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

5

SYMMETRY GROUPS IN THREE DIMENSIONS

In this chapter we determine the symmetry groups that can be realized in twoand three-dimensional space. We rely heavily on geometric intuition, not only to simplify arguments but also to give geometric flavor to the group theory. Because we live in a three-dimensional world, these symmetry groups play a crucial role in the application of modern algebra to physics and chemistry.

We first show how the group of isometries of Rn can be broken down into translations and orthogonal transformations fixing the origin. Since the orthogonal transformations can be represented as a group of matrices, we look at the properties of matrix groups. We then use these matrix groups to determine all the finite rotation groups in two and three dimensions, and we find polyhedra that realize these symmetry groups.

TRANSLATIONS AND THE EUCLIDEAN GROUP

Euclidean geometry in n dimensions is concerned with those properties that are preserved under isometries (rigid motions) of euclidean n-space, that is, bijections α: Rn Rn that preserve distance. The group of all isometries of Rn is called the euclidean group in n dimensions and is denoted E(n). Given w Rn, the map Rn Rn with v v + w is called translation by w, and we begin by showing that the group T (n) of all translations is a normal subgroup of E(n), and that the factor group is isomorphic to the group of all orthogonal n × n matrices (that is, matrices A such that A1 = AT , the transpose of A—reflection of A in its

main diagonal).

λ

Rn

 

Rn

 

 

 

 

 

 

 

λ(a

 

 

 

 

 

is called a linear transformation if

v +

Recall that a function :

 

 

 

 

n

 

bw)

=

aλ(v)

+

bλ(w) for all a, b

 

R

and all

(column) vectors v, w

 

R

 

. Let

 

 

denote the

 

 

 

 

R

n

, that is,

 

 

 

 

the

{e1, e2, . . . , en}

standard basis of

 

the columns of

Modern Algebra with Applications, Second Edition, by William J. Gilbert and W. Keith Nicholson ISBN 0-471-41451-4 Copyright 2004 John Wiley & Sons, Inc.

= AT ,

TRANSLATIONS AND THE EUCLIDEAN GROUP

105

n

×

n

identity

matrix. Then the action of λ is matrix multiplication

λ(v)

=

 

 

 

n

 

 

Av for all v in R

 

, where the matrix A is given in terms of its columns by

A = [λ(e1)λ(e2) · · · λ(en)] and is

called the standard matrix of α. Moreover,

the correspondence λ A is a

bijection that preserves addition, multiplica-

tion, and the identity. So we may (and sometimes shall) identify λ with the

matrix A.

 

 

If v and w are vectors in Rn, let vžw

 

 

vT w denote their inner product. Then

||

|| =

 

ž

n

 

 

n

 

 

 

||

v

 

||

is the distance between v and w.

v

 

v v is the length of v, and

 

= w

 

Thus a function α: R R is an isometry if

 

 

 

 

 

 

 

||

α(v)

α(w)

 

 

v

w

||

 

for all v, w

 

Rn.

( )

 

 

 

 

 

|| = ||

 

 

 

 

 

 

 

 

 

Since ||v w||2 = ||v||2 + 2(vžw) + ||w||2

for any v, w Rn, it follows from ( )

that every isometry α preserves inner products in the sense that

 

 

 

 

 

 

α(v)

α(w)

=

v w

 

 

for all v, w

 

Rn.

 

( )

 

 

 

 

 

 

 

ž

 

 

 

ž

 

 

 

 

 

 

 

 

 

 

 

Lemma 5.1. If α: Rn Rn is an isometry such that α(0) = 0, then α is linear.

Proof. It follows from ( ) that

{

α(e1), α(e2), . . . , α(en) is an orthonormal

 

 

}

basis of Rn. If a R and v Rn, then ( ) implies that

 

[α(av) aα(v)]žα(ei ) = (av)žei a(vžei ) = 0

for each i.

Hence α(av) = naα(v), and α(v + w) = α(v) + α(w) follows in the same way

for all v, w R .

 

Hence the isometries of Rn that fix the origin are precisely the linear isometries. An n × n matrix A is called orthogonal if it is invertible and A1

equivalently if the columns of A are an orthonormal basis of Rn. These matrices form a subgroup of the group of all invertible matrices, called the orthogonal group and denoted O(n).

