphism from H1 of the orthogonal group, to the H2 of the kernel of the spin
covering. The cohomology is non-abelian, so that this is as far as we can go, at least with the conventional
definitions.

Related groups
The orthogonal groups and special orthogonal groups have a number of important subgroups, supergroups, quotient
groups, and covering groups. These are listed below.
The inclusions and are part of a
sequence of 8 inclusions used in a geometric proof of the Bott periodicity theorem, and the corresponding quotient
spaces are symmetric spaces of independent interest – for example, is the Lagrangian Grassmannian.

Lie subgroups
In physics, particularly in the areas of Kaluza–Klein compactification, it is important to find out the subgroups of the
orthogonal group. The main ones are:
– preserves an axis
– U(n) are those that preserve a compatible complex structure or a compatible
symplectic structure – see 2-out-of-3 property; SU(n) also preserves a complex orientation.
Orthogonal group 29

Lie supergroups
The orthogonal group O(n) is also an important subgroup of various Lie groups:

Discrete subgroups
As the orthogonal group is compact, discrete subgroups are equivalent to finite subgroups.[6] These subgroups are
known as point group and can be realized as the symmetry groups of polytopes. A very important class of examples
are the finite Coxeter groups, which include the symmetry groups of regular polytopes.
Dimension 3 is particularly studied – see point groups in three dimensions, polyhedral groups, and list of spherical
symmetry groups. In 2 dimensions, the finite groups are either cyclic or dihedral – see point groups in two
dimensions.
Other finite subgroups include:
• Permutation matrices (the Coxeter group An)
• Signed permutation matrices (the Coxeter group Bn); also equals the intersection of the orthogonal group with the
integer matrices.[7]

Covering and quotient groups


The orthogonal group is neither simply connected nor centerless, and thus has both a covering group and a quotient
group, respectively:
• Two covering Pin groups, Pin+(n) → O(n) and Pin−(n) → O(n),
• The quotient projective orthogonal group, O(n) → PO(n).
These are all 2-to-1 covers.
For the special orthogonal group, the corresponding groups are:
• Spin group, Spin(n) → SO(n),
• Projective special orthogonal group, SO(n) → PSO(n).
Spin is a 2-to-1 cover, while in even dimension, PSO(2k) is a 2-to-1 cover, and in odd dimension PSO(2k+1) is a
1-to-1 cover, i.e., isomorphic to SO(2k+1). These groups, Spin(n), SO(n), and PSO(n) are Lie group forms of the
compact special orthogonal Lie algebra, – Spin is the simply connected form, while PSO is the centerless
form, and SO is in general neither.[8]
In dimension 3 and above these are the covers and quotients, while dimension 2 and below are somewhat degenerate;
see specific articles for details.
Orthogonal group 30

Applications to string theory


The group O(10) is of special importance in superstring theory because it is the symmetry group of 10 dimensional
space-time.

Principal homogeneous space: Stiefel manifold


The principal homogeneous space for the orthogonal group O(n) is the Stiefel manifold of orthonormal
bases (orthonormal n-frames).
In other words, the space of orthonormal bases is like the orthogonal group, but without a choice of base point: given
an orthogonal space, there is no natural choice of orthonormal basis, but once one is given one, there is a one-to-one
correspondence between bases and the orthogonal group. Concretely, a linear map is determined by where it sends a
basis: just as an invertible map can take any basis to any other basis, an orthogonal map can take any orthogonal
basis to any other orthogonal basis.
The other Stiefel manifolds for of incomplete orthonormal bases (orthonormal k-frames) are still
homogeneous spaces for the orthogonal group, but not principal homogeneous spaces: any k-frame can be taken to
any other k-frame by an orthogonal map, but this map is not uniquely determined.

Notes
[1] Away from 2, it is equivalent to use bilinear forms or quadratic forms, but at 2 these differ – notably in characteristic 2, but also when
generalizing to rings where 2 is not invertible, most significantly the integers, where the notions of even and odd quadratic forms arise.
[2] The analogy is stronger: Weyl groups, a class of (representations of) Coxeter groups, can be considered as simple algebraic groups over the
field with one element, and there are a number of analogies between algebraic groups and vector spaces on the one hand, and Weyl groups and
sets on the other.
[3] John Baez "This Week's Finds in Mathematical Physics" week 105 (http:/ / math. ucr. edu/ home/ baez/ week105. html)
[4] (Taylor 1992, page 160)
[5] (Grove 2002, Theorem 6.6 and 14.16)
[6] Infinite subsets of a compact space have an accumulation point and are not discrete.
[7] equals the signed permutation matrices because an integer vector of norm 1 must have a single non-zero entry,
which must be ±1 (if it has two non-zero entries or a larger entry, the norm will be larger than 1), and in an orthogonal matrix these entries
must be in different coordinates, which is exactly the signed permutation matrices.
[8] In odd dimension, SO(2k+1)   PSO(2k+1) is centerless (but not simply connected), while in even dimension SO(2k) is neither centerless
nor simply connected.

References
• Grove, Larry C. (2002), Classical groups and geometric algebra, Graduate Studies in Mathematics, 39,
Providence, R.I.: American Mathematical Society, ISBN 978-0-8218-2019-3, MR1859189
• Taylor, Donald E. (1992), The Geometry of the Classical Groups, 9, Berlin: Heldermann Verlag,
ISBN 3-88538-009-9, MR1189139

External links
• John Baez "This Week's Finds in Mathematical Physics" week 105 (http://math.ucr.edu/home/baez/week105.
html)
• John Baez on Octonions (http://math.ucr.edu/home/baez/octonions/node10.html)
• (Italian) n-dimensional Special Orthogonal Group parametrization (http://ansi.altervista.org)
Lie group 31

Lie group
In mathematics, a Lie group ( /ˈliː/) is a group which is also a differentiable manifold, with the property that the
group operations are compatible with the smooth structure. Lie groups are named after Sophus Lie, who laid the
foundations of the theory of continuous transformation groups.
Lie groups represent the best-developed theory of continuous symmetry of mathematical objects and structures,
which makes them indispensable tools for many parts of contemporary mathematics, as well as for modern
theoretical physics. They provide a natural framework for analysing the continuous symmetries of differential
equations (Differential Galois theory), in much the same way as permutation groups are used in Galois theory for
analysing the discrete symmetries of algebraic equations. An extension of Galois theory to the case of continuous
symmetry groups was one of Lie's principal motivations.

