








































































Study with the several resources on Docsity
Earn points by helping other students or get them with a premium plan
Prepare for your exams
Study with the several resources on Docsity
Earn points to download
Earn points by helping other students or get them with a premium plan
Community
Ask the community for help and clear up your study doubts
Discover the best universities in your country according to Docsity users
Free resources
Download our free guides on studying techniques, anxiety management strategies, and thesis advice from Docsity tutors
algebra group theory dummit foote, david dummit, richard foote, group theory extracted
Typology: Summaries
1 / 80
This page cannot be seen from the preview
Don't miss anything!
Let S be a set. A mapping
sxs.:....s
is sometimes called a law of composition (of S into itself). If x, y are elements of S, the image of the pair (x, y) under this mapping is also called their product under the law of composition, and will be denoted by xy. (Sometimes, we also write x ·^ y, and in many cases it is also convenient to use an additive notation, and thus to write x +^ y. In that case, we call this element the^ sum^ of x and y. It is customary to use the notation x + y only when the relation x + y = y +^ x holds.) Let S be a set with a law of composition. If x, y, z are elements of S, then we may form their product in two ways : (xy)z and x(yz). If (xy)z = x(yz) for all x, y, z in S then we say that the law of composition is a ssociative. An element e of S such that ex = x = xe for all x e S is called a unit element. (When the law of composition is written additively, the unit element is denoted by 0, and is called a zero element.) A unit element is unique, for if e'^ is another unit element, we have
e = ee^ I = eI
by assumption. In most cases, the unit element is written simply 1 (instead of e). For most of this chapter, however, we shall write e so as to avoid confusion in proving the most basic properties. A monoid is a set G, with a law of composition which is associative, and having a unit element (so that in particular, G is not empty).
3
4 G ROUPS^ I,^ §
We define their product inductively : n
v= 1
m n^ m+n nxJl· nxm+v= nxv, 11= 1 v= 1 v= 1
leave it as an exercise. One also writes
and we define
m+n n
m+ 1 v= 1
0
v= 1 As a matter of convention, we agree also that the empty product is equal to the unit element. It would be possible to define more general laws of composition, i.e. maps
commutativity in any setting for which they make sense. For instance, for commutativity we need a law of composition
where the two sets of departure are the same. Commutativity then means
associativity, we leave it to the reader to formulate the most general combination of sets under which it will work. We shall meet special cases later, for instance arising from maps
above we have (^) (xy)z (^) = x(yz). If the law of composition of G is commutative, we also say that G is com mutative (or abelian).
6 G ROUPS I,^ §
n r n f(i,j)] =^ n^ [n J(i,j)]. i e l Le J (^) j e J^ i e l We leave the proof as an exercise.
to be n n 1 X,
If S, S' are two subsets of a monoid G, t hen we define SS' to be the subset consisting of all elements xy, with xES and^ yES'.^ Inductively, we can define the product of a finite number of subsets, and we have associativity. For in
(because G^ has a unit element).^ If^ xE G, then we define xS to be^ {x }S, where {x} is the set consisting of the single element (^) x. Thus xS consists of all elements xy,^ with^ yES. By a submonoid of G , we shall mean a subset H of G containing the unit element e, and such that, if x, y E H then xy E H (we say that H is closed under the law of composition). It is then clear that H is itself a monoid , under the law of composition induced by that of G.
is a submonoid of G. The set of integers > 0 under addition is a monoid. Later we shall define rings. If R is a commutative ring, we shall deal with multiplicative subsets S, that is subsets containing the unit element, and such that if (^) x, yES then (^) xy E S. Such subsets are monoids. A routine example. Let N be the natural numbers , i. e. the integers > 0. Then N is an additive monoid. In some applications , it is useful to deal with a multiplicative version. See the definition of polynomials in Chapter II , §3 , where a higher-dimensional version is also used for polynomial s in several variables. An interesting example. We assume that the reader is familiar with the terminology of elementary topology. Let M be the set of homeomorphism classes of compact (connected) surfaces. We shall define an addition in M. Let S, S' be compact surfaces. Let D be a small disc in S, and D' a small disc in S'. Let C, C' be the circles which form the boundaries of D and D' respectively. Let D0 , D� be the interiors of D and D' respectively, and glue S-D0 to S'-D0 by identifying C with C'. It can be shown that the resulting surface is independent,
I, §2 G ROUPS^7
up to homeomorphism, of the various choices made in the preceding construc tion. If u, u'^ denote the homeomorphism classes of S and S' respectively, we define u +^ u'^ to be the class of the surface obtained by the preceding gluing process. It can be shown that this addition defines a monoid structure on M,
denotes the class of the torus, and 1t denotes the class of the projective plane, then every element u of M has a unique expression of the form
(The reasons for inserting the preceding example are twofold : First to relieve the essential dullness of the section. Second to show the reader that monoids exist in nature. Needless to say, the example will not be used in any way throughout the rest of the book.)