Proposition 5.2. Let λ: Rn Rn be a linear transformation with standard matrix A.

(i)λ is an isometry if and only if A is an orthogonal matrix.

(ii)The group of linear isometries of Rn is isomorphic to O(n).

Proof. If A is orthogonal, then for all v, w Rn,

 

||

λ(v)

λ(w)

2

=

A(v

w)žA(v

w)

=

(v

w)T AT A(v

w)

= ||

v

w

2

 

 

 

 

 

||

 

 

 

 

 

 

 

 

||

 

and it

 

follows

that

λ is

an isometry. Conversely, if λ is an

isometry and

{

e1, e2

, . . . , en

}

is the standard basis of Rn, then ( ) gives

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ei žej = λ(ei )žλ(ej ) = Aei žAej = eTi (AT A)ej = the (i, j )-entry of A

106 5 SYMMETRY GROUPS IN THREE DIMENSIONS

for all i and j . It follows that AT A = I , so A is orthogonal, proving (i). But then the correspondence λ A between the linear transformation λ and its standard matrix A induces a group isomorphism between the (linear) isometries fixing the origin and the orthogonal matrices. This proves (ii).

Given a vector w Rn, define τw: Rn Rn by τw(v) = v w for all v Rn. Thus τw is the unique translation that carries w to 0. Because

τw Ž τw = τw+w for all w, w Rn,

=

the correspondence w

τw is a group isomorphism (Rn,

+

 

 

) T (n). In partic-

ular, T (n) is an abelian group.

Theorem 5.3. For each n 1, T (n) is an abelian normal subgroup of E(n) and

=

O(n) given by

E(n)/T (n) O(n). In fact, the map E(n)

 

α the standard matrix of τα(0) Ž α

is a surjective group morphism E(n) O(n) with kernel T (n).

Proof. Write G(n) for the group of all linear isometries of Rn. Observe that if α E(n), then τα(0) Ž α is linear (it is an isometry that fixes 0), so we have a map

φ: E(n) G(n) given by φ(α) = τα(0) Ž α for all α E(n).

By Proposition 5.2 it suffices to show that φ is a surjective group morphism with kernel T (n). To see that φ is a group morphism, observe first that

 

α

Ž

τw

=

τα(w)

Ž

α

for all w

 

Rn.

 

( )

 

 

 

 

 

 

 

 

 

 

[Indeed,

Ž τw)(v) = α(v w) = α(v) α(w) = α(w) Ž α)(v)

for

all

v.]

Hence, given α and β in E(n), we have

 

 

 

 

 

 

φ(α) Ž φ(β) = τα(0) Ž α Ž τβ(0) Ž β = τα(0) Ž Ž τβ(0) Ž α1) Ž Ž β),

 

 

so it suffices to show that

τα(0) Ž Ž τβ(0) Ž α1) = τŽ β)(0). But

this

follows

because

α Ž τβ(0) Ž α1 is

a translation by

( ), so

τα(0) Ž Ž τβ(0) Ž α1) is

the

unique translation that carries Ž β)(0) to 0. Hence φ is a group morphism.

 

φ

 

φ(λ)

=

λ

 

λ

 

G(n)

Ker(φ)

=

Moreover,

 

is surjective because

 

 

for every

 

, and

 

n

 

T (n) because φ(α) = 1Rn if and only if α(v) = v + α(0) for all v R

 

, that is,

if and only if α = τα(0).

 

 

 

 

 

 

 

 

 

 

 

If α E(n) and α(0) = w, the proof of Theorem 5.3 shows that τw Ž α G(n). Hence every isometry α of Rn is the composition of a linear isometry τw1 Ž α followed by a translation τw.

MATRIX GROUPS

107

Proposition 5.4. Every finite subgroup of isometries of n-dimensional space fixes at least one point.

Proof. Let G be a finite subgroup of isometries, and let x be any point of n-dimensional space. The orbit of x consists of a finite number of points that are permuted among themselves by any element of G. Since all the elements of G are rigid motions, the centroid of Orb x must always be sent to itself. Therefore, the centroid is a fixed point under G.

If the fixed point of any finite subgroup G of isometries is taken as the origin, then G is a subgroup of O(n), and all its elements can be written as orthogonal matrices. We now look at the structure of groups whose elements can be written as matrices.