Overview
Lie groups are smooth manifolds and, therefore, can be studied using
differential calculus, in contrast with the case of more general
topological groups. One of the key ideas in the theory of Lie groups,
from Sophus Lie, is to replace the global object, the group, with its
local or linearized version, which Lie himself called its "infinitesimal
group" and which has since become known as its Lie algebra.

Lie groups play an enormous role in modern geometry, on several


different levels. Felix Klein argued in his Erlangen program that one
can consider various "geometries" by specifying an appropriate
transformation group that leaves certain geometric properties invariant.
Thus Euclidean geometry corresponds to the choice of the group E(3)
of distance-preserving transformations of the Euclidean space R3, The circle of center 0 and radius 1 in the complex
conformal geometry corresponds to enlarging the group to the plane is a Lie group with complex multiplication.

conformal group, whereas in projective geometry one is interested in


the properties invariant under the projective group. This idea later led to the notion of a G-structure, where G is a Lie
group of "local" symmetries of a manifold. On a "global" level, whenever a Lie group acts on a geometric object,
such as a Riemannian or a symplectic manifold, this action provides a measure of rigidity and yields a rich algebraic
structure. The presence of continuous symmetries expressed via a Lie group action on a manifold places strong
constraints on its geometry and facilitates analysis on the manifold. Linear actions of Lie groups are especially
important, and are studied in representation theory.

In the 1940s–1950s, Ellis Kolchin, Armand Borel and Claude Chevalley realised that many foundational results
concerning Lie groups can be developed completely algebraically, giving rise to the theory of algebraic groups
defined over an arbitrary field. This insight opened new possibilities in pure algebra, by providing a uniform
construction for most finite simple groups, as well as in algebraic geometry. The theory of automorphic forms, an
important branch of modern number theory, deals extensively with analogues of Lie groups over adele rings; p-adic
Lie groups play an important role, via their connections with Galois representations in number theory.
Lie group 32

Definitions and examples


A real Lie group is a group which is also a finite-dimensional real smooth manifold, and in which the group
operations of multiplication and inversion are smooth maps. Smoothness of the group multiplication

means that μ is a smooth mapping of the product manifold G×G into G. These two requirements can be combined to
the single requirement that the mapping

be a smooth mapping of the product manifold into G.

First examples
• The 2×2 real invertible matrices form a group under multiplication, denoted by GL2(R):

This is a four-dimensional noncompact real Lie group. This group is disconnected; it has two connected
components corresponding to the positive and negative values of the determinant.
• The rotation matrices form a subgroup of GL2(R), denoted by SO2(R). It is a Lie group in its own right:
specifically, a one-dimensional compact connected Lie group which is diffeomorphic to the circle. Using the
rotation angle as a parameter, this group can be parametrized as follows:

Addition of the angles corresponds to multiplication of the elements of SO2(R), and taking the opposite angle
corresponds to inversion. Thus both multiplication and inversion are differentiable maps.
• The orthogonal group also forms an interesting example of a Lie group.
All of the previous examples of Lie groups fall within the class of classical groups

Related concepts
A complex Lie group is defined in the same way using complex manifolds rather than real ones (example: SL2(C)),
and similarly one can define a p-adic Lie group over the p-adic numbers. Hilbert's fifth problem asked whether
replacing differentiable manifolds with topological or analytic ones can yield new examples. The answer to this
question turned out to be negative: in 1952, Gleason, Montgomery and Zippin showed that if G is a topological
manifold with continuous group operations, then there exists exactly one analytic structure on G which turns it into a
Lie group (see also Hilbert–Smith conjecture). If the underlying manifold is allowed to be infinite dimensional (for
example, a Hilbert manifold) then one arrives at the notion of an infinite-dimensional Lie group. It is possible to
define analogues of many Lie groups over finite fields, and these give most of the examples of finite simple groups.
The language of category theory provides a concise definition for Lie groups: a Lie group is a group object in the
category of smooth manifolds. This is important, because it allows generalization of the notion of a Lie group to Lie
supergroups.
Lie group 33

More examples of Lie groups


Lie groups occur in abundance throughout mathematics and physics. Matrix groups or algebraic groups are (roughly)
groups of matrices (for example, orthogonal and symplectic groups), and these give most of the more common
examples of Lie groups.

