Notes FRM
Notes FRM
SERGEY MOZGOVOY
Contents
1. Rings 2
1.1. Basic definitions 2
1.2. Ideals and quotient rings 4
1.3. Ring homomorphisms 7
1.4. Algebras 10
2. Integral domains 14
2.1. Basic definitions 14
2.2. Unique factorization domains (UFD) 15
2.3. Principal ideal domains (PID) 17
2.4. GCD and LCM 18
2.5. Euclidean domains 20
2.6. The field of fractions 22
2.7. Factorization in polynomial rings 23
3. Fields 25
3.1. Basic definitions 25
3.2. Field extensions 27
3.3. Splitting fields, Finite fields, Algebraically closed fields 30
3.4. Constructions with compass and straightedge 33
4. Symmetric polynomials 37
4.1. Discriminant 39
5. Modules 40
5.1. Definition and examples 40
5.2. Homomorphisms and submodules 42
5.3. Simple and indecomposable modules 44
5.4. Chinese remainder theorem 47
5.5. Modules over PID 49
5.6. Noetherian modules 51
Appendix A. Cyclotomic polynomials 53
Appendix B. RSA algorithm 55
1. Rings
1.1. Basic definitions. Consider the set Z of integer numbers. It has two binary operations:
+ (addition) and · (multiplication) compatible with each other
We will use this example as a motivation for a formal description of a structure, called a ring,
consisting of a set and two compatible binary operations as above.
Definition 1.1. An abelian group is a pair (A, +), where A is a set and + : A × A → A is a
+
map (we write (a, b) 7−
→ a + b) such that
(1) (Associativity) (a + b) + c = a + (b + c).
(2) (Commutativity) a + b = b + a.
(3) (Existence of zero) ∃ 0 ∈ A such that 0 + a = a + 0 = a ∀a ∈ A.
(4) (Existence of negative) For any a ∈ A there exists an element b ∈ A such that a + b = 0.
It is denoted by −a.
Example 1.2. The set of natural numbers N = {0, 1, 2, 3, . . .} has an obvious addition operation.
But it is not a group: it does not contain negatives of nonzero elements. For example −1 6∈ N.
The following are examples of abelian groups
(1) The set Z of integer numbers.
(2) The set Q of rational numbers.
(3) The set R of real numbers.
(4) The set C of complex numbers.
(3) The set Mn (R) of n × n matrices with real entries is a ring. Addition and multiplication
of matrices A = (aij ) and B = (bij ) is defined by
n
X
A + B = (cij ), cij = aij + bij , AB = (dij ), dij = aik bkj .
k=1
The zero element of this ring is the zero matrix. The identity element of this ring is the
identity matrix
1 0 ··· 0
0 1 · · · 0
In =
. . . . . . . . . . . .
0 ...... 1
Similarly, the set Mn (C) of n × n matrices with complex entries is a ring. These rings
are not commutative for n ≥ 2. For example
( 00 10 )( 10 00 ) = ( 00 00 ), ( 10 00 )( 00 10 ) = ( 00 10 ).
They are also not division rings for n ≥ 2. For example, the matrix ( 10 00 ) is not invertible.
Lemma 1.6. Let A be a ring. Then
(1) The zero element is unique.
(2) The negative of any element is unique.
(3) The identity element is unique.
Proof. (1) If 00 is another zero, then 0 + 00 = 0 and 0 + 00 = 00 . Therefore 0 = 00 .
(2) Assume that an element a has two negatives b, b0 . Then
b = b + (a + b0 ) = (b + a) + b0 = 0 + b0 = b0 .
(3) If 10 is another identity then 1 · 10 = 1 = 10 .
Lemma 1.7. Let A be a ring. Then
(1) 0a = a0 = 0.
(2) (−a)b = a(−b) = −ab.
Proof. (1) We have
0a + a = 0a + 1a = (0 + 1)a = 1a = a.
Therefore 0a = 0. Similarly a0 = 0.
(2) We have
(−a)b + ab = (−a + a)b = 0b = 0.
Therefore (−a)b = −ab. Similarly a(−b) = −ab.
4 SERGEY MOZGOVOY
Remark 1.10. The subsets {0} ⊆ A and A ⊆ A are ideals of A. An ideal I ⊆ A is called
proper if it is a proper subset of A, meaning that I 6= A.
Example 1.11. Let A be a commutative ring and a ∈ A. Then the set (a) = aA = {ab | b ∈ A }
is an ideal of A. Indeed
(1) given two elements ab1 , ab2 ∈ (a), we have ab1 + ab2 = a(b1 + b2 ) ∈ (a).
(2) 0 = a0 ∈ (a).
(3) if ab ∈ (a), then −ab = a(−b) ∈ (a).
(4) A(aA) = (aA)A ⊆ aA.
Ideals of this form are called principal ideals. Note that 0A = {0} and 1A = A.
Definition 1.12. Let A be a commutative ring and let a, b ∈ A. We say that a divides b (or
that b is divisible by a, or b is a multiple of a) and we write a | b if there exists c ∈ A such that
b = ac. Note that a | b if and only if b ∈ (a) = aA.
Lemma 1.13. All ideals of Z are of the form (n) = nZ for some n ∈ Z.
Proof. Let I ⊆ Z be an ideal. If I = {0} then I = (0). Assume that I is nonzero. Let n be the
minimal positive element of I. We will prove that I = (n). Inclusion (n) ⊆ I is clear. Assume
that m ∈ I\(n). Dividing m by n with a remainder, we can write m = qn + r for some integers
q, r with 0 ≤ r < n. Actually 0 < r < n as m ∈/ (n). As m, n ∈ I, we obtain
r = m − qn ∈ I.
This contradicts to the minimality of the positive element n ∈ I.
(2) The element [1] ∈ A/I is the identity element: [a] · [1] = [a] = [1] · [a].
(3) Distributivity: [a]([b] + [c]) = [a] · [b + c] = [ab + ac] = [ab] + [ac] = [a] · [b] + [a] · [c].
Similarly, one can prove the second distributivity property.
Example 1.18. Let us consider the ring Z and the ideal (n) = nZ, for n ≥ 2. Then we have
the quotient ring Zn = Z/nZ, called the ring of congruence classes of integers modulo n. It
consists of n elements that are congruence classes of the elements 0, 1, . . . , n − 1 in Z. Indeed, for
any m ∈ Z, we can write m = qn + r for some q, r ∈ Z with 0 ≤ r < n. Then m − r = qn ∈ (n),
hence m ∼ r and [m] = [r]. If 0 ≤ r < r0 < n, then 0 < r0 − r < n, hence r0 − r ∈ / (n). Therefore
0 0
r 6∼ r and [r ] 6= [r].
