Advanced Linear Algebra With an Introduction to Module Theory 1st Edition Shou-Te Chang pdf download
Advanced Linear Algebra With an Introduction to Module Theory 1st Edition Shou-Te Chang pdf download
https://ebookgate.com/product/advanced-linear-algebra-with-an-
introduction-to-module-theory-1st-edition-shou-te-chang/
https://ebookgate.com/product/advanced-calculus-an-introduction-to-
linear-analysis-1st-edition-leonard-f-richardson/
ebookgate.com
https://ebookgate.com/product/introduction-to-linear-algebra-3rd-
edition-gilbert-strang/
ebookgate.com
https://ebookgate.com/product/concise-introduction-to-linear-
algebra-1st-edition-qingwen-hu/
ebookgate.com
https://ebookgate.com/product/advanced-linear-algebra-2nd-edition-
bruce-n-cooperstein/
ebookgate.com
Advanced Linear Algebra Second Edition Nicholas A. Loehr
https://ebookgate.com/product/advanced-linear-algebra-second-edition-
nicholas-a-loehr/
ebookgate.com
https://ebookgate.com/product/linear-algebra-a-modern-
introduction-2nd-edition-david-poole/
ebookgate.com
https://ebookgate.com/product/elementary-linear-algebra-with-
applications-9th-edition-bernard-kolman/
ebookgate.com
https://ebookgate.com/product/linear-algebra-with-applications-global-
edition-steven-j-leon/
ebookgate.com
https://ebookgate.com/product/an-introduction-to-generalized-linear-
models-third-edition-barnett/
ebookgate.com
Advanced
Linear Algebra
With an Introduction to
Module Theory
This page intentionally left blank
Advanced
Linear Algebra
With an Introduction to
Module Theory
Shou-Te Chang
National Chung Cheng University, Taiwan
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
Preface
This book is suitable for students who want to go further into any
algebraic disciplines. Previously, the author had published two textbooks
(both published by World Scientific) with Professor Minking Eie, National
Chung Cheng University, Taiwan. In the first book, A First Course in
Linear Algebra,1 the reader will familiarize oneself with concepts such as
bases, dimension, matrices, and what not. There is a detailed discussion on
the spectral theorem and a brief introduction to Cayley-Hamilton theorem
and Jordan forms without proof. The topics of the second book, A Course
on Abstract Algebra, 2e,2 include introductions to group theory, ring theory
and some field theory. For a solid foundation of abstract algebra, the author
strongly feels that, in addition to familiarity with contents of the previously
mentioned two books (or of other similar textbooks), a third textbook is
necessary. The author believes that, equipped with knowledge covered by
these three books, a student of mathematics can delve into any algebraic
disciplines with complete confidence.
Linear algebra is considered one of the most basic mathematical tools,
even for non-mathematicians. However, a student who aspires to a career
in mathematics often finds a gaping gap between a first course on linear
algebra and the linear algebra needed in graduate schools. For example, an
elementary textbook on linear algebra will not provide proofs to results on
topics about or related to canonical forms of a square matrix. The proofs
would be hard to follow even if they are provided. Another example is
1 ISBN 9789813143111
2 ISBN 9789813229624
v
vi Preface
that such a textbook can only focus on finite dimensional vector spaces,
while graduate students often have to deal with infinite dimensional vector
spaces. These materials are considered too advanced for undergraduates
while the same materials are simply assumed in the graduate school. This
book fills the need for students who want to fill the gap themselves. It will
provide a gentle guidance to motivated students for self-study. It is also a
suitable textbook for an advanced undergraduate course on linear algebra
or on module theory. Its introduction to sets and modules would be of
particular interest to students who aspire to becoming algebraists.
One of the main purposes in this book is the introduction of module
theory. The concept of modules is a generalization of the concept of vector
spaces. For modules, the scalars only need to form a commutative ring
with unity instead of a field. We will find most concepts regarding vector
spaces have their counterparts in module theory. Why can’t we consider a
matrix with entries only in Z? How do we determine whether a matrix over
Z is invertible? And if the inverse matrix exists, how do we find it? We
see that many questions answered in linear algebra remain to be explored
in module theory. The study of modules opens up a whole new world and
new adventures. The results and insights gained in module theory actually
also help us further understand linear algebra.