Examples
• Euclidean space Rn with ordinary vector addition as the group operation becomes an n-dimensional noncompact
abelian Lie group.
• The circle group S1 consisting of angles mod 2π under addition or, alternately, the complex numbers with
absolute value 1 under multiplication. This is a one-dimensional compact connected abelian Lie group.
• The group GLn(R) of invertible matrices (under matrix multiplication) is a Lie group of dimension n2, called the
general linear group. It has a closed connected subgroup SLn(R), the special linear group, consisting of matrices
of determinant 1 which is also a Lie group.
• The orthogonal group On(R), consisting of all n × n orthogonal matrices with real entries is an n(n −
1)/2-dimensional Lie group. This group is disconnected, but it has a connected subgroup SOn(R) of the same
dimension consisting of orthogonal matrices of determinant 1, called the special orthogonal group (for n = 3, the
rotation group).
• The Euclidean group En(R) is the Lie group of all Euclidean motions, i.e., isometric affine maps, of
n-dimensional Euclidean space Rn.
• The unitary group U(n) consisting of n × n unitary matrices (with complex entries) is a compact connected Lie
group of dimension n2. Unitary matrices of determinant 1 form a closed connected subgroup of dimension n2 − 1
denoted SU(n), the special unitary group.
• Spin groups are double covers of the special orthogonal groups, used for studying fermions in quantum field
theory (among other things).
• The symplectic group Sp2n(R) consists of all 2n × 2n matrices preserving a nondegenerate skew-symmetric
bilinear form on R2n (the symplectic form). It is a connected Lie group of dimension 2n2 + n. The fundamental
group of the symplectic group is Z and this fact is related to the theory of Maslov index.
• The 3-sphere S3 forms a Lie group by identification with the set of quaternions of unit norm, called versors. The
only other spheres that admit the structure of a Lie group are the 0-sphere S0 (real numbers with absolute value 1)
and the circle S1 (complex numbers with absolute value 1). For example, for even n > 1, Sn is not a Lie group
because it does not admit a nonvanishing vector field and so a fortiori cannot be parallelizable as a differentiable
manifold. Of the spheres only S0, S1, S3, and S7 are parallelizable. The latter carries the structure of a Lie
quasigroup (a nonassociative group), which can be identified with the set of unit octonions.
• The group of upper triangular n by n matrices is a solvable Lie group of dimension n(n + 1)/2.
• The Lorentz group and the Poincare group are the groups of linear and affine isometries of the Minkowski space
(interpreted as the spacetime of the special relativity). They are Lie groups of dimensions 6 and 10.
• The Heisenberg group is a connected nilpotent Lie group of dimension 3, playing a key role in quantum
mechanics.
• The group U(1)×SU(2)×SU(3) is a Lie group of dimension 1+3+8=12 that is the gauge group of the Standard
Model in particle physics. The dimensions of the factors correspond to the 1 photon + 3 vector bosons + 8 gluons
of the standard model.
• The (3-dimensional) metaplectic group is a double cover of SL2(R) playing an important role in the theory of
modular forms. It is a connected Lie group that cannot be faithfully represented by matrices of finite size, i.e., a
nonlinear group.
• The exceptional Lie groups of types G2, F4, E6, E7, E8 have dimensions 14, 52, 78, 133, and 248. There is also a
group E7½ of dimension 190.
Lie group 34

Constructions
There are several standard ways to form new Lie groups from old ones:
• The product of two Lie groups is a Lie group.
• Any topologically closed subgroup of a Lie group is a Lie group. This is known as Cartan's theorem.
• The quotient of a Lie group by a closed normal subgroup is a Lie group.
• The universal cover of a connected Lie group is a Lie group. For example, the group R is the universal cover of
the circle group S1. In fact any covering of a differentiable manifold is also a differentiable manifold, but by
specifying universal cover, one guarantees a group structure (compatible with its other structures).

Related notions
Some examples of groups that are not Lie groups (except in the trivial sense that any group can be viewed as a
0-dimensional Lie group, with the discrete topology), are:
• Infinite dimensional groups, such as the additive group of an infinite dimensional real vector space. These are not
Lie groups as they are not finite dimensional manifolds
• Some totally disconnected groups, such as the Galois group of an infinite extension of fields, or the additive group
of the p-adic numbers. These are not Lie groups because their underlying spaces are not real manifolds. (Some of
these groups are "p-adic Lie groups"). In general, only topological groups having similar local properties to Rn for
some positive integer n can be Lie groups (of course they must also have a differentiable structure)

Early history
According to the most authoritative source on the early history of Lie groups (Hawkins, p. 1), Sophus Lie himself
considered the winter of 1873–1874 as the birth date of his theory of continuous groups. Hawkins, however,
suggests that it was "Lie's prodigious research activity during the four-year period from the fall of 1869 to the fall of
1873" that led to the theory's creation (ibid). Some of Lie's early ideas were developed in close collaboration with
Felix Klein. Lie met with Klein every day from October 1869 through 1872: in Berlin from the end of October 1869
to the end of February 1870, and in Paris, Göttingen and Erlangen in the subsequent two years (ibid, p. 2). Lie stated
that all of the principal results were obtained by 1884. But during the 1870s all his papers (except the very first note)
were published in Norwegian journals, which impeded recognition of the work throughout the rest of Europe (ibid,
p. 76). In 1884 a young German mathematician, Friedrich Engel, came to work with Lie on a systematic treatise to
expose his theory of continuous groups. From this effort resulted the three-volume Theorie der
Transformationsgruppen, published in 1888, 1890, and 1893.
Lie's ideas did not stand in isolation from the rest of mathematics. In fact, his interest in the geometry of differential
equations was first motivated by the work of Carl Gustav Jacobi, on the theory of partial differential equations of
first order and on the equations of classical mechanics. Much of Jacobi's work was published posthumously in the
1860s, generating enormous interest in France and Germany (Hawkins, p. 43). Lie's idée fixe was to develop a theory
of symmetries of differential equations that would accomplish for them what Évariste Galois had done for algebraic
equations: namely, to classify them in terms of group theory. Lie and other mathematicians showed that the most
important equations for special functions and orthogonal polynomials tend to arise from group theoretical
symmetries. Additional impetus to consider continuous groups came from ideas of Bernhard Riemann, on the
foundations of geometry, and their further development in the hands of Klein. Thus three major themes in 19th
century mathematics were combined by Lie in creating his new theory: the idea of symmetry, as exemplified by
Galois through the algebraic notion of a group; geometric theory and the explicit solutions of differential equations
of mechanics, worked out by Poisson and Jacobi; and the new understanding of geometry that emerged in the works
of Plücker, Möbius, Grassmann and others, and culminated in Riemann's revolutionary vision of the subject.
Lie group 35