(1) In the ring Z2 = Z/2Z we have [1] + [1] = [0] and [1] · [1] = [1]. This ring is a field.
(2) In the ring Z3 = Z/3Z we have [2] · [2] = [4] = [1] (as 4 ≡ 1 mod 3). This means that [2]
is invertible in Z3 and Z3 is a field.
(3) On the other hand, in Z4 = Z/4Z we have [2] · [2] = [4] = [0] (as 4 ≡ 0 mod 4). Therefore
[2] is not invertible in Z4 and Z4 is not a field.
The difference between the above rings stems from the fact that 2 and 3 are prime numbers,
while 4 is not. The general situation is described by the following theorem.
Remark 1.19. An integer p ≥ 2 is called a prime number if a | p implies a = ±1 or a = ±p.
Remark 1.20. Let a, b ∈ A be nonzero elements such that ab = 0. Then a is not invertible.
Indeed, if a has the inverse a−1 , then b = a−1 ab = a−1 0 = 0, which contradicts to b 6= 0.
Theorem 1.21. For n ≥ 2, the ring Zn = Z/nZ is a field if and only if n is a prime number.
Proof. =⇒ : Assume that n is not prime. Then n = km for some 1 < k, m < n. Therefore
[k], [m] 6= 0, but [k][m] = [n] = 0 in Zn . This implies that [k] is not invertible, hence Zn is not a
field.
⇐= : Assume that n = p is a prime number. Let 1 ≤ k < p be a number that represents
some nonzero element [k] in Zp . Then the multiplication map
k : Zp → Zp , [m] 7→ [k] · [m] = [km],
is injective: if not, then [km] = 0 for some [m] 6= 0. But this would imply that km ∈ (p), hence
p | km and therefore p | k or p | m (see the next lemma), a contradiction. As Zp is finite,
the map k : Zp → Zp should be bijective. This implies that there exists [m] ∈ Zp such that
[k] · [m] = [1], hence [k] is invertible. Therefore Zp is a field.
Lemma 1.22. Let p ∈ Z be a prime number and a, b ∈ Z. If p | ab, then p | a or p | b.
Proof. Let I = {n ∈ Z | p | na}. Then p, b ∈ I. The set I is an ideal in Z (exercise). Therefore
I = dZ for some d ≥ 1. As p ∈ I = dZ, we conclude that d | p and therefore d = 1 or d = p. If
d = 1, then p | da = a. If d = p then b ∈ I = pZ, hence p | b.
FIELDS, RINGS AND MODULES 7
φ φ̄
B
The map φ̄ is injective and induces an isomorphism φ̄ : A/ Ker φ → Im φ.
Proof. Uniqueness. As φ = φ̄ ◦ π, we have φ(a) = φ̄π(a) = φ̄([a]) ∀a ∈ A. This means that for
any equivalence class [a], we require φ̄([a]) = φ(a) and φ̄ is uniquely determined.
Existence. For any equivalence class [a], we define φ̄([a]) = φ(a).
The map φ̄ is well-defined: if a ∼ b then a − b ∈ Ker φ =⇒ φ(a − b) = 0 =⇒ φ(a) = φ(b).
FIELDS, RINGS AND MODULES 9
For any a ∈ A we have φ(a) = φ̄([a]) = φ̄π(a). Therefore φ = φ̄ ◦ π and the diagram commutes.
The map φ̄ is a ring homomorphism:
(1) φ̄([a] + [b]) = φ̄([a + b]) = φ(a + b) = φ(a) + φ(b) = φ̄([a]) + φ̄([b]).
(2) φ̄([a] · [b]) = φ̄([ab]) = φ(ab) = φ(a)φ(b) = φ̄([a])φ̄([b]).
(3) φ̄([1A ]) = φ(1A ) = 1B .
The map φ̄ : A/ Ker φ → B is injective: if φ̄([a]) = 0, then φ(a) = φ̄([a]) = 0 =⇒ a ∈ Ker φ
=⇒ [a] = 0 in A/ Ker φ. Therefore Ker φ̄ = 0 and φ̄ is injective.
Consider the map φ̄ : A/ Ker φ → Im φ ⊆ B. It is injective by the previous discussion. It is also
surjective: for any φ(a) ∈ Im φ, we have φ̄([a]) = φ(a). This implies that φ̄ : A/ Ker φ → Im φ is
bijective, hence is an isomorphism.
10 SERGEY MOZGOVOY
1.4.1. The algebra of matrices. Let R be a commutative ring and let Mn (R) be the set of
n × n matrices with entries in R. Then Mn (R) is a ring with respect to the usual addition and
multiplication: given matrices A = (aij ) and B = (bij ) in Mn (R), we define
n
X
A + B = (cij ), cij = aij + bij , AB = (dij ), dij = aik bkj .
k=1
The ring Mn (R) is an algebra over R. Indeed, we have an embedding (injective ring homomor-
phism)
R ,→ Mn (R), r 7→ rIn ,
where In is the identity matrix in Mn (R). In this way we can identify A with a subring of Mn (R).
The elements of A commute with all matrices
(rIn )A = A(rIn ) = rA ∀r ∈ R, A ∈ Mn (R).
The algebra Mn (R) over R is called the matrix algebra (or the matrix ring).
1.4.2. The algebra of quaternions. The algebra H of quaternions is the algebra over R having a
basis 1, i, j, k and the multiplication defined on this basis so that 1 is the identity element and
i2 = j 2 = k 2 = −1, ij = −ji = k, jk = −kj = i, ki = −ik = j.
Remark 1.38. According to a legend, this algebra was invented by William Rowan Hamilton on
October 16, 1843 while walking near the Broome Bridge, Dublin. This event is commemorated
by a stone plaque near the bridge. For a long time quaternions were a mandatory exam topic in
Trinity College Dublin.
Actually it is enough to require that
i2 = j 2 = k 2 = ijk = −1.
Indeed, i, j, k are invertible, hence ijk = k 2 implies ij = k =⇒ kj = ij 2 = −i and so on. For
any element x = a + bi + cj + dk ∈ H, we define the absolute value and the conjugate of x by
√
|x| = a2 + b2 + c2 + d2 , x̄ = a − bi − cj − dk.
Lemma 1.39. We have xx̄ = x̄x = |x|2 .