isomorphism classes of modules over a ring form a monoid under the direct sum. In Chapter XV , § 1 , we shall consider a monoid consisting of equivalence classes of quadratic forms.
A group G is a monoid, such that for every element x E G there exists an
We denote this inverse by x-^1 (or by -x when the law of composition is written additively).
exponentiation hold for all integers, not only for integers > 0 (^) (as we pointed out for mono ids in § 1 ). The trivial proofs are left to the reader. In the definitions of unit elements and in verses, we could also define left units and left inverses (in the obvious way). One can easily prove that these are also units and inverses respectively under suitable conditions. Namely:
Multiplying on the left by a left inverse for b yields
I, §2 G ROUPS 9
with r E Z and r prime to n. A generator for this group is called a primitive n-th root of unity. Example. The direct product. Let G I , G2 be groups. Let G I X G2 be the direct product as sets , so G I x G2 is the set of all pairs (xi , x2 ) with X; E G;. We define the product componentwise by
(xi , x2)( yi , Y2)^ =^ (xi YI , X2 Y2 ). Then G 1 X G2 is a group , whose unit element is ( e 1 , e2) (where e; is the unit element of G; ). Similarly , for n groups we define G I X • · · X G (^) n to be the set of n-tuples with X; E G; (i^ = 1 ,... , n) , and componentwise multiplication. Even more generally , let I be a set , and for each i E I, let G; be a group. Let G = fi^ G; be the set-theoretic product of the sets G;. Then G is the set of all families (x; );E1 with X; E G;. We can define a group structure on G by compo nentwise multiplication , namely , if (x; );EJ and ( Y; );EJ^ are^ two^ elements of^ G,^ we define their product to be ( X;Y; );EJ · We define the inverse of (x; );EJ to be (xj^1 );EJ· It is then obvious that G is a group called the direct product of the family.
Let G be a group. A subgroup H of G is a subset of G containing the unit
trivial if it consists of the unit element �lone. The intersection of an arbitrary non-empty family of subgroups is a subgroup (trivial verification). Let G be a group and S a subset of G. We shall say that S generates G, or that S is a set of generators for G, if every element of G can be expressed as a product of elements of S or inverses of elements of S, i.e. as a product x 1 • • • xn where each xi or xi- 1^ is in S. It is clear that the set of all such products is a subgroup of G (the empty product is t he unit element), and is the smallest sub group of G containing S. Thus S generates G if and only if the smallest subgroup of G containing S is G itself. If G is generated by S, then we write G = (S). By definition , a cyclic group is a group which has one generator. Given elements xi ,... , Xn E G , these elements generate a subgroup (xb... , xn), namely the set of all elements of G of the form
X�l · · · x�; with k1 , • • • , krE Z.
A single element x E G generates a cyclic subgroup.
Example. There are two non-abelian groups of order 8. One is the group of symmetries of the square , generated by two elements u, T such that
u4 = T^2 = e and TUT- I^ = u3.
The other is the quaternion group , generated by two elements , i, j such that if we put k = ij and m = i^2 , then lJ^.^. = mJl ..^.
After you know enough facts about groups, you can easily do Exercise 35.
1 0 GROUPS I, §
Let G, G' be monoids. A monoid-homomorphism (or simply (^) homomorphism)
and mapping the unit element of G into that of G'. If G, G' are groups, a group homomorphism of G into G' is simply a monoid-homomorphism. We sometimes say : "Let f : G � G' be a group-homomorphism " to mean : "Let G, G' be groups, and let fbe a homomorphism from G into G'." Letf: G � G' be a group-homomorphism. Then f(x - 1 ) = f(x) - 1
Furthermore, if G, G' are groups and f: G -+ G' is a map such that f(xy) = f(x)f(y)
Let G, G' be monoids. A homomorphism! : G -+ G' is called an isomorphism
identity mappings (in G' and G respectively). It is trivially verified that f is
between two groups G and G' is sometimes denoted by G � G'. If G = G', we say that isomorphism is an automorphism. A homomorphism of G into itself is also called an endomorphism.