Although today Sophus Lie is rightfully recognized as the creator of the theory of continuous groups, a major stride
in the development of their structure theory, which was to have a profound influence on subsequent development of
mathematics, was made by Wilhelm Killing, who in 1888 published the first paper in a series entitled Die
Zusammensetzung der stetigen endlichen Transformationsgruppen (The composition of continuous finite
transformation groups) (Hawkins, p. 100). The work of Killing, later refined and generalized by Élie Cartan, led to
classification of semisimple Lie algebras, Cartan's theory of symmetric spaces, and Hermann Weyl's description of
representations of compact and semisimple Lie groups using highest weights.
Weyl brought the early period of the development of the theory of Lie groups to fruition, for not only did he classify
irreducible representations of semisimple Lie groups and connect the theory of groups with quantum mechanics, but
he also put Lie's theory itself on firmer footing by clearly enunciating the distinction between Lie's infinitesimal
groups (i.e., Lie algebras) and the Lie groups proper, and began investigations of topology of Lie groups (Borel
(2001), ). The theory of Lie groups was systematically reworked in modern mathematical language in a monograph
by Claude Chevalley.

The concept of a Lie group, and possibilities of classification


Lie groups may be thought of as smoothly varying families of symmetries. Examples of symmetries include rotation
about an axis. What must be understood is the nature of 'small' transformations, e.g., rotations through tiny angles,
that link nearby transformations. The mathematical object capturing this structure is called a Lie algebra (Lie himself
called them "infinitesimal groups"). It can be defined because Lie groups are manifolds, so have tangent spaces at
each point.
The Lie algebra of any compact Lie group (very roughly: one for which the symmetries form a bounded set) can be
decomposed as a direct sum of an abelian Lie algebra and some number of simple ones. The structure of an abelian
Lie algebra is mathematically uninteresting (since the Lie bracket is identically zero); the interest is in the simple
summands. Hence the question arises: what are the simple Lie algebras of compact groups? It turns out that they
mostly fall into four infinite families, the "classical Lie algebras" An, Bn, Cn and Dn, which have simple descriptions
in terms of symmetries of Euclidean space. But there are also just five "exceptional Lie algebras" that do not fall into
any of these families. E8 is the largest of these.

Properties
• The diffeomorphism group of a Lie group acts transitively on the Lie group
• Every Lie group is parallelizable, and hence an orientable manifold (there is a bundle isomorphism between its
tangent bundle and the product of itself with the tangent space at the identity)

Types of Lie groups and structure theory


Lie groups are classified according to their algebraic properties (simple, semisimple, solvable, nilpotent, abelian),
their connectedness (connected or simply connected) and their compactness.
• Compact Lie groups are all known: they are finite central quotients of a product of copies of the circle group S1
and simple compact Lie groups (which correspond to connected Dynkin diagrams).
• Any simply connected solvable Lie group is isomorphic to a closed subgroup of the group of invertible upper
triangular matrices of some rank, and any finite dimensional irreducible representation of such a group is 1
dimensional. Solvable groups are too messy to classify except in a few small dimensions.
• Any simply connected nilpotent Lie group is isomorphic to a closed subgroup of the group of invertible upper
triangular matrices with 1's on the diagonal of some rank, and any finite dimensional irreducible representation of
such a group is 1 dimensional. Like solvable groups, nilpotent groups are too messy to classify except in a few
small dimensions.
Lie group 36

• Simple Lie groups are sometimes defined to be those that are simple as abstract groups, and sometimes defined to
be connected Lie groups with a simple Lie algebra. For example, SL2(R) is simple according to the second
definition but not according to the first. They have all been classified (for either definition).
• Semisimple Lie groups are Lie groups whose Lie algebra is a product of simple Lie algebras.[1] They are central
extensions of products of simple Lie groups.
The identity component of any Lie group is an open normal subgroup, and the quotient group is a discrete group.
The universal cover of any connected Lie group is a simply connected Lie group, and conversely any connected Lie
group is a quotient of a simply connected Lie group by a discrete normal subgroup of the center. Any Lie group G
can be decomposed into discrete, simple, and abelian groups in a canonical way as follows. Write
Gcon for the connected component of the identity
Gsol for the largest connected normal solvable subgroup
Gnil for the largest connected normal nilpotent subgroup
so that we have a sequence of normal subgroups
1 ⊆ Gnil ⊆ Gsol ⊆ Gcon ⊆ G.
Then
G/Gcon is discrete
Gcon/Gsol is a central extension of a product of simple connected Lie groups.
Gsol/Gnil is abelian. A connected abelian Lie group is isomorphic to a product of copies of R and the circle
group S1.
Gnil/1 is nilpotent, and therefore its ascending central series has all quotients abelian.
This can be used to reduce some problems about Lie groups (such as finding their unitary representations) to the
same problems for connected simple groups and nilpotent and solvable subgroups of smaller dimension.

The Lie algebra associated with a Lie group


To every Lie group, we can associate a Lie algebra, whose underlying vector space is the tangent space of G at the
identity element, which completely captures the local structure of the group. Informally we can think of elements of
the Lie algebra as elements of the group that are "infinitesimally close" to the identity, and the Lie bracket is
something to do with the commutator of two such infinitesimal elements. Before giving the abstract definition we
give a few examples:
• The Lie algebra of the vector space Rn is just Rn with the Lie bracket given by
[A, B] = 0.
(In general the Lie bracket of a connected Lie group is always 0 if and only if the Lie group is abelian.)
• The Lie algebra of the general linear group GLn(R) of invertible matrices is the vector space Mn(R) of square
matrices with the Lie bracket given by
[A, B] = AB − BA.
If G is a closed subgroup of GLn(R) then the Lie algebra of G can be thought of informally as the matrices m of
Mn(R) such that 1 + εm is in G, where ε is an infinitesimal positive number with ε2 = 0 (of course, no such real
number ε exists). For example, the orthogonal group On(R) consists of matrices A with AAT = 1, so the Lie algebra
consists of the matrices m with (1 + εm)(1 + εm)T = 1, which is equivalent to m + mT = 0 because ε2 = 0.
• Formally, when working over the reals, as here, this is accomplished by considering the limit as ε → 0; but the
"infinitesimal" language generalizes directly to Lie groups over general rings.
The concrete definition given above is easy to work with, but has some minor problems: to use it we first need to
represent a Lie group as a group of matrices, but not all Lie groups can be represented in this way, and it is not
Lie group 37