FIELDS, RINGS AND MODULES 11
Proof. We have
xx̄ = (a + bi + cj + dk)(a − bi − ci − dk) = a2 + b2 + c2 + d2 = |x|2
and similarly x̄x = |x|2 .
The above result implies that for any x 6= 0 we have
x̄ x̄
x · 2 = 2 · x = 1,
|x| |x|
hence x is invertible in H. This means that all nonzero elements of H are invertible, hence H is
a division ring. It is non-commutative (for example ij 6= ji).
The ring H is an algebra over R, where we embed R ⊆ H by the rule a 7→ a1 (note that the
elements of R commute with all elements of H). We can also embed C ⊆ H by the rule
a + bi 7→ a1 + bi ∈ H.
6 ji, implying
This makes C into a subring of H. But H is not an algebra over C: we have ij =
that i ∈ C does not commute with all elements of H.
1.4.3. The algebra of polynomials. Let A be a commutative ring. We define the algebra A[x] of
polynomials in one variable x with coefficients in A to be the set of sequences
f = (f0 , f1 , f2 , . . . ),
where fk ∈ A, k ≥ 0, and all but a finite number of elements fk are zero. We will write elements
f ∈ A[x] in a more customary form
X
f = f 0 + f 1 x + f 2 x2 + · · · = f k xk .
k≥0
Given two polynomials f, g ∈ A[x], we define their sum f + g ∈ A[x] and product f g ∈ A[x] by
k
!
X X X
f +g = (fk + gk )xk , fg = fi gk−i xk .
k≥0 k≥0 i=0
be a polynomial.
(1) We define the evaluation of f at b ∈ B (or the substitution of b into f ) to be
X
f (b) = fi bi ∈ B.
i≥0
12 SERGEY MOZGOVOY
Remark 1.43. A polynomial f ∈ A[x] induces a (polynomial) map f¯: A → A, a 7→ f (a). But
this map doesn’t determine f uniquely in general. For example, the polynomial f = x2 +x ∈ Z2 [x]
induces the map f¯: Z2 → Z2 , a 7→ 0, the same map as for the zero polynomial. We will see
later that polynomials over infinite fields are uniquely determined by the corresponding maps.
φb : A[x] → B, f 7→ f (b)
is a ring homomorphism, called the evaluation map. It is a unique ring homomorphism that
satisfies
φb (a) = a ∀a ∈ A, φb (x) = b.
Proof. It is clear that φb preserves the additive structures. Concerning the product, we have
k
! ! k
!
X X X X
φb (f g) = φb fi gk−i xk = fi gk−i bk
k≥0 i=0 k≥0 i=0
! !
X X
= f i bi gj bj = φb (f )φb (g).
i≥0 j≥0
We proved that p(A) = 0, hence p ∈ Ker φA and (p) = R[x]p ⊆ Ker φA . One can show that
actually Ker φA = (p).
Definition 1.46. For any commutative ring A, we define the algebra of polynomials in several
variables inductively by the rule
where fi1 ...in ∈ A and all but a finite number of these elements are zero.
FIELDS, RINGS AND MODULES 13
2. Integral domains
2.1. Basic definitions.
Remark 2.1. There is the following chain of commutative ring classes
and therefore
√ |zi |2 are equal to 2 or 3. But there are no such elements in A. This implies that
2, 3, 1 ± −5 are irreducible. They are not associates of each other as A× = {±1}. This implies
that the above factorizations are not equivalent.
Equality √ √
2 · 3 = 6 = (1 + −5)(1 − −5)
16 SERGEY MOZGOVOY
√ √ √
implies that 2 | (1 + √ −5)(1 − −5). But 2 does not divide 1 ± −5. This means that 2 is not
prime in the ring Z[ −5], although we have seen that 2 is irreducible.
Theorem 2.15. Let A be a factorization domain (satisfies just the first axiom of a UFD). Then
A is a UFD if and only if every irreducible element of A is prime.
Proof. ⇒. Let A be a UFD, p ∈ A be irreducible and p | ab, that is, ab = pc for some a, b, c ∈ A.
We want to show that p | a or p | b. Let
Y Y Y
a= ai , b= bi , c= ci
be factorizations into irreducible elements. Then the element ab = pc has two factorizations
Y Y Y
ai bi = p ci .
From the uniqueness of factorizations we conclude that p equals (up to a unit) to one of ai or bj .
This means that p divides a or b.
⇐. Assume that an element a has two decompositions into irreducibles
m
Y Yn
a= pi = qi .
i=1 i=1
By assumption all elements pi are prime. Let p = pm . Then p should divide one of qi (we can
assume that it is qn ). But qn is irreducible, therefore qn = up for some u ∈ A× . Dividing both
sides by p we obtain m−1
Q Qn−1
i=1 pi = u i=1 qi . By induction on m, these factorizations are the same
up to the permutation of factors and multiplications by units.
FIELDS, RINGS AND MODULES 17
where u, v ∈ A× , p1 , . . . , pn are distinct prime elements (not associate to each other) and
ki , li ≥ 0. Define
n
Y Yn
si
d= pi , si = min{ki , li }, m= ptii , ti = max{ki , li }.
i=1 i=1
We claim that d = gcd(a, b) and m = lcm(a, b). Let is show this just for d. It is clear that d | a
and d | b. Assume that c | a and c | b. Up to a unit, we can write c = ni=1 pri i , where ri ≤ ki
Q
2.5. Euclidean domains. Euclidean domains are rings where an analogue of the Euclidean
division of integers (that is, division with a remainder) is possible.
Definition 2.28. An integral domain A is called a Euclidean domain if there exists a function
δ : R\{0} → N = {0, 1, 2, . . .}
(called a Euclidean function or a degree function) such that for any a, b ∈ A\{0} there exist
elements q, r ∈ A (quotient and remainder) such that a = bq + r and either r = 0 or δ(r) < δ(b).
Example 2.29. The ring Z with the degree function δ(n) = |n| is a Euclidean domain.
Example 2.30. For any field K, the algebra of polynomials K[x] with the degree function
δ(f ) = deg(f ) is a Euclidean domain. This follows from the following result.
Theorem 2.31. Let A be a commutative ring and let f, g ∈ A[x] be polynomials such that g
is monic of degree d, meaning that g(x) = xd + gd−1 xd−1 + · · · + g0 . Then there exist unique
polynomials q, r ∈ A[x] such that f = gq + r and deg r < d.