Example. Let G be a monoid and x an element of G. Let N denote the (additive) monoid of integers 2 0. Then the mapf: N -+ G such thatf(n) = xn is a homomorphism. If G is a group, we can extend fto a homomorphism of Z
are left to the reader.
easily that the map
from G into itself is a homomorphism. So is the map x r--+ x-^1. The map x r--+ xn is called the n-th power map. Example. Let I = { i} be an indexing set, and let { G;} be a family of groups. Let G = fi G;^ be their direct product. Let p;: G � G;
be the projection on the i-th factor. Then P; is a homomorphism.
1 2 G ROU PS I, §
We observe that Proposition 2. 1 generalizes by induction to a finite number
and such that
In that case, G is isomorphic to the direct product
Hence any two left cosets have the same cardinality.
From the above conclusion, we get :
Hence
j
(disjoint),
(disjoint).
cosets. Suppose
I , §
Y·X· KJ I = Y··X·· KJ I
NORMAL SUBGROUPS 1 3
for a pair of indices {j, i) and (j', i'). Multiplying by H on the right, and noting that X;, xi' are in H, we get Y·HJ = y (^) J.. H' whence Yi = Yr · From this it follows that xi K = X;· K and therefore that X; = X;·, as was to be shown.
The formula of Proposition 2. 2 may therefore be generalized by writing
(G : K) = (G : H)(H : K) , with the understanding that if two of the three indices appearing in this formula are finite , then so is the third and the formula holds. The above results are concerned systematically with left cosets. For the right cosets , see Exercise 1 0. Example. A group of prime order is cyclic. Indeed , let G have order p and let a E G, a =I= e. Let H be the subgroup generated by a. Then #(H) divides p and is =I= 1 , so #(H) = p and so H = G , which is therefore cyclic. Example. Let Jn = { 1 ,... , n }. Let Sn be the group of permutations of ln. We define a transposition to be a permutation T such that there exist two elements r =I= s in Jn for which T(r)^ =^ s, T(s) =^ r, and T(k) =^ k for all k =I= r, s. Note that the transpositions generate Sn. Indeed , say u is a permutation , u (n) = k =I= n. Let T be the transposition interchanging k, n. Then TU leaves n fixed , and by induction , we can write TU as a product of transpositions in Perm(] n- I)' thus proving that transpositions generate s n. Next we note that #(Sn) = n!. Indeed, let H be the subgroup of Sn consisting of those elements which leave n fixed. Then H may be identified with Sn-t· If U; (i^ =^ 1 ,^...^ , n)^ is an element of^ Sn^ SUCh that^ U;(n)^ =^ i,^ then it is immediately verified that u1, • • • , un are coset representatives of H. Hence by induction (Sn : 1 ) = n(H : 1 ) =^ n!. Observe that for u; we could have taken the transposition T;, which interchanges i (^) and n (except for i (^) = n, where we could take un to be the identity).
We have already observed that the kernel of a group-homomorphism is a subgroup. We now wish to characterize such subgroups. Letf : G __.. G' be a group-homomorphism, and let H be its kernel. If x is an
I, §3 NORMAL SUBGROU PS^ 1 5
Second , let G be the set of all maps Ta,b : R � R such that
morphism of G onto the multiplicative group , whose kernel is the group of
Let H be a subgroup of G. Then H is obviously a normal subgroup of its
x = y (mod H)
if x and y lie in the same coset of H, or equivalently if xy -^1 (or y - 1x) lie in H. We read this relation " x and y are congruent modulo H." When G is an additive group, then x = 0 (mod H) means that x lies in H, and x = y (mod H) means that x - y (or y - x) lies in H. This notation of congruence is used mostly for additive groups. Let
be a sequence of homomorphisms. We shall say that this sequence is exact if
H ..i.. G � G/H
morphisms having more than one term, like
is called exact if it is exact at each joint, i.e. if.
1 6 G ROU PS I, §
is exact means that f is injective, that lm f = Ker g, and that g is surjective. If
More precisely, there exists a commutative diagram
in which the vertical maps are isomorphisms, and the rows are exact. Next we describe some homomorphisms, all of which are called canonical. (i) Let G, G' be groups and f: G __.. G' a homomorphism whose kernel
representative x, and it is then trivially verified that f* is a homomorphism, is injective, and is the unique homomorphism satisfying our requirements. We shall say that f* is induced by f Our homomorphism !* induces an iso111orphism
sion of homomorphisms :
diagram commutative :
G/N As before, cp is the canonical map. We can define f* as in ( 1) by the rule
This is well defined, and is trivially verified to satisfy all our requirements.
