Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
Алгебра_1 / Gilbert - Modern algebra with applications.pdf
Скачиваний:
34
Добавлен:
23.02.2015
Размер:
2.55 Mб
Скачать

48 3 GROUPS

GROUPS AND SYMMETRIES

A group (G, ·) is a set G together with a binary operation · satisfying the following axioms.

(i) G is closed under the operation ·; that is, a · b G for all a, b G. (ii) The operation · is associative; that is, (a · b) · c = a · (b · c) for all

a, b, c G.

(iii) There is an identity element e G such that e · a = a · e = a for all a G.

(iv)Each element a G has an inverse element a1 G such that a1 · a = a · a1 = e.

The closure axiom is already implied by the definition of a binary operation; however, it is included because it is often overlooked otherwise.

If the operation is commutative, that is, if a · b = b · a for all a, b G, the group is called commutative or abelian, in honor of the mathematician Niels Abel.

Let G be the set of complex numbers {1, 1, i, i} and let · be the standard multiplication of complex numbers. Then (G, ·) is an abelian group. The product of any two of these elements is an element of G; thus G is closed under the operation. Multiplication is associative and commutative in G because multiplication of complex numbers is always associative and commutative. The identity element is 1, and the inverse of each element a is the element 1/a. Hence 11 = 1, (1)1 = −1, i1 = −i, and (i)1 = i. The multiplication of any two elements of G can be represented by Table 3.1.

The set of all rational numbers, Q, forms an abelian group (Q, +) under addition. The identity is 0, and the inverse of each element is its negative. Similarly, (Z, +), (R, +), and (C, +) are all abelian groups under addition.

If Q , R , and C denote the set of nonzero rational, real, and complex numbers, respectively, (Q , ·), (R , ·), and (C , ·) are all abelian groups under multiplication.

For any set X, (P(X), ) is an abelian group. The group axioms follow from Proposition 2.3; the empty set, Ø, is the identity, and each element is its own inverse.

TABLE 3.1. Group {1, 1, i, i}

·

1

1

i

i

1

1

1

i

i

1

1

1

i

i

i

i

i

1

1

i

i

i

1

1

GROUPS AND SYMMETRIES

49

Every group must have at least one element, namely, its identity, e. A group with only this one element is called trivial. A trivial group takes the form ({e}, ·), where e · e = e.

Many important groups consist of functions. Given functions f : X Y and g: Y Z, their composite g Ž f : X Z is defined by

(g Ž f )(x) = g(f (x)) for all x X.

Composition is associative; that is, if h: Z W , then h Ž (g Ž f ) = (h Ž g) Ž f . Indeed,

(h Ž (g Ž f ))(x) = h(g(f (x))) = ((h Ž g) Ž f )(x)

for all x X, as is readily verified. In particular, if X is a set, then Ž is an associative binary operation on the set of all functions f : X X. Moreover, this operation has an identity: The identity function 1X : X X is defined by

 

 

1X (x) = x

for all x X.

Then

1X Ž f = f = f Ž 1X for all f : X X. Hence, we say that a function

f : X

X is an inverse of f : X

X if

 

 

 

 

f

Ž f = 1X

and f Ž f = 1X ;

equivalently if f (f (x)) = x and f (f (x)) = x for all x X. This inverse is unique when it exists. For if f is another inverse of f , then

f = f Ž 1X = f Ž (f Ž f ) = (f Ž f ) Ž f = 1X Ž f = f .

When it exists (see Theorem 3.3) the inverse of f is denoted f 1.

Example 3.1. A translation of the plane R2 in the direction of the vector (a, b) is a function f : R2 R2 defined by f (x, y) = (x + a, y + b). The composition of this translation with a translation g in the direction of (c, d) is the function f Ž g: R2 R2, where

f Ž g(x, y) = f (g(x, y)) = f (x + c, y + d) = (x + c + a, y + d + b).

This is a translation in the direction2 of (c + a, d + b). It can easily be verified

that the set of all translations in R

forms an abelian group, (T(2), Ž ), under

composition. The identity is the identity transformation 1R2 : R2 R2, and the inverse of the translation in the direction (a, b) is the translation in the opposite direction (a, b).

A function f : X Y is called injective or one-to-one if f (x1) = f (x2) implies that x1 = x2. In other words, an injective function never takes two different points to the same point. The function f : X Y is called surjective or

50 3 GROUPS

onto if for any y Y , there exists x X with y = f (x), that is, if the image f (X) is the whole set Y . A bijective function or one-to-one correspondence is a function that is both injective and surjective. A permutation or symmetry of a set X is a bijection from X to itself.