Proof. Let us choose a polynomial q ∈ A[x] such that the polynomial r := f − gq has the
minimal possible degree. Assume that r = re xe + · · · + r0 has degree e ≥ d. Then
e−1
X d−1
X
e−d e−d i
f − g(q + re x ) = r − re gx = ri x − re gi xe−d+i
i=0 i=0
Proof. We can divide with a remainder f = (x − a)q + r, where deg r < deg(x − a) = 1. This
means that r ∈ A. If a is a root of f then r = f (a) = 0. Therefore (x − a) divides f . Conversely,
if (x − a) divides f then clearly f (a) = 0.
Remark 2.33. If A is an integral domain, then a polynomial f ∈ A[x] has at most deg f roots.
Indeed, if a ∈ A is a root of f , then we can write f = (x − a)q with deg q = deg f − 1. Any root
of f different from a is a root of q. By induction on degree, q has at most deg q = deg f − 1
roots. Therefore f has at most deg f roots. If A is not an integral domain, then f ∈ A[x] can
have more than deg f roots. For example, the polynomial x2 − 1 in Z8 [x] has four roots.
Note that if K is an infinite field and f, g ∈ K[x] are polynomials such that f (a) = g(a) for
all a ∈ K, then f − g has infinitely many roots, hence f − g = 0.
Example 2.34. The ring Z[i] = {x + yi | x, y ∈ Z} ⊆ C is called the ring of Gaussian integers.
It is a Euclidean domain with a degree function δ(x + yi) = |x + yi|2 = x2 + y 2 . Such degrees
are precisely those natural numbers that can be written as a sum of two squares.
Proof. Let a, b ∈ Z[i]\{0}. Consider the element ab−1 = x + yi ∈ C. There exist m, n ∈ Z such
that |x − m| ≤ 21 and |y − n| ≤ 21 . Let q = m + ni and
r = a − bq = b(ab−1 − q) = b((x − m) + (y − n)i).
FIELDS, RINGS AND MODULES 21
We have
δ(r) = |r|2 = |b|2 · ((x − m)2 + (y − n)2 ) ≤ |b|2 · (1/4 + 1/4) < |b|2 = δ(b).
Theorem 2.36. A Euclidean domain is a principal ideal domain and therefore also a unique
factorization domain.
Proof. Let A be a Euclidean domain with a degree function δ : A\{0} → N and let I ⊆ A be a
nonzero ideal. Let b ∈ I\{0} have a minimal possible value δ(b). Then (b) ⊆ I and we claim
that I = (b). If a ∈ I\(b), then we can write a = bq + r with r = 0 or δ(r) < δ(b). If r = 0, then
a = bq and a ∈ (b), a contradiction. If r 6= 0 then δ(r) < δ(b) and r = a − bq ∈ I, contradicting
to the minimality of δ(b). This implies that I = (b).
Example 2.37. This theorem implies in particular, that any polynomial ring K[x] over a field K
is a principal ideal domain, hence a unique factorization domain. Therefore one has GCD and
LCM in this ring. For example, the polynomials x2 + 1, x + 1 over Q have gcd = 1. On the other
hand, the same polynomials over Z2 have gcd = x + 1 (note that x2 + 1 = x2 − 1 = (x − 1)(x + 1)).
Remark 2.38 (Euclidean algorithm). As we have seen, any Euclidean domain A is a UFD and
therefore its elements have the greatest common divisors. There is an algorithm, called the
Euclidean algorithm, that allows one to find gcd(a, b) for any a, b ∈ A. It is a generalization
of a similar algorithm for integers. Namely, we apply the following sequence of division with
remainders, until we obtain the zero remainder:
a = bq1 + r1 , δ(r1 ) < δ(b),
b = r1 q2 + r2 , δ(r2 ) < δ(r1 ),
r1 = r2 q3 + r3 , δ(r3 ) < δ(r2 ),
................
rk−2 = rk−1 qk + rk , δ(rk ) < δ(rk−1 ),
rk−1 = rk qk+1 , δ(rk+1 ) = 0.
Then
gcd(a, b) = gcd(b, r1 ) = gcd(r1 , r2 ) = · · · = gcd(rk−1 , rk ) = rk ,
where we use the fact that if a = bq + r, then gcd(a, b) = gcd(b, r). This algorithm can also be
used in order to find x, y ∈ A such that xa + yb = gcd(a, b) = rk . Indeed, first we can write
r1 = a − q1 b.
Then
r2 = b − q2 r1 = b − q2 (a − q1 b) = −q2 a + (1 + q1 q2 )b.
Continuing this process, we obtain an expression of rk as a linear combination of a and b.
22 SERGEY MOZGOVOY
f f¯
K
Proof. Let as ∈ F(A) be a nonzero element. Then as 6= 01 , that is, a = 6 0. This implies that
a
s
∈ F(A). We have s · a = 1 and this means that s is invertible. Therefore F(A) is a field.
a s 1 a
2.7. Factorization in polynomial rings. We know that if K is a field, then K[x] is a UFD.
However, this is not enough to show that K[x1 , . . . , xn ] is a UFD. Our goal will be to show that
if A is a UFD, then also A[x] is a UFD. This implies that A[x1 , . . . , xn ] is a UFD by induction.
To show that A[x] is a UFD, we will consider the field of fractions K = F(A) and embed A[x]
into K[x]. Then we use the fact that K[x] is a UFD in order to prove the same for A[x].
We have seen earlier that for any two elements a, b of a UFD A, there exists their greatest
common divisor gcd(a, b). Similarly, there exists the greatest common divisor of several elements
a1 , . . . , an ∈ A which we denote by gcd(a1 , . . . , an ).
Pn i
Definition 2.42. Given a polynomial f = i=0 fi x ∈ A[x], we define its content to be
d(f ) = gcd(f0 , . . . , fn ). We say that f is primitive if d(f ) = 1 (up to a unit). Note that we can
always write f = d(f ) · f ∗ , where f ∗ ∈ A[x] is primitive.
Proof. (1) Assume that f g is not primitive. Then there exists a prime element p ∈ A such that
all coefficients of f g are divisible by p. The ring Ā = A/(p) is an integral domain. Indeed, if
a, b ∈ A are such that [a][b] = 0 in Ā, then ab ∈ (p) =⇒ p | ab =⇒ p | a or p | b =⇒ [a] = 0
or [b] = 0 in Ā. This implies that Ā[x] is also an integral domain.