1 8 GROU PS I,^ §
again called canonical, giving rise to the commutative diagram
We shall now describe some applications of our homomorphism statements.
form an abelian tower (resp. cyclic tower), because we have an injective homo morphism
abelian ( resp. cyclic). A refinement of a tower G = G0 ::J G 1 ::J •^ •^ • ::J Gm is a tower which can be obtained by inserting a finite number of subgroups in the given tower. A group is said to be solvable if it has an abelian tower, whose
we obtain the desired cyclic tower.
Example. In Theorem 6.4 it will be proved that a group whose order is a prime power is solvable.
I , §3 NORMAL SUBG ROU PS 1 9
Example. One of the major results of group theory is the Fe it-Thompson theorem that all finite groups of odd order are solvable. Cf. [Go 68]. Example. Solvable groups will occur in field theory as the Galois groups of solvable extensions. See Chapter VI , Theorem 7. 2.
Example. We assume the reader knows the basic notions of linear algebra. Let k^ be a field. Let G^ =^ GL(n^ ,^ k)^ be the group of invertible^ n^ x^ n^ matrices in k. Let T = T(n, k) be the upper triangular group; that is , the subgroup of matrices which are 0 below the diagonal. Let D be the diagonal group of diagonal matrices with non-zero components on the diagonal. Let N be the additive group of matrices which are 0 on and below the diagonal , and let U = I + N, where I is the unit n x n matrix. Then U is a subgroup of G. (Note that N consists of nilpotent matrices , i. e. matrices A such that Am = 0 for some positive integer m. Then (I - A) - I^ = I + A + A2 +... + Am - I is computed using the geometric series. ) Given a matrix A E T, let diag(A) be the diagonal matrix which has the same diagonal components as A. Then the reader will verify that we get a surjective homomorphism (^) T � D given by A � diag(A).
The kernel of this homomorphism is precisely U. More generally , observe that for r > 2 , the set Nr- I consists of all matrices of the form 0 0 0 a (^) Ir.....^ a^ In 0 0 0 0 a (^2) ,r+ 1 a 2 n
M - 0 0................^ an -r+ l ,n 0 0................^0
0 0................^0
Let Ur = I +^ N r. Then U1 U and Ur :J Ur + 1 • Furthermore, Ur + 1 is normal
the mapping which sends I + M to the n - r-tuple (a lr+ l '... , an -r,n ) E kn- r. This n - r-tuple could be called the r-th upper diagonal. Thus we obtain an abelian tower
Theorem 3.2.^ Let^ G^ be a group and^ H^ a normal subgroup. Then^ G^ is solvable if and only if H and G /H are solvable.
Proof. We prove that G solvable implies that H is solvable. Let G = G0 :J G I :J... :J Gr = {e} be a tower of groups with G; + I normal in G; and such that G; /G;+ I is abelian. Let H; = H n G;. Then H;+ I is normal in H;, and we have an embedding H;/H;+ t (^) � G;/G;+ (^) I , whence H; /H;+ t is abelian , whence proving that H is solvable. We leave the proofs of the other statements to the reader.
I, §
u( U n V )
u
u
u n v
NORMAL SUBGROUPS 21
v
( U n V ) v
v
u n v
correspond to certain groups which can be determined as follows. The inter section of two line segments going downwards represents the intersection of groups. Two lines going upwards meet in a point which represents the product of two subgroups (i.e. the smallest subgroup containing both of them). We consider the two parallelograms representing the wings of the butterfly , and we shall give isomoqJhisms of the factor groups as follows:
----^ u(^ u n^ V)^ ,....... ------u n^ v^ = (....;_^ u n^ V)v u(U n v) (u n V)( U n v) (u n V)v.
morphism
This is obtained from the isomorphism theorem
by setting H = U (^) n V and N = u( U n v). This gives us the isomorphism on the left. By symmetry we obtain the corresponding isomorphism on the right , which proves the Butterfly lemma.
be normal towers of subgroups, ending with the trivial group. We shall say that these towers are equivalent if r = s and if there exists a permutation of the
22 G ROU PS
I , §
In other words, the sequences of factor groups in our two towers are the same, up to isomorphisms, and a permutation of the indices.
ending with the trivial group have equivalent refinements.
Similarly, we define
have isomorphisms
Thus the sequence of non-trivial factors for the original tower, or the refined tower, is the same. This proves our theorem.