What preparation do we need for using this book? A basic understand-
ing of set theory, group theory and ring theory is preferred. Here is a
checklist.
• Some basic set theory, including the equivalence relation, the well-
ordering principle, the pigeonhole principle and the induction princi-
ple.
• Definitions of groups, rings, fields and vector spaces. The basic prop-
erties of these algebraic structures. Morphisms of each structure and
their basic properties.
• The Euclidean domain, the PID (principal ideal domain) and the
UFD (factorial domain or unique factorization domain).
Shou-Te Chang
March 31, 2023
About the Author
viii
Contents
Preface v
Special Notes xi
2 Linear Maps 59
2.1 Linear maps . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.2 Quotient modules and quotient spaces . . . . . . . . . . . . 67
2.3 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3 Determinant 89
3.1 Basics of the determinant . . . . . . . . . . . . . . . . . . . 90
ix
x Contents
Index 255
Special Notes
3 There are two schools of mathematicians in the world, one who views 0 as a natural
xi
This page intentionally left blank
CHAPTER 1
1
2 1. Modules and Vector Spaces
• (a + b) · c = a · c + b · c for all a, b, c ∈ R;
• a · (b + c) = a · b + a · c for all a, b, c ∈ R.
In this book, the only rings of interest are the commutative rings with
unity. Unless otherwise noted, when we say that R is a ring we mean that
R is a commutative ring with unity.
If F is a commutative ring with unity which further satisfies the as-
sumption that every nonzero element has a multiplicative inverse, we say
that F is a field.
In short, a ring is where you can add, subtract and multiply. A field is
where you can add, subtract, multiply and divide (by a nonzero element).
Common examples of rings include Z, Zn and R[ x ], the polynomial ring
of one variable over the ring R. Common examples of fields include Q, R,
C and Zp where p is a positive prime integer.
R×M −→ M
(a, m) 7−→ a·m
such that
(i) 1 · m = m,
(ii) a · (m + n) = a · m + a · n,
(iii) (a + b) · m = a · m + b · m, and
(iv) a · (b · m) = (ab) · m
However, for lack of a better term, we will also call a · m the scalar
multiplication of a and m even when R is not a field. When there leaves
no confusion, the element a · m is often abbreviated as am. In fact, we will
do just that for the rest of the book.
Always remember that a vector space is also a module. All properties
of modules apply to vector spaces.
Next are some examples of modules and vector spaces.
Example 1.1.2. The trivial group {0} has a natural module structure.
We call it the trivial module.
0Z · a = 0G ,
k · a = a + · · · + a, and
| {z }
k copies
for any a ∈ G and any positive integer k. The verification that G is thus a
Z-module is tedious but straightforward. A question regarding an abelian
group may often be translated as a question on a Z-module. Conversely, a
Z-module is an additive (abelian) group by definition. Hence, the concepts
of Z-modules and abelian groups are basically the same.
Example 1.1.5. Let V be the set of all continuous real-valued functions
defined on an interval I of R. There is a natural addition in V defined by
This follows from the fact (in a course of Advanced Calculus) that f + g
and af remain continuous on I. We leave it to the reader to verify that V
is an R-vector space.
Similarly, the set of differentiable real-valued functions on an open in-
terval I, or the integrable real-valued functions on a finite closed interval I
can both be regarded as R-vector spaces.
Let D be an open region in Rm . Consider the set W of vector fields
(vector-valued functions) with D as the domain and Rn as the codomain.
With the vector addition and scalar multiplication defined as in (1.1.1) and
(1.1.2), W is an R-vector space.
Example 1.1.6. The space Cn is a C-vector space. It is also an R-vector
space, since the scalar multiplication may be induced by restricting the
mapping C × Cn → Cn to R × Cn → Cn .
6 1. Modules and Vector Spaces
Definition 1.1.8. We will use (aij )m×n to denote an array of m rows and
n columns
a11 a12 ··· a1n
a21 a22 ··· a2n
.
. .. ..
. ..
. . . .
am1 am2 ··· amn
The aij ’s are called the entries of this array. If all the entries are elements
in a ring R, we call it an m × n matrix over R. We will use Mm×n (R) to
denote the set of all m × n matrices over R. An n × n matrix is called a
square matrix of size n. We will use Mn (R) to denote the set of square
matrices of size n.