For any polynomial h ∈ A[x], we denote its image in Ā[x] by h. Then f g = 0 in Ā[x]. From
f · g = f g = 0 we obtain that f = 0 or g = 0. This means that p divides all coefficients of f or
all coefficients of g, that is, either f or g is not primitive, a contradiction.
(2) Let a = d(f ) and b = d(g). Then f = af ∗ and g = bg ∗ for primitive f ∗ , g ∗ ∈ A[x]. Therefore
Remark 2.44. Given a polynomial f ∈ F(A)[x], we can write it in the form f = ab f ∗ , where
a, b ∈ A and f ∗ ∈ A[x] is primitive. If f is irreducible in F(A)[x], then f ∗ is irreducible in A[x].
Corollary 2.45. Let A be a UFD and f ∈ A[x] be a primitive non-constant polynomial. Then
f is irreducible in A[x] if and only if f is irreducible in F(A)[x].
Proof. Assume that f is irreducible in F(A)[x] and f = gh, where g, h ∈ A[x] are not invertible.
If deg g > 0 and deg h > 0, then f is also reducible in F(A), a contradiction. If deg g = 0, then
g ∈ A and g | d(f ). But d(f ) = 1, hence g is invertible in A and A[x], a contradiction. Similarly
for deg h = 0.
Assume that f is irreducible in A[x] and f = gh, where g, h ∈ F(A)[x] have positive degrees.
We can write
a c
g = g∗, h = h∗ ,
b d
∗ ∗
where a, b, c, d ∈ A and g , h ∈ A[x] are primitive. Then f = gh = ac bd
g ∗ h∗ , hence
Corollary 2.46. Let f, g, h ∈ Q[x] be monic and f = gh. If f ∈ Z[x], then g, h ∈ Z[x].
24 SERGEY MOZGOVOY
3. Fields
3.1. Basic definitions. Recall that a field is a commutative ring such that all of its nonzero
elements are invertible. We know the fields Q, R, C. We also know that, for any prime integer p,
the quotient ring Fp = Z/pZ is a field. The next lemma gives an important source of new fields.
Lemma 3.1. Let K be a field and p ∈ K[x] be irreducible. Then K[x]/(p) is a field.
Proof. Note that A = K[x] is a PID. We will prove generally that if A is a PID and p ∈ A
is irreducible, then A/(p) is a field. Let f ∈ A be such that [f ] 6= 0 in A/(p). We can write
(f, p) = (d) for some d ∈ A. Then d | p, hence d = p or d = 1 (up to a unit) as p is irreducible. If
d = p then p | f =⇒ f ∈ (p) =⇒ [f ] = 0, a contradiction. If d = 1, then there exist u, v ∈ A
such that f u + pv = 1. This implies that [f ] · [u] = 1 in A/(p). Therefore [f ] is invertible and
A/(p) is a field.
Definition 3.2. Let L be a field. A subring K ⊆ L is called a subfield if K is a field (equipped
with the induced ring structure). The field L is called a field extension of K. We write L/K in
this case.
T
Lemma 3.3. Let (Ki )i∈I be a collection of subfields of a field L. Then i∈I Ki is a subfield
of L.
T
Proof. Let K = i∈I Ki . It is clear that 0, 1 ∈ K. For any a, b ∈ K we have a, b ∈ Ki ∀i ∈ I.
Therefore a + b, ab ∈ Ki ∀i ∈ I. This implies a + b, ab ∈ K. Therefore K ⊆ L is a subring.
Finally, if a ∈ K is nonzero, then a−1 ∈ Ki ∀i ∈ I. Therefore a−1 ∈ K. This means that K ⊆ L
is a subfield.
Definition 3.4. Given a field K and a subset S ⊆ K, we define the subfield generated by S to
be the intersection of all subfields of K that contain S. It is the minimal subfield that contains S.
The subfield of K generated by ∅ (or by {0, 1}) is called the prime subfield of K. It is the
minimal subfield contained in K.
Given a field K and n ∈ Z, a ∈ K, we define
n·a=a
| + ·{z
· · + a} ∈ K
n summands
If p = 0, then f is injective and therefore n · 1K 6= 0 for all n > 0. This means that char K = 0.
The injective map f : Z → K can be extended to f : Q → K. We obtain a subfield Q ⊆ K
generated by 0, 1. This implies that Q is the prime subfield of K.
If p > 0, then p is the minimal positive integer such that p · 1 = 0 in K. Therefore char K = p.
Moreover, there is an injective ring homomorphism Z/(p) = Z/ Ker f → K. Therefore Z/(p) is
an integral domain. If p is not prime, then there exist 1 < a, b < p such that ab = p. Then the
corresponding congruence classes a, b ∈ Z/(p) are nonzero and ab = 0. This contradicts to the
fact that Z/(p) is an integral domain. We conclude that p is prime and therefore Z/(p) is a
field. It is a subfield of K generated by 0, 1. Therefore it is the prime subfield of K.
FIELDS, RINGS AND MODULES 27
(1) Among all monic polynomials in K[x] that have a root a, there exists a unique polyno-
mial p having minimal degree. It is called the minimal polynomial of a over K.
(2) The minimal polynomial p ∈ K[x] of a is irreducible. If a is a root of f ∈ K[x], then
p | f.
(3) There is an isomorphism of fields
(4) We have [K(a) : K] = deg p. If d = deg p, then (1, a, . . . , ad−1 ) is a basis of K(a) over K.
Proof. (1) Consider the ring homomorphism φa : K[x] → L given by f 7→ f (a). As a is algebraic,
the kernel Ker φa ⊆ K[x] is a non-trivial prinicipal ideal. Let p ∈ K[x] be the monic polynomial
such that Ker φa = (p). If f ∈ K[x] is another monic polynomial that has a root a, then f ∈ (p)
and p | f . This implies that deg p ≤ deg f and if deg p = deg f then p = f .
(2) We have K[x]/(p) ⊆ L, hence K[x]/(p) is an integral domain. This implies that p is
irreducible. Indeed, if p = f g then f g = 0 in K[x]/(p) =⇒ f = 0 or g = 0 =⇒ p | f or p | g
and this means that p = f or p = g up to a unit. We have seen already that if f ∈ K[x] has
root a then p | f .
(3) The kernel of the map φa : K[x] → L is (p) and its image is K[a]. This implies that
K[x]/(p) ' K[a]. We know that K[x]/(p) is a field as p is irreducible. Therefore K[a] is a field
and K(a) = K[a].