obvious that the Lie algebra is independent of the representation we use. To get round these problems we give the
general definition of the Lie algebra of any Lie group (in 4 steps):
1. Vector fields on any smooth manifold M can be thought of as derivations X of the ring of smooth functions on the
manifold, and therefore form a Lie algebra under the Lie bracket [X, Y] = XY − YX, because the Lie bracket of any
two derivations is a derivation.
2. If G is any group acting smoothly on the manifold M, then it acts on the vector fields, and the vector space of
vector fields fixed by the group is closed under the Lie bracket and therefore also forms a Lie algebra.
3. We apply this construction to the case when the manifold M is the underlying space of a Lie group G, with G
acting on G = M by left translations Lg(h) = gh. This shows that the space of left invariant vector fields (vector
fields satisfying Lg*Xh = Xgh for every h in G, where Lg* denotes the differential of Lg) on a Lie group is a Lie
algebra under the Lie bracket of vector fields.
4. Any tangent vector at the identity of a Lie group can be extended to a left invariant vector field by left translating
the tangent vector to other points of the manifold. Specifically, the left invariant extension of an element v of the
tangent space at the identity is the vector field defined by v^g = Lg*v. This identifies the tangent space Te at the
identity with the space of left invariant vector fields, and therefore makes the tangent space at the identity into a
Lie algebra, called the Lie algebra of G, usually denoted by a Fraktur Thus the Lie bracket on is given
explicitly by [v, w] = [v^, w^]e.
This Lie algebra is finite-dimensional and it has the same dimension as the manifold G. The Lie algebra of G
determines G up to "local isomorphism", where two Lie groups are called locally isomorphic if they look the same
near the identity element. Problems about Lie groups are often solved by first solving the corresponding problem for
the Lie algebras, and the result for groups then usually follows easily. For example, simple Lie groups are usually
classified by first classifying the corresponding Lie algebras.
We could also define a Lie algebra structure on Te using right invariant vector fields instead of left invariant vector
fields. This leads to the same Lie algebra, because the inverse map on G can be used to identify left invariant vector
fields with right invariant vector fields, and acts as −1 on the tangent space Te.
The Lie algebra structure on Te can also be described as follows: the commutator operation
(x, y) → xyx−1y−1
on G × G sends (e, e) to e, so its derivative yields a bilinear operation on TeG. This bilinear operation is actually the
zero map, but the second derivative, under the proper identification of tangent spaces, yields an operation that
satisfies the axioms of a Lie bracket, and it is equal to twice the one defined through left-invariant vector fields.

Homomorphisms and isomorphisms


If G and H are Lie groups, then a Lie-group homomorphism f : G → H is a smooth group homomorphism. (It is
equivalent to require only that f be continuous rather than smooth.) The composition of two such homomorphisms is
again a homomorphism, and the class of all Lie groups, together with these morphisms, forms a category. Two Lie
groups are called isomorphic if there exists a bijective homomorphism between them whose inverse is also a
homomorphism. Isomorphic Lie groups are essentially the same; they only differ in the notation for their elements.
Every homomorphism f : G → H of Lie groups induces a homomorphism between the corresponding Lie algebras
and . The association G is a functor (mapping between categories satisfying certain axioms).
One version of Ado's theorem is that every finite dimensional Lie algebra is isomorphic to a matrix Lie algebra. For
every finite dimensional matrix Lie algebra, there is a linear group (matrix Lie group) with this algebra as its Lie
algebra. So every abstract Lie algebra is the Lie algebra of some (linear) Lie group.
The global structure of a Lie group is not determined by its Lie algebra; for example, if Z is any discrete subgroup of
the center of G then G and G/Z have the same Lie algebra (see the table of Lie groups for examples). A connected
Lie group is simple, semisimple, solvable, nilpotent, or abelian if and only if its Lie algebra has the corresponding
Lie group 38

property.
If we require that the Lie group be simply connected, then the global structure is determined by its Lie algebra: for
every finite dimensional Lie algebra over F there is a simply connected Lie group G with as Lie algebra, unique
up to isomorphism. Moreover every homomorphism between Lie algebras lifts to a unique homomorphism between
the corresponding simply connected Lie groups.

The exponential map


The exponential map from the Lie algebra Mn(R) of the general linear group GLn(R) to GLn(R) is defined by the
usual power series:

for matrices A. If G is any subgroup of GLn(R), then the exponential map takes the Lie algebra of G into G, so we
have an exponential map for all matrix groups.
The definition above is easy to use, but it is not defined for Lie groups that are not matrix groups, and it is not clear
that the exponential map of a Lie group does not depend on its representation as a matrix group. We can solve both
problems using a more abstract definition of the exponential map that works for all Lie groups, as follows.
Every vector v in determines a linear map from R to taking 1 to v, which can be thought of as a Lie algebra
homomorphism. Because R is the Lie algebra of the simply connected Lie group R, this induces a Lie group
homomorphism c : R → G so that

for all s and t. The operation on the right hand side is the group multiplication in G. The formal similarity of this
formula with the one valid for the exponential function justifies the definition