We use In , or simply I when n is understood, to denote the identity
matrix of size n. To be precise, In = (aij )n×n where
0, if i ̸= j;
aij = the Kronecker deltaδij =
1, if i = j.
Example 1.1.9. Inside Mm×n (R) there is the usual matrix addition and
scalar multiplication:
r(aij ) = (raij )
where aij , bij and r ∈ R. These operations make Mm×n into an R-module.
The reasoning is similar to that of Example 1.1.3.
Let S and T be sets. We often use T S to denote the set of all functions
from S to T . In general, if S and T are both finite sets, we know that
(1.1.3) |T S | = |T ||S| .
Lemma 1.1.14 (Test for submodules and subspaces). Let R be a ring and
M be an R-module. A subset N of M is a submodule of M if and only if
the following three conditions are satisfied:
(i) 0 ∈ N ;
Proof. The “only if” part: We already have (ii) and (iii) from the Remark
to Definition 1.1.13. To show (i), we may find the additive identity n in N
since N is a module. We should have n + n = n. Hence 0 + n = n = n + n
in M . We have that 0 = n ∈ N by the cancelation law in M .
The “if” part: Condition (i) guarantees that N is nonempty. Condi-
tion (ii) guarantees that there is an inherited addition in N . Condition (iii)
further assures that −n = (−1)n ∈ N for each n ∈ N . (See Exercise 1(c).)
Hence N is an additive subgroup of M . Condition (iii) also guarantees
that there is an inherited scalar multiplication in N . Conditions (i)–(iv) in
Definition 1.1.1 hold in N since they hold true in M . We have shown that
N is an R-module.
Example 1.1.16. Note that the three conditions in Lemma 1.1.14 are
also used for testing ideals. This is because ideals of a ring R are also
submodules of R (see Exercise 3).
10 1. Modules and Vector Spaces
Exercises 1.1
(a) a0M = 0M .
(b) 0R m = 0M . (Normally we will just use the symbol 0 for either
0R or 0M when no confusion arises.)
(c) (−1)m = −m. (Here −1 stands for the additive inverse of 1 in
R and −m is the additive inverse of m in M .)
(d) If R is a field and am = 0, then a = 0 or m = 0.
Give an example to show that the assertion (d) is not necessarily true
if R is not a field.
1.1 Definitions and examples of modules and vector spaces 11
(i) N is nonempty;
(ii) an + n′ ∈ N whenever a ∈ R and n, n′ ∈ N .
is an R-submodule of Mn (R).
8. Suppose given an index set I and a family of sets {Xi }i∈I . Define
Y
Xi = {(xi )i∈I : xi ∈ Xi for each i ∈ I}
i∈I
L
9. Let Mi be an R-module for each i ∈ I. Define i∈I Mi to be the set
We call it the direct sum of the Mi ’s. This is a subset of the direct
Q L Q
product i∈I Mi . Show that i∈I Mi is an R-submodule of i∈I Mi
under the inherited addition and scalar multiplication.
Note that the direct sum and the direct product are the same when I
is finite, while the direct product and the direct sum are not the same
Q∞
when I is infinite. For example, (1, 1, 1, . . . ) is an element in i=1 Z
L∞
but not in i Z.
13. Let (M, +) be an additive (abelian) group. Show that the scalar
multiplication given in Example 1.1.4 is the only mapping that makes
M into a Z-module.
1.1 Definitions and examples of modules and vector spaces 13
14. Let V be a Q-vector space. Show that the given scalar multiplication
is the only mapping that makes (V, +) into a Q-vector space.
19. For this problem, we will assume that the ring R is not necessarily
commutative. In this case Definition 1.1.1 defines what is called a
left R-module or a left module over R.
We say that an additive group M is a right R-module or a right
module over R if there is a mapping
M ×R −→ M
(m, a) 7−→ m·a
such that
(i) m · 1 = m,
14 1. Modules and Vector Spaces
(ii) (m + n) · a = m · a + n · a,
(iii) m · (a + b) = m · a + m · b, and
(iv) (m · b) · a = m · (ab)
M ×R −→ M
(m, a) 7−→ m ·r a = a−1 ·l m.