(4) The basis of K[x]/(p) is given by (1, x, . . . , xd−1 ). Therefore the basis of K[a] = K(a) is
given by (1, a, . . . , ad−1 ).
Lemma 3.15. Let L/K be a finite field extension. Then any element a ∈ L is algebraic over K.
f0 + f1 a + · · · + fn an = 0.
Pn
Let f = i=0 fi xi ∈ K[x]. Then f 6= 0 and f (a) = 0.
Example 3.16. This lemma implies that any complex number a ∈ C is algebraic over R.
Moreover, the elements 1, a, a2 are linearly dependent over R, hence there exists a nonzero
polynomial
f = f2 x2 + f1 x + f0 ∈ R[x]
such that f (a) = 0. This means that every complex number a is a root of a quadratic polynomial
with real coefficients.
Lemma 3.17. Let L/K be a field extension and a ∈ L be transcendental. Then K(a) ' K(x).
[M : K] = [M : L] · [L : K].
FIELDS, RINGS AND MODULES 29
Proof. Let r = [L : K], s = [M : L], let l1 , . . . , lr be a basis of L/K and let m1 , . . . , ms be a basis
of M/L. We claim that the products li mj form a basis of M/K. Given x ∈ M we can write
X
x= yj mj , yj ∈ L
j=1
hence is unique.
Let us prove existence of a field with q = pn elements. Let K be the splitting field of xq − x
over Fp . We claim that K is the set of q distinct roots of xq − x, hence contains q elements. Let
L ⊆ K be the set of all roots of f (x) = xq − x. Then L is a subfield of K:
FIELDS, RINGS AND MODULES 31
(1) 0, 1 ∈ L as 0q = 0 and 1q = 1.
(2) If a, b ∈ L, then aq = a, bq = b =⇒ (ab)q = aq bq = ab =⇒ ab ∈ L.
(3) If 0 6= a ∈ L, then aq = a =⇒ (a−1 )q = a−q = a−1 =⇒ a−1 ∈ L.
(4) Let a, b ∈ L. We have (a + b)q = aq + bq (binomial coefficients kq are divisible by p if
0 < k < q). This implies (a + b)q = aq + bq = a + b =⇒ a + b ∈ L.
As L is a field containing all roots of xq − x (and nothing else), it is the splitting field of this
polynomial, hence K = L. We only need to show that f (x) = xq − x does not have multiple
roots. Indeed, assume that f (x) = (x − a)2 · g(x) for some g ∈ K[x] and a ∈ K. We can formally
define a derivative of polynomials. Then f 0 (x) = 2(x − a)g(x) + (x − a)2 g 0 (x) and f 0 (a) = 0.
But f 0 (x) = qxq−1 − 1 = −1 as q = pn = 0 in K. A contradiction. We conclude that all roots of
xq − x are distinct, hence #K = #L = q.
Proof. 1 =⇒ 2. Let f ∈ K[x] have degree n and c ∈ K be its root. Then f (x) = (x − c)g(x)
for some polynomial g ∈ K[x] of degree n − 1. By induction on n, we can write g(x) =
a(x − c1 ) . . . (x − cn−1 ). Therefore
f (x) = a(x − c)(x − c1 ) . . . (x − cn−1 ).
2 =⇒ 3. Let p ∈ K[x] be an irreducible polynomial. By assumption it is a product of linear
polynomials. Therefore p is linear.
3 =⇒ 4. Let L/K be a field extension and a ∈ L be algebraic over K. Then the minimal
polynomial of a over K is irreducible. By assumption it is linear. This implies that a ∈ K.
4 =⇒ 5. Let L/K be a finite field extension. Then any element a ∈ L is algebraic. By
assumption a ∈ K. This implies that L = K.
5 =⇒ 1. We can assume that f ∈ K[x] is irreducible. Then the field K[x]/(f ) is finite over K.
By assumption L = K[x]/(f ) = K and therefore
deg f = [L : K] = 1,
that is, f is linear.
Theorem 3.26 (Fundamental Theorem of Algebra). The field C of complex numbers is alge-
braically closed.
Proof. Let
f (z) = z n + fn−1 z n−1 + · · · + f0
be a non-constant polynomial over C that does not have roots over C. There exist r > 0 such
that for all z ∈ C with |z| > r, we have
|f (z)| > |f (0)| .
Let Dr = {z ∈ C | |z| ≤ r }. Then
inf |f (z)| = inf |f (z)| = |f (z0 )| > 0
z∈C z∈Dr
32 SERGEY MOZGOVOY
for some z0 ∈ Dr (as Dr is compact). If f (z) 6= 0 for all z ∈ C, then 1/f (z) is holomorphic and
bounded over C:
1 1
|f (z)| ≥ |f (z0 )| =⇒ ≤ < +∞.
|f (z)| |f (z0 )|
By the Liouville’s Theorem such function should be constant. Therefore f (z) is constant. A
contradiction.
FIELDS, RINGS AND MODULES 33
3.4. Constructions with compass and straightedge. In this section we discuss construc-
tions on the plane R2 using a compass and a straightedge. We will usually identify R2 with C.
Our goal is to understand the set K ⊆ C of points that can be constructed using a compass and
a straightedge starting from the points 0 and 1. The following problems were studied already in
Ancient Greece:
(1) Duplication of a cube (construct a > 0 such that a3 = 2 · 13 = 2).
(2) Trisection of an arbitrary angle (given an angle φ ∈ [0, 2π] construct an angle φ/3, that
is, given a point eiφ ∈ C, construct a point eiφ/3 ).
(3) Quadrature of a circle (construct a > 0 such that a square with sides of length a has an
area of a radius one circle, that is, a2 = π)
We will show that these constructions are not possible with a compass and a straighedge.
Definition 3.27. Let K ⊆ C be a subset.
(1) A line through two distinct points in K is called an elementary K-constructible line.
(2) A circle that has some point in K and a center in K is called an elementary K-
constructible circle.
(3) All elementary K-constructible lines and circles are called elementary K-constructible
objects.
(4) A point in the intersection of two different elementary K-constructible objects is called
an elementary K-constructible point.
Lemma 3.28. Let L/K be a field extension and char K 6= 2. Then the following conditions are
equivalent
(1) [L : K] = 2.
(2) L = K[a] for some a ∈ L\K such that a2 ∈ K.
Such fields extension is called quadratic.
Proof. 1 =⇒ 2. Let b ∈ L\K. Its minimal polynomial has degree 2 and can be written in the
form
p(x) = x2 + cx + d = (x + c/2)2 + (d − c2 /4), c, d ∈ K.