Lemma 3.2. If f : X Y and g: Y Z are two functions, then:

(i) If f and g are injective, g Ž f is injective.

(ii)If f and g are surjective, g Ž f is surjective.

(iii)If f and g are bijective, g Ž f is bijective.

Proof. (i) Suppose that (g Ž f )(x1) = (g Ž f )(x2). Then g(f (x1)) = g(f (x2)) so, since g is injective, f (x1) = f (x2). Since f is also injective, x1 = x2, proving that g Ž f is injective.

(ii) Let z Z. Since g is surjective, there exists y Y with g(y) = z, and since f is also surjective, there exists x X with f (x) = y. Hence (g Ž f )(x) =

g(f (x)) = g(y) = z, so g Ž f is surjective.

 

(iii) This follows from parts (i) and (ii).

The following theorem gives a necessary and sufficient condition for a function to have an inverse.

Theorem 3.3. Inversion Theorem. The function f : X Y has an inverse if and only if f is bijective.

Proof. Suppose that h: Y X is an inverse of f . The function f is injective because if f (x1) = f (x2), it follows that (h Ž f )(x1) = (h Ž f )(x2), and so x1 = x2. The function f is surjective because if y is any element of Y and x = h(y), it follows that f (x) = f (h(y)) = y. Therefore, f is bijective.

Conversely, suppose that f is bijective. We define the function h: Y X as follows. For any y Y , there exists x X with y = f (x). Since f is injective, there is only one such element x. Define h(y) = x. This function h is an inverse to f because f (h(y)) = f (x) = y, and h(f (x)) = h(y) = x.

Theorem 3.4. If S(X) is the set of bijections from any set X to itself, then (S(X), Ž ) is a group under composition. This group is called the symmetric group or permutation group of X.

Proof. It follows from Lemma 3.1 that the composition of two bijections is a bijection; thus S(X) is closed under composition. The composition of functions is always associative, and the identity of S(X) is the identity function 1X : X X. The inversion theorem (Theorem 3.3) proves that any bijective function f S(X) has an inverse f 1 S(X). Therefore, (S(X), Ž ) satisfies all the axioms for a group.

GROUPS AND SYMMETRIES

51

TABLE 3.2. Symmetry

Group of {a, b}

Ž

1X

f

1X

1X

f

f

f

1X

For example, if X = {a, b} is a two-element set, the only bijections from X to itself are the identity 1X and the symmetry f : X X, defined by f (a) = b, f (b) = a, that interchanges the two elements. The use of the term symmetry to describe the bijection f agrees with one of our everyday uses of the word. In the phrase “the boolean expression (a b) (a b ) is symmetrical in a and b” we mean that the expression is unchanged when we interchange a and b. The symmetric group of X, S(X) = {1X, f } and its group table is given in Table 3.2. The composition f Ž f interchanges the two elements a and b twice; thus it is the identity.

Since the composition of functions is not generally commutative, S(X) is not usually an abelian group. Consider the elements f and g in the permutation group of {1, 2, 3}, where f (1) = 2, f (2) = 3, f (3) = 1 and g(1) = 1, g(2) =

3, g(3) = 2. Then f Ž g(1) = 2, f Ž g(2) = 1, f Ž g(3) = 3,

while g Ž f (1) = 3,

g Ž f (2)

=

2, g Ž f (3)

=

1; hence f Ž g

=

{

1, 2, 3

}

) is not abelian.

 

 

 

g Ž f , and S(

 

A nonsingular linear transformation of the plane is a bijective function of the form f : R2 R2, where f (x, y) = (a11x + a12y, a21x + a22y) with the determinant a11a22 a12a21 = 0. It can be verified that the composition of two such linear transformations is again of the same type. The set of all nonsingular linear transformations, L, forms a non-abelian group (L, Ž ).

Besides talking about the symmetries of a distinct set of elements, we often refer, in everyday language, to a geometric object or figure as being symmetrical. We now make this notion more mathematically precise.

If F is a figure in the plane or in space, a symmetry of the figure F or isometry of F is a bijection f : F F which preserves distances; that is, for all points p, q F , the distance from f (p) to f (q) must be the same as the distance from p to q.

One can visualize this operation by imagining F to be a solid object that can be picked up and turned in some manner so that it assumes a configuration identical to the one it had in its original position. For example, the design on the left of Figure 3.1 has two symmetries: the identity and a half turn about a vertical axis, called an axis of symmetry. The design in the center of Figure 3.1

Figure 3.1. Symmetrical designs.