MATRIX GROUPS

In physical sciences and in mathematical theory, we frequently encounter multiplicative group structures whose elements are n × n complex matrices. Such a group is called a matrix group if its identity element is the n × n identity matrix I . To investigate these groups, we have at our disposal, and shall freely apply, the machinery of linear algebra.

For example, if

Ak =

sin(2π k/m)

cos(2π k/m)

,

 

cos(2π k/m)

sin(2π k/m)

 

then ({A0, A1, . . . , Am1}, ·) is a real matrix group of order m isomorphic to Cm. The matrix Ak represents a counterclockwise rotation of the plane about the origin through an angle (2π k/m).

The matrices

0

 

1

 

0 i

 

 

0

 

 

 

 

 

 

 

0

 

0

1

,

 

,

,

 

i

,

 

1

,

1

0

 

 

1

0

 

i

0

 

i

0

 

 

0

1

 

 

 

0

,

 

0

, and

 

i

 

 

 

 

 

 

 

 

 

1

i

 

0

 

 

 

 

 

 

 

 

 

 

0

1

 

 

0

 

i

 

 

 

0

i

 

 

 

 

 

 

 

 

 

 

form a group under matrix multiplication. This is a complex matrix group of order 8 that is, in fact, isomorphic to the quaternion group Q of Exercise 3.47.

Since the identity of any matrix group is the identity matrix I and every element of a matrix group must have an inverse, every element must be a nonsingular matrix. All the nonsingular n × n matrices over a field F form a group (GL(n, F ), ·) called the general linear group of dimension n over F . Any matrix group over the field F must be a subgroup of GL(n, F ).

Proposition 5.5. The determinant of any element of a finite matrix group must be an integral root of unity.

108 5 SYMMETRY GROUPS IN THREE DIMENSIONS

Proof. Let A be an element of a matrix group of order m. Then, by Corollary 4.10, Am = I . Hence (detA)m = detAm = detI = 1.

Hence, if G is a real matrix group, the determinant of any element of G is either +1 or 1. If G is a complex matrix group, the determinant of any element is of the form e2π ik/m.

The orthogonal group O(n) is a real matrix group, and therefore any element must have determinant +1 or 1. The determinant function

det: O(n) → {1, 1}

is a group morphism from (O(n), ·) to ({1, 1}, ·). The kernel, consisting of orthogonal matrices with determinant +1, is called the special orthogonal group of dimension n and is denoted by

SO(n) = {A O(n)|detA = +1}.

This is a normal subgroup of O(n) of index 2. The elements of SO(n) are called proper rotations, whereas the elements in the other coset of O(n) by SO(n), consisting of orthogonal matrices with determinant 1, are called improper rotations.

An n × n complex matrix A is called unitary if it is invertible and A1 is the conjugate transpose of A. Thus the real unitary matrices are precisely the

orthogonal matrices. Indeed,

if x, y

xT

 

 

in Cn,

y denotes the inner product

 

=

 

n

(that

the matrix A is unitary if and only if it preserves inner products in C

 

is, Ax, Ay = x, y for all

x, y Cn),

if and only if the columns of

A are

orthonormal.

 

 

 

 

 

 

The unitary group of dimension n, U(n), consists of all n × n complex unitary matrices under multiplication. The special unitary group, SU(n), is the subgroup of U(n) consisting of those matrices with determinant +1. The group SU(3) received some publicity in 1964 when the Brookhaven National Laboratory discovered the fundamental particle called the omega-minus baryon. The existence and properties of this particle had been predicted by a theory that used SU(3) as a symmetry group of elementary particles.

Proposition 5.6. If λ C is an eigenvalue of any unitary matrix, then |λ| = 1.

Proof. Let λ be an eigenvalue and x a corresponding nonzero eigenvector of

the unitary matrix A. Then Ax = λx, and since A preserves distances, ||x|| = ||Ax|| = ||λx|| = |λ|||x||. Since x is nonzero, it follows that |λ| = 1.

The group {Ak |k = 0, 1, . . . , m 1} of rotations of the plane is a subgroup of SO(2), and the eigenvalues that occur are e±(2π ik/m). The matrix group isomorphic to the quaternion group Q is a subgroup of SU(2), and the eigenvalues that occur are ±1 and ±i.