From p(b) = 0 we obtain
(b + c/2)2 = c2 /4 − d.
Therefore the element a = b + c/2 ∈ L\K satisfies
a2 = c2 /4 − d ∈ K
and we have L = K[a].
2 =⇒ 1. The minimal polynomial of a over K is p(x) = x2 − a2 . Therefore
[L : K] = [K[a] : K] = deg p = 2.
Theorem 3.29. The following subsets of C coincide
(1) The smallest subset K ⊆ C that contains 0, 1 and all its elementary constructible points.
(2) The minimal subfield L ⊆ C that contains all its square roots.
Proof. L ⊆ K: To prove this we need to show that K is a field closed under taking square roots.
Then L ⊆ K as L is the minimal field with this property. Let us show first that K is a field.
Given z, w ∈ C, one can construct a parallelogram with vertices 0, z, w, z + w. Therefore one can
construct z + w. It is easy to construct −z. To construct zw or z/w, we represent z = aeiφ and
34 SERGEY MOZGOVOY
w = aeiψ , where a, b ∈ R are the lengths of z, w and φ, ψ ∈ [0, 2π] are the angles between z, w
and the x-axis. Then zw = abei(φ+ψ) and similarly for z/w. One can easily add and subtract
angles. Therefore we just have to multiply and divide real positive numbers. In the following
picture the lines AC and BD are parallel and therefore a1 = cb , that is, c = ab. This implies that
if we know real a, b > 0, then we can construct c = ab. And if we know real b, c > 0, then we
can construct a = cb . This implies that K is a field.
C
b
O 1 A a B
Let us show that K is stable under taking square roots. We have to prove that given z = aeiφ ,
√
we can also construct aeiφ/2 . It is easy to construct a bisector of an angle. Therefore we can
√
construct an angle φ/2. To construct a, consider the following picture, where we start with
intervals OA and AB, construct a circle with the diameter OB, and raise a perpendicular to
OB at the point A. The angle OCB is right.
C
O 1 A a B
We have
(12 + h2 ) + (a2 + h2 ) = OC 2 + BC 2 = OB 2 = (1 + a)2
√
which implies h2 = a, that is, h = a.
K ⊆ L: To prove this we will show that L 3 0, 1 is closed under elementary constructions.
Then K ⊆ L as K is the minimal set with this property. First, we claim that z ∈ L if and only
if <z, Im z ∈ L ∩ R. If L ⊆ C is closed under taking square roots, then so is its conjugate L and
the intersection L ∩ L (if x2 = a for a ∈ L ∩ L then x ∈ L ∩ L). From the minimality of L, we
obtain L = L ∩ L =⇒ L = L. This implies that if z = x + iy ∈ L =⇒ z = x − iy ∈ L =⇒
x = 12 (z + z) ∈ L and iy = 12 (z − z) ∈ L. Note that i ∈ L as i2 = −1 ∈ L. Therefore y ∈ L.
Conversely, if x, y ∈ L ∩ R then also x + iy ∈ L.
An elementary L-constructible circle consists of points x + iy satisfying
(x − a)2 + (y − b)2 = c2
for some a, b, c ∈ L ∩ R. An elementary L-constructible line through the points x1 + iy1 ∈ L
and x2 + iy2 ∈ L (with x1 6= x2 , y1 6= y2 ) has an equation
x − x1 y − y1
= .
x2 − x1 y2 − y1
It can be written in the form
ax + by + c = 0,
FIELDS, RINGS AND MODULES 35
This implies that Kn is a finite field extension of Q. Therefore a is algebraic over Q. If p ∈ Q[x]
is the minimal polynomial of a, then
deg p = [Q[a] : Q]
is a divisor of [Kn : Q] = 2n as
[Kn : Q] = [Kn : Q[a]] · [Q[a] : Q].
Therefore deg p is a power of 2.
Corollary 3.32. The following constructions are impossible with a compass and a straightedge
(1) Duplication of a cube.
(2) Trisection of an arbitrary angle.
(3) Quadrature of a circle.
Proof. 1. Assume that we can construct a > 0 such that a3 = 2. The minimal polynomial
of a = 21/3 over Q is x3 − 2. Its degree is 3 which is not a power of 2. Therefore a is not
constructible.
2. Assume that we can trisect an arbitrary angle. One can construct angles π/3 and φ = 2π/3.
Therefore the point eiφ = e2πi/3 is constructible. We will show that the angle φ/3 is not
constructible, that is, the point z = eiφ/3 = e2πi/9 is not constructible. We have
z 9 = e9·2πi/9 = e2πi = 1.
Therefore z is a root of the polynomial
x9 − 1 = (x3 − 1)(x6 + x3 + 1).
It is clear that z 3 6= 1. Therefore z is a root of the polynomial p(x) = x6 + x3 + 1. This
polynomial is irreducible. To see this we apply the Eisenstein’s criterion to the polynomial
p(x + 1) = (x + 1)6 + (x + 1)3 + 1 = x6 + 6x5 + 15x4 + 21x3 + 18x2 + 9x + 3
with prime 3. The irreducibility of p implies that p is a minimal polynomial of z. Its degree is 6
which is not a power of 2. Therefore z is not constructible.
36 SERGEY MOZGOVOY
3. If the quadrature of the circle is possible, then the value a > 0 satisfying a2 = π is
algebraic. This implies that a2 is also algebraic. But by the theorem of Lindemann (1882) π is
not algebraic.
FIELDS, RINGS AND MODULES 37
4. Symmetric polynomials
Let Kbe a field. For every permutation σ ∈ Sn , define a map
Λn = K[x1 , . . . , xn ]Sn
Example 4.1.
(1) For every 1 ≤ k ≤ n, the polynomial
X
ek = xi1 . . . xik
1≤i1 <···<ik ≤n
e2 = x1 x2 + x1 x3 + x2 x3 + · · · + xn−1 xn ,
en = x1 . . . xn .
We can write
n
Y
(x − xi ) = xn − e1 xn−1 + e2 xn−2 + · · · + (−1n )en .
i=1
pk = xk1 + · · · + xkn
Theorem 4.2. The elements e1 , . . . , en generate Λn over Kand are algebraically independent.
This means that every element in Λn is a polynomial in e1 , . . . en and if
X
f (e1 , . . . , en ) = fi1 ,...,in ei11 . . . einn = 0, fi1 ,...,in ∈ K,
i1 ,...,in ≥0
then f = 0.