Linear combinations
Definition 1.2.1. Let M be an R-module and let m1 , m2 , . . . , mn be
elements in M . We say an element of the form
n
X
ai mi , ai ∈ R,
i=1
Example 1.2.2. Consider u = (1, −1, 3), v = (1, −3, 4) and w = (1, 1, 2)
in Z3 . Is u a linear combination of v and w over Z? How about over Q?
If we solve the equation over Q, we will see that a = b = 1/2 is the only
solution. Hence (1, −1, 3) is not a linear combination of v and w over Z.
However, it is a linear combination of v and w over Q. ⋄
(−20)6 + (20)8 = 40 = 1
Remark. It is easy to see that ⟨∅⟩ = ⟨0⟩ = {0} is the trivial module.
Proof. (a) This follows from Definition 1.2.4 that ⟨S⟩ is the smallest sub-
module of M which contains S.
(b) Since S ⊆ T ⊆ ⟨T ⟩, the result follows form (a).
(c) The “if” part: This is true since m ∈ S ∪ {m} ⊆ ⟨S ∪ {m}⟩ = ⟨S⟩.
The “only if” part: It remains to verify ⟨S⟩ ⊇ ⟨S ∪ {m}⟩ from (b). Since
m ∈ ⟨S⟩, we have that ⟨S⟩ ⊇ S ∪ {m}. The result follows from (a).
Solution. At this point we have no adequate tools for this problem yet, and
so we will use brutal force for now. However, since we have mentioned in
Example 1.1.4 that Z-modules may be viewed as abelian groups, we may
use Chinese remainder theorem (CRT) for Z to give us a hint.
(a) From CRT we have that Z57 × Z8 ≃ Z57·8 is cyclic as a group. By
reviewing the proof of CRT, we can see that (1, 1) is a generator. To view
this as a Z-module, observe that 57 − 7 · 8 = 1. We have
(1, 0) = − 56(1, 1)
(0, 1) = 57(1, 1)
in Z57 × Z8 , and
ei = (0, . . . , 0, 1, 0, . . . , 0)
↑
(a1 , a2 , . . . , an ) = a1 e1 + · · · + an en .
Linear independence
Definition 1.2.12. Let R be a ring and let M be an R-module. We say
that m1 , m2 , . . . , mn are linearly independent over R if for a1 , . . . , an
in R we have
a1 m1 + · · · + an mn = 0 =⇒ a1 = · · · = an = 0.
Proof. (b) This follows from the fact that any non-trivial relation (among
finitely many elements) in S is also a non-trivial relation in T .
The statement (a) is equivalent to (b).
Example 1.2.14. In Z2 , the elements (2, 3) and (3, −5) are linearly inde-
pendent over Z. This is true since
Then a1 = a2 = · · · = an = 0.
From the examples above we can see that the concept of linear indepen-
dence is more complicated for modules. For vector spaces, the situation is
more straightforward.
• v1 ̸= 0, and
• vi ̸∈ Sp{v1 , v2 , . . . , vi−1 } for i = 2, . . . , n.
a1 v1 + a2 v2 + · · · + ai vi = 0
vi = −a−1
i (a1 v1 + a2 v2 + · · · + ai−1 vi−1 ) ∈ Sp{v1 , v2 , . . . , vi−1 }.
(c) From (b), the set {v1 , v2 , . . . , vn−1 } is linearly independent over F
and vn ̸∈ Sp{v1 , v2 , . . . , vn−1 } if and only if
• v1 ̸= 0,
• vn ̸∈ Sp{v1 , v2 , . . . , vn−1 }.
Example 1.2.19. The set {2, 3} is clearly linearly dependent over Z since
3 · 2 + (−2) · 3 = 0.
However, 3 ̸∈ Z2, which is the module of even integers. This example shows
that Proposition 1.2.18 does not apply to modules.
Exercises 1.2
1. Consider the R-vector space RR . Are the two functions cos x and sin x
linearly independent over R?
M + N = { m + n ∈ L : m ∈ M, n ∈ N }
L ∩ (M + N ) = (L ∩ M ) + (L ∩ N )
M = M1 ⊕ · · · ⊕ Mn
8. Let V be a vector space over the field F and let S be a (not neces-
sarily finite) subset of V . Let v ∈ V . Show that S ∪ {v} is linearly
independent over F if and only if S is linearly independent over F
and v ̸∈ Sp S. This is a generalization of Proposition 1.2.18(c).