This is called the exponential map, and it maps the Lie algebra into the Lie group G. It provides a diffeomorphism
between a neighborhood of 0 in and a neighborhood of e in G. This exponential map is a generalization of the
exponential function for real numbers (because R is the Lie algebra of the Lie group of positive real numbers with
multiplication), for complex numbers (because C is the Lie algebra of the Lie group of non-zero complex numbers
with multiplication) and for matrices (because Mn(R) with the regular commutator is the Lie algebra of the Lie group
GLn(R) of all invertible matrices).
Because the exponential map is surjective on some neighbourhood N of e, it is common to call elements of the Lie
algebra infinitesimal generators of the group G. The subgroup of G generated by N is the identity component of G.
The exponential map and the Lie algebra determine the local group structure of every connected Lie group, because
of the Baker–Campbell–Hausdorff formula: there exists a neighborhood U of the zero element of , such that for u,
v in U we have
exp(u) exp(v) = exp(u + v + 1/2 [u, v] + 1/12 [[u, v], v] − 1/12 [[u, v], u] − ...)
where the omitted terms are known and involve Lie brackets of four or more elements. In case u and v commute, this
formula reduces to the familiar exponential law exp(u) exp(v) = exp(u + v).
The exponential map from the Lie algebra to the Lie group is not always onto, even if the group is connected (though
it does map onto the Lie group for connected groups that are either compact or nilpotent). For example, the
exponential map of SL2(R) is not surjective. Also, exponential map is not surjective nor injective for
infinite-dimensional (see below) Lie groups modelled on C∞ Fréchet space, even from arbitrary small neighborhood
of 0 to corresponding neighborhood of 1.
Lie group 39

Infinite dimensional Lie groups


Lie groups are often defined to be finite dimensional, but there are many groups that resemble Lie groups, except for
being infinite dimensional. The simplest way to define infinite dimensional Lie groups is to model them on Banach
spaces, and in this case much of the basic theory is similar to that of finite dimensional lie groups. However this is
inadequate for many applications, because many natural examples of infinite dimensional Lie groups are not Banach
manifolds. Instead one needs to define Lie groups modeled on more general locally convex topological vector
spaces. In this case the relation between the Lie algebra and the Lie group becomes rather subtle, and several results
about finite dimensional Lie groups no longer hold.
Some of the examples that have been studied include:
• The group of diffeomorphisms of a manifold. Quite a lot is known about the group of diffeomorphisms of the
circle. Its Lie algebra is (more or less) the Witt algebra, which has a central extension called the Virasoro algebra,
used in string theory and conformal field theory. Very little is known about the diffeomorphism groups of
manifolds of larger dimension. The diffeomorphism group of spacetime sometimes appears in attempts to
quantize gravity.
• The group of smooth maps from a manifold to a finite dimensional Lie group is an example of a gauge group
(with operation of pointwise multiplication), and is used in quantum field theory and Donaldson theory. If the
manifold is a circle these are called loop groups, and have central extensions whose Lie algebras are (more or
less) Kac–Moody algebras.
• There are infinite dimensional analogues of general linear groups, orthogonal groups, and so on. One important
aspect is that these may have simpler topological properties: see for example Kuiper's theorem.

Notes
[1] Helgason, Sigurdur (1978). Differential Geometry, Lie Groups, and Symmetric Spaces. New York: Academic Press. p. 131.
ISBN 0123384605.

References
• Adams, John Frank (1969), Lectures on Lie Groups, Chicago Lectures in Mathematics, Chicago: Univ. of
Chicago Press, ISBN 0-226-00527-5.
• Borel, Armand (2001), Essays in the history of Lie groups and algebraic groups (http://books.google.com/
books?isbn=0821802887), History of Mathematics, 21, Providence, R.I.: American Mathematical Society,
ISBN 978-0-8218-0288-5, MR1847105
• Bourbaki, Nicolas, Elements of mathematics: Lie groups and Lie algebras. Chapters 1–3 ISBN 3-540-64242-0,
Chapters 4–6 ISBN 3-540-42650-7, Chapters 7–9 ISBN 3-540-43405-4
• Chevalley, Claude (1946), Theory of Lie groups, Princeton: Princeton University Press, ISBN 0-691-04990-4.
• Fulton, William; Harris, Joe (1991), Representation theory. A first course, Graduate Texts in Mathematics,
Readings in Mathematics, 129, New York: Springer-Verlag, ISBN 978-0-387-97495-8, MR1153249, ISBN
978-0-387-97527-6
• Hall, Brian C. (2003), Lie Groups, Lie Algebras, and Representations: An Elementary Introduction, Springer,
ISBN 0-387-40122-9.
• Hawkins, Thomas (2000), Emergence of the theory of Lie groups (http://books.google.com/
books?isbn=978-0-387-98963-1), Sources and Studies in the History of Mathematics and Physical Sciences,
Berlin, New York: Springer-Verlag, ISBN 978-0-387-98963-1, MR1771134 Borel's review (http://www.jstor.
org/stable/2695575)
• Knapp, Anthony W. (2002), Lie Groups Beyond an Introduction, Progress in Mathematics, 140 (2nd ed.), Boston:
Birkhäuser, ISBN 0-8176-4259-5.
Lie group 40

• Rossmann, Wulf (2001), Lie Groups: An Introduction Through Linear Groups, Oxford Graduate Texts in
Mathematics, Oxford University Press, ISBN 978-0198596837. The 2003 reprint corrects several typographical
mistakes.
• Serre, Jean-Pierre (1965), Lie Algebras and Lie Groups: 1964 Lectures given at Harvard University, Lecture
notes in mathematics, 1500, Springer, ISBN 3-540-55008-9.
• Steeb, Willi-Hans (2007), Continuous Symmetries, Lie algebras, Differential Equations and Computer Algebra:
second edition, World Scientific Publishing, ISBN 981-270-809-X.
Article Sources and Contributors 41