52

3 GROUPS

has three symmetries: the identity and rotations of one-third and two-thirds of a revolution about its center.

However, both the one-third rotation and interchanging two vertices are symmetries of the equilateral triangle on the right in Figure 3.1, but there is a subtle difference: The rotation can be performed as a physical motion within the plane of the triangle (and so is called a proper symmetry or a proper rotation), while the reflection can only be accomplished as a physical motion by moving the triangle outside its plane (an improper symmetry or an improper rotation).

The set of all symmetries of a geometric figure forms a group under composition because the composition and inverse of two distance-preserving functions is distance preserving.

Example 3.5. Write down the table for the group of symmetries of a rectangle with unequal sides.

Solution. Label the corners of the rectangle 1, 2, 3, and 4 as in Figure 3.2. Any symmetry of the rectangle will send corner points to corner points and so will permute the corners among themselves. Denote the (improper) symmetry obtained by reflecting the rectangle in the horizontal axis through the center, by a; then a(1) = 4, a(2) = 3, a(3) = 2, and a(4) = 1. This symmetry can also be considered as a rotation of the rectangle through half a revolution about this horizontal axis. There is a similar symmetry, b, about the vertical axis through the center. A third (proper) symmetry, c, is obtained by rotating the rectangle in its plane through half a revolution about its center. Finally, the identity map, e, is a symmetry. These are the only symmetries because it can be verified that any other bijection between the corners will not preserve distances.

The group of symmetries of the rectangle is ({e, a, b, c}, Ž ), and its table, as shown in Table 3.3, can be calculated as follows. The symmetries a, b, and c are all half turns, so a Ž a, b Ž b, and c Ž c are full turns and are therefore equal to the identity. The function a Ž b acts on the corner points by a Ž b(1) = a(b(1)) = a(2) = 3, a Ž b(2) = 4, a Ž b(3) = 1, and a Ž b(4) = 2. Therefore, a Ž b = c. The other products can be calculated similarly.

This group of symmetries of a rectangle is sometimes called the Klein 4- group, after the German geometer Felix Klein (1849–1925).

We have seen that the group operation can be denoted by various symbols, the most common being multiplication, composition, and addition. It is conventional

 

b

1

2

 

a

 

c

4

3

Figure 3.2. Symmetries of a rectangle.

GROUPS AND SYMMETRIES

 

 

 

53

 

TABLE 3.3. Symmetry Group of

 

 

 

a Rectangle

 

 

 

 

 

 

 

 

 

 

 

 

 

Ž

 

e

a

b

c

 

 

e

 

e

a

b

c

 

a

 

a

e

c

b

 

b

 

b

c

e

a

 

c

 

c

b

a

e

 

 

 

 

 

 

 

 

to use addition only for abelian groups. Furthermore, the identity under addition is usually denoted by 0 and the inverse of a by a. Hence expressions of the form a · b1 and an = a · · · a, in multiplicative notation, would be written as a b and na = a + · · · + a, respectively, in additive notation.

In propositions and theorems concerning groups in general, it is conventional to use multiplicative notation and also to omit the dot in writing a product; therefore, a · b is just written as ab.

Whenever the operation in a group is clearly understood, we denote the group just by its underlying set. Therefore, the groups (Z, +), (Q, +), (R, +), and (C, +) are usually denoted just by Z, Q, R, and C, respectively. This should cause no confusion because Z, Q, R, and C are not groups under multiplication (since the element 0 has no multiplicative inverse). The symmetric group of X is denoted just by S(X), the operation of composition being understood. Moreover, if we refer to a group G without explicitly defining the group or the operation, it can be assumed that the operation in G is multiplication.

We now prove two propositions that will enable us to manipulate the elements of a group more easily. Recall from Proposition 2.10 that the identity of any binary operation is unique. We first show that the inverse of any element of a group is unique.

Proposition 3.6. Let be an associative binary operation on a set S that has identity e. Then, if an element a has an inverse, this inverse is unique.

Proof. Suppose that b and c are both inverses of a; thus a b = b a = e, and a c = c a = e. Now, since e is the identity and is associative,

b = b e = b (a c) = (b a) c = e c = c.

Hence the inverse of a is unique.

Note that if ab = e in a group G with identity e, then a1 = b and b1 = a. Indeed, b has an inverse b1 in G, so b1 = eb1 = (ab)b1 = ae = a. Similarly, a1 = b.

Proposition 3.7. If a, b, and c are elements of a group G, then

(i)

(a1)1 = a.

 

(ii)

(ab)1 = b1a1.

 

(iii)

ab = ac or ba = ca implies that b = c.

(cancellation law)