1.3 Bases
The concept of basis plays an essential role in the study of vector spaces.
In this section we will examine whether this concept also makes sense for
modules.
24 1. Modules and Vector Spaces
Bases
Definition 1.3.1. In an R-module, we say a set B is a base, basis or free
basis for M over R if
• B generates M , and
Posets
Before we give a characterization to the concept of bases, we need to
introduce the concept of “order” or “partial order”.
(i) (Reflexivity) a ≤ a;
Example 1.3.7. Let S be an arbitrary set and let P(S) be the power set
of S (the set of all the subsets of S). Then (P(S), ⊆) is a poset.
Example 1.3.9. Let ≤ be the relation “less than or equal to” in the usual
sense. Let | denote the relation of divisibility, that is, a|b if and only if
b = ac for some c. Then (Z+ , ≤) and (Z+ , | ) are both posets. However,
note that (Z, ≤) is a poset while (Z, | ) is not! The relation | does not satisfy
antisymmetry in Z. For example, we have 1| − 1 and −1|1 but 1 ̸= −1 in Z.
Remark. One may notice the word “the” in the phrases “the greatest” and
“the least”. The word “the” in mathematics means “the unique” or “one
and only”. Suppose g and h are both greatest elements in a subset T within
a poset S. Then g ≤ h and h ≤ g by definition. This implies that g = h by
antisymmetry.
Example 1.3.11. Consider the poset (Z, ≤) in Example 1.3.9. There are
neither maximal nor minimal elements in it.
In this example, we can see that maximal or minimal elements may not
exist in a poset.
Example 1.3.12. Let S = {1, 2, 3, 4}. In the poset (P(S), ⊆), the empty
set ∅ is the only minimal element; it is also the least element in P(S).
Similarly, the set S is the greatest element and the only minimal element
in P(S).
Consider the poset (A , ⊆) where
A = {{1}, {3}, {1, 2}, {2, 3}, {2, 4}, {2, 3, 4}} ⊆ P(S).
The elements {1}, {3} and {2, 4} are minimal and the elements {1, 2} and
{2, 3, 4} are maximal. There are neither least nor greatest elements in A .
1.3 Bases 27
In this example, we can see that in a poset: (i) maximal elements (or
minimal elements respectively) may not be unique even when they exist,
and (ii) a minimal (or maximal) element may not be the least (or greatest,
respectively) element.
v = a−1 (−a1 v1 − · · · − an vn ) ∈ Sp B,
v = a1 v1 + · · · + an vn , for a1 , a2 , . . . , an ∈ F.
28 1. Modules and Vector Spaces
v1 = a−1
1 (−a2 v2 − · · · − an vn ) ∈ Sp{v2 , . . . , vn } ⊆ Sp(B \ {v1 }).
Proof. Find a finite generating set for the given finite dimensional vector
space. If this generating set is not minimal, we may find a proper generating
subset within this generating set. Continue with the procedure and we may
eventually reach a minimal generating subset since we started with a finite
set. From Proposition 1.3.13 this minimal generating subset is a basis for
the given vector space.
Example 1.3.17. Let ≤ be the relation “less than or equal to” in the usual
sense. The posets (Z+ , ≤) and (Z, ≤) are totally ordered.
(1.3.1) a1 v1 + a2 v2 + · · · + an vn = 0.
Even though we are sure of the existence of bases, it does not mean that
Q∞
we can actually find them. Good luck finding a specific basis for i=1 R
over R. Good luck finding a specific basis for R over Q.
Exercises 1.3
L∞
2. Is i=1 R a free module over R? If yes, find an R-basis for it. Is
L
i∈R R a free module over R? If yes, find an R-basis for it.
5. Let S = {1, 2, 3}. Show that (R2 )S is free over R by finding an R-basis
for it.
Replacement theorem
Theorem 1.4.1. All bases of a finite dimensional vector space are finite
and have the same number of elements in them.
B ∩ S = {v1 , v2 , . . . , vs }
such that
Bs+1 = Bs \ {ws+1 } ∪ {vs+1 }
= B \ {w1 , w2 , . . . , ws+1 } ∪ {v1 , v2 , . . . , vs+1 }
(1.4.1) avs+1 + a1 v1 + a2 v2 + · · · + as vs + b1 u1 + b2 u2 + · · · + bd ud = 0
ebookgate.com