Article Sources and Contributors


Group (mathematics)  Source: http://en.wikipedia.org/w/index.php?oldid=429706344  Contributors: 0, 132.146.192.xxx, A legend, AGK, APH, AYSH AYSH AY AY AY AY, Alecobbe,
Algebraist, Alksentrs, Allanhalme, Amahoney, Andres, Anonymous Dissident, Antzervos, Archelon, Arneth, Art LaPella, Arthur Rubin, Ashwin, Awadewit, AxelBoldt, Basploeger, Bcrowell,
Bender235, Bh3u4m, BigDunc, Bobrek, Braveorca, Brews ohare, Brion VIBBER, Brusegadi, Bryan Derksen, C S, CBM, CBM2, CRGreathouse, Calliopejen1, Cenarium, Charles Matthews,
Charvest, Chas zzz brown, Chenxlee, Chinju, Cj67, Classicalecon, Conversion script, CàlculIntegral, DAGwyn, DMacks, DYLAN LENNON, Dan Hoey, Danielh, Dank, Daran, Darius Bacon,
David Eppstein, David Underdown, Dcoetzee, Debivort, Delaszk, Dispenser, Donarreiskoffer, DoubleBlue, Dr.enh, Drschawrz, Dtrebbien, Dungodung, Dysprosia, ELApro, EdJohnston,
Eeekster, EnumaElish, Epbr123, Eriatarka, Ettrig, Eubulides, Farvin111, Fbaggins, Felizdenovo, Fibonacci, Frankenpuppy, GabeAB, Gaius Cornelius, Gandalf61, Geometry guy, Giftlite, Giggy,
Gilliam, Ginpasu, Git2010, Glenn, Goochelaar, Graham87, Grubber, Guy Harris, Hairy Dude, Hans Adler, Harry007754, Harryboyles, Hayabusa future, Headbomb, Helix84, HenryLi,
Hippophaë, IPonomarev, Ilya, Indeed123, JackSchmidt, Jakob.scholbach, Jarmiz, Jarretinha, JasonASmith, Jimfbleak, Jitse Niesen, Joanjoc, Jonik, Jorgen W, Josh Parris, Jrtayloriv, Jshadias,
Jujutacular, Kakila, Kareekacha, KarlHeg, Kinser, Kku, Konradek, Kozuch, Lambiam, LarryLACa, Lee J Haywood, Lesnail, Ling.Nut, LittleDan, LkNsngth, Loadmaster, LokiClock, M759,
Mani1, MarSch, Marc van Leeuwen, Marco Krohn, Mark Wolfe, Markan, Markus Poessel, Mav, Mee26, Melchoir, Michael Devore, Michael Hardy, Michael Kinyon, Mike Christie,
MithrandirMage, Mongreilf, Mormegil, Mostlyharmless, Mpatel, Mscalculus, Msh210, Myasuda, Natkeeran, Nbarth, Nergaal, Newone, Nibuod, Nihiltres, Nsk92, Nwbeeson, OdedSchramm,
OktayD, Oleg Alexandrov, Oli Filth, Olivier, OneWeirdDude, Ozob, PAR, Pako, Patrick, Paul August, Paul Matthews, Paxinum, Pbroks13, Pcap, Pdenapo, Peiresc, Penumbra2000, Philip
Trueman, Piledhigheranddeeper, Pmanderson, Point-set topologist, Pomte, Puckly, Python eggs, Qwfp, R.e.b., RJHall, Randomblue, Raul654, Rawling, Redgolpe, Revolver, Ricardo sandoval,
Rich Farmbrough, Rick Norwood, Rjwilmsi, RobHar, Robin S, Rursus, Rush Psi, Ryan Reich, S2000magician, SP-KP, Salix alba, SandyGeorgia, Sekky, Shahab, Silly rabbit, Skippydo, Sl,
Slawekk, Soru81, Spindled, Stephen Ducret, Stifle, Swpb, TKD, Tango, Tarquin, Taxman, Template namespace initialisation script, Teresol, TheOtherJesse, Thehotelambush, Thingg,
TimothyRias, Tobias Bergemann, Toby, Toby Bartels, Tosha, Treisijs, Uriyan, VKokielov, Vanished User 0001, Wayne Slam, WikiDan61, Wikitanvir, WillowW, WinoWeritas, WolfmanSF,
Wood Thrush, Woodstone, Wshun, X Fallout X, X Pacman X, XJamRastafire, Xantharius, Xaos, Zippy, Zundark, Zzuuzz, 274 anonymous edits

Orthogonal group  Source: http://en.wikipedia.org/w/index.php?oldid=429952786  Contributors: Algebraist, AxelBoldt, BenFrantzDale, Chadernook, Charles Matthews, DreamingInRed,
Drschawrz, Ettrig, Fropuff, Gauge, Giftlite, Guardian of Light, HappyCamper, JackSchmidt, Jim.belk, KnightRider, Kwamikagami, Lambiam, Looxix, Loren Rosen, Masnevets, MathMartin,
Mathiscool, Michael Hardy, Monsterman222, Msh210, Nbarth, Niout, Noideta, Oleg Alexandrov, Patrick, Paul D. Anderson, Phys, Pt, R.e.b., Renamed user 1, Salix alba, Schmloof, Shell
Kinney, Softcafe, Somethingcompletelydifferent, Technohead1980, The Anome, Thehotelambush, TooMuchMath, Ulner, Unyoyega, Weialawaga, Wolfrock, Zundark, 45 anonymous edits

Lie group  Source: http://en.wikipedia.org/w/index.php?oldid=428324516  Contributors: 212.29.241.xxx, Abdull, Akriasas, Alex Varghese, AnmaFinotera, Anterior1, Arcfrk, Archelon,
Arkapravo, AxelBoldt, BMF81, Badger014, Barak, Bears16, Beastinwith, Beland, BenFrantzDale, Bender235, Benjamin.friedrich, Bobblewik, Bongwarrior, Borat fan, Brian Huffman, Buster79,
CBM, CRGreathouse, Cacadril, Charles Matthews, Cherlin, ChrisJ, Cmelby, Conversion script, Dablaze, Darkfight, Davewild, David Eppstein, David Shay, Deciwill, DefLog, Deflective,
Dorftrottel, Dr.enh, Drorata, Dysprosia, Dzordzm, Ekeb, Eubulides, FlashSheridan, Fropuff, GTBacchus, Genuine0legend, Geometry guy, Giftlite, Graham87, HappyCamper, Headbomb,
Hesam7, Hillman, Homeworlds, Ht686rg90, Incnis Mrsi, Inquisitus, Isnow, Itai, JDspeeder1, JackSchmidt, James.r.a.gray, JamesMLane, Jason Quinn, Jesper Carlstrom, Jim.belk, Jitse Niesen,
Joriki, Josh Cherry, Josh Grosse, JustAGal, KSmrq, Kaoru Itou, KbReZiE 12, KeithB, Kier07, Krasnoludek, Kwamikagami, Len Raymond, Leontios, Lethe, Linas, Lockeownzj00, Logical2u,
Looxix, Lseixas, MarSch, Marc van Leeuwen, Masnevets, Mathchem271828, Mhss, Michael Hardy, Michael Kinyon, Michael Slone, Miguel, MotherFunctor, Msh210, MuDavid, Myasuda,
NatusRoma, Ndbrian1, Ninte, Niout, Oleg Alexandrov, Orthografer, Oscarbaltazar, Ozob, PAR, Paul August, Phys, Pidara, Pmanderson, Porcher, Pred, R.e.b., Rgdboer, RichardVeryard, RobHar,
Rocket71048576, RodVance, S2000magician, Saaska, Salgueiro, Salix alba, Shanes, Sidiropo, Silly rabbit, Siva1979, Smaines, Smylei, Stevertigo, Sullivan.t.j, Suslindisambiguator, Sławomir
Biały, Tanath, Tide rolls, Tobias Bergemann, Tom Lougheed, Tompw, TomyDuby, Topology Expert, Tosha, Trevorgoodchild, Ulner, Unifey, VKokielov, Weialawaga, Wgmccallum,
WhatamIdoing, XJamRastafire, Xantharius, Xavic69, Yggdrasil014, Zoicon5, Zundark, 114 anonymous edits
Image Sources, Licenses and Contributors 42

Image Sources, Licenses and Contributors


Image:Rubik's cube.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Rubik's_cube.svg  License: GNU Free Documentation License  Contributors: User:Booyabazooka
Image:group D8 id.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_id.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 90.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_90.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 180.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_180.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 270.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_270.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 fv.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_fv.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 fh.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_fh.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 f13.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_f13.svg  License: Public Domain  Contributors: User:TimothyRias
Image:group D8 f24.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Group_D8_f24.svg  License: Public Domain  Contributors: User:TimothyRias
Image:Wallpaper group-cm-6.jpg  Source: http://en.wikipedia.org/w/index.php?title=File:Wallpaper_group-cm-6.jpg  License: Public Domain  Contributors: Unknown
Image:Fundamental group.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Fundamental_group.svg  License: Creative Commons Attribution-Sharealike 3.0  Contributors:
User:Pbroks13
Image:Cyclic group.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Cyclic_group.svg  License: Creative Commons Attribution-Sharealike 3.0  Contributors: User:Jakob.scholbach,
User:Pbroks13
Image:Sixteenth stellation of icosahedron.png  Source: http://en.wikipedia.org/w/index.php?title=File:Sixteenth_stellation_of_icosahedron.png  License: Attribution  Contributors: User:Jim2k
Image:C60a.png  Source: http://en.wikipedia.org/w/index.php?title=File:C60a.png  License: GNU Free Documentation License  Contributors: Original uploader was Mstroeck at en.wikipedia
Later versions were uploaded by Bryn C at en.wikipedia.
Image:Ammonia-3D-balls-A.png  Source: http://en.wikipedia.org/w/index.php?title=File:Ammonia-3D-balls-A.png  License: Public Domain  Contributors: Ben Mills
Image:Cubane-3D-balls.png  Source: http://en.wikipedia.org/w/index.php?title=File:Cubane-3D-balls.png  License: Public Domain  Contributors: Ben Mills
Image:Hexaaquacopper(II)-3D-balls.png  Source: http://en.wikipedia.org/w/index.php?title=File:Hexaaquacopper(II)-3D-balls.png  License: Public Domain  Contributors: Ben Mills
Image:Uniform tiling 73-t2 colored.png  Source: http://en.wikipedia.org/w/index.php?title=File:Uniform_tiling_73-t2_colored.png  License: Creative Commons Attribution-Sharealike
3.0,2.5,2.0,1.0  Contributors: Jakob.scholbach (talk) Original uploader was Jakob.scholbach at en.wikipedia
Image:Matrix multiplication.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Matrix_multiplication.svg  License: Creative Commons Attribution-Sharealike 3.0  Contributors:
User:Pbroks13
Image:Circle as Lie group2.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Circle_as_Lie_group2.svg  License: Public Domain  Contributors: User:Oleg Alexandrov
File:Loudspeaker.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Loudspeaker.svg  License: Public Domain  Contributors: Bayo, Gmaxwell, Husky, Iamunknown, Myself488,
Nethac DIU, Omegatron, Rocket000, The Evil IP address, Wouterhagens, 9 anonymous edits
Image:Circle as Lie group.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Circle_as_Lie_group.svg  License: Public Domain  Contributors: User:Oleg Alexandrov
License 43

License
Creative Commons Attribution-Share Alike 3.0 Unported
http:/ / creativecommons. org/ licenses/ by-sa/ 3. 0/

Você também pode gostar