38 SERGEY MOZGOVOY
which is symmetric in x1 , . . . , xn . Note that ∆ is zero if and only if f has a multiple root. Let
us show that ∆ ∈ K. We can assume that a = 1. By the previous results ∆ is a polynomial in
e1 , . . . , en , where
f (x) = xn − e1 xn−1 + e2 xn−2 + · · · + (−1)n en .
This means that ∆ is a polynomial in the coefficients of f , hence ∆ ∈ K.
Let n = 2 and f = x2 − e1 x + e2 = x2 + bx + c. Then we get the usual discriminant
∆ = (x1 − x2 )2 = e21 − 4e2 = b2 − 4c.
Let n = 3 and assume for simplicity that
f = x3 − e1 x2 + e2 x − e3 = x3 + bx + c,
implying that e1 = x1 + x2 + x3 = 0. As ∆ has degree 6, we have
∆ = ue23 + ve32 ,
for some u, v ∈ Z.
(1) If x1 = 1 = −x2 and x3 = 0, then ∆ = 4 = ve32 = −v, hence v = −4.
(2) If x1 = x2 = 1 and x3 = −2, then ∆ = 0 = 4u − 4(1 − 2 − 2)3 , hence u = −27.
We conclude that
∆ = −27e23 − 4e32 = −27c2 − 4b3 .
Example 4.6. Let us compute the discriminant of xn − 1. Let ξ = e2πi/n . Then
n−1
n n
(ξ − ξ ) = (−1)( 2 ) (ξ − ξ ) = (−1)( 2 )
Y Y YY
i j 2 i j
∆= ξ i (1 − ξ j−i )
0≤i<j<n i6=j i=0 j6=i
n−1
n n n n
= (−1)( 2 ) (1 − ξ k ) = (−1)( 2 ) ξ ( 2 ) nn = (−1)( 2 ) eπi(n−1) nn = ±nn .
Y Y
ξi
i=0 k6=0
−1 n
where we used the fact that k6=0 (x − ξ k ) = xx−1 = 1 + x + · · · + xn−1 and substituted x = 1.
Q
5. Modules
5.1. Definition and examples. Modules over rings generalize the notion of a vector space
over a field.
Definition 5.1. A module over a ring A (or an A-module) is an abelian group (M, +) together
with a map
A × M → M, (a, m) 7→ a · m = am,
called a multiplication, such that for all a, b ∈ A and m, n ∈ M
(1) a(bm) = (ab)m,
(2) 1m = m,
(3) a(m + n) = am + an,
(4) (a + b)m = am + bm.
Remark 5.2. If A is a field, then an A-module is usually called an A-vector space or a vector
space over A.
Example 5.3. Any ring A is a module over itself. More generally, for any n ≥ 1, the product
An is an A-module with the multiplication defined by a(x1 , . . . , xn ) = (ax1 , . . . , axn ).
Example 5.4. Let I ⊆ A be an ideal. Then I is a module over A with the multiplication
A × I 3 (a, b) 7→ ab ∈ I.
a · m = f (a) · m, a ∈ A, m ∈ M.
a · b = f (a) · b, a ∈ A, b ∈ B.
Remark 5.6. Let M be an A-module, 0A be the zero element of A and 0M be the zero element
of M . Then
(1) 0A m = 0M for any m ∈ M .
(2) (−1)m = −m for any m ∈ M .
(3) a0M = 0M for any a ∈ A.
Indeed,
0A m + 0A m = (0A + 0A )m = 0A m.
Therefore 0A m = 0M . Similarly,
Therefore a0M = 0M .
FIELDS, RINGS AND MODULES 41
Example 5.7. Let M be an abelian group. Then M is automatically a Z-module. Indeed, for
any k ≥ 0 and m ∈ M , we define
k·m=m
| + ·{z
· · + m}
k summands
and (−k)m = −km. Note that this is a unique possible structure of a Z-module on M because
we should have 1 · m = m, 2 · m = (1 + 1)m = m + m and generally (k + 1)m = km + m.
42 SERGEY MOZGOVOY
The map
Mn (K) → EndK (V ), A 7→ fA
is an isomorphism of rings. Given a linear map f : V → V , one reconstructs the matrix
P
A = (aij ) ∈ Mn (K) by the rule f ej = i aij ei , where (e1 , . . . , en ) is the standard basis of
V = K n . We will often identify A ∈ Mn (K) and fA ∈ EndK (V ).
Remark 5.11. If M is an abelian group, then the ring EndZ (M ) is denoted by End(M ). If M
is an A-module, then there is a ring homomorphism
φ : R → End(M ), φ(a)(m) = am, a ∈ A, m ∈ M.
For example
φ(ab)m = (ab)m = a(bm) = φ(a)(bm) = φ(a)(φ(b)m)
and therefore φ(ab) = φ(a) ◦ φ(b). Conversely, given an abelian group M and a ring homomor-
phism φ : R → End(M ), we can equip M with an A-module structure
a · m = φ(a)(m) ∈ M, a ∈ A, m ∈ M.
Example 5.12. Let V be a vector space over a field K. Given a K-linear map A ∈ EndK (V ),
the evaluation map
K[x] → EndK (V ), f 7→ f (A)
is a ring homomorphism, hence V gets a structure of a K[x]-module
f · v = f (A)(v), f ∈ K[x], v ∈ V.
Conversely, if V is a K[x]-module (extending the K-vector space structure on V ), then multipli-
cation by x induces a K-linear map A : V → V .
FIELDS, RINGS AND MODULES 43
f f¯
0
M
∼
The map f¯ induces an isomorphism f¯: M/ Ker f −
→ Im f .
44 SERGEY MOZGOVOY
Remark 5.20. Given two A-modules N, N 0 , define a new A-module N ⊕N 0 , called an (external)
direct sum of N and N 0 , to be the product of abelian groups N ×N 0 equipped with a multiplication
a(n, n0 ) = (an, an0 ), a ∈ A, n ∈ N, n0 ∈ N 0 .
A module M is indecomposable if and only if it is not isomorphic to a direct sum N ⊕ N 0 for
some nonzero modules N, N 0 .
Lemma 5.23 (Schur’s lemma). For any simple A-module M , the endomorphism ring EndA (M )
is a division ring, that is, any endomorphism M → M is either zero or invertible.
5.4. Chinese remainder theorem. We have seen that given a PID A and coprime elements
p, q ∈ A, we have
A/(pq) ' A/(p) ⊕ A/(q).
We will generalize this statement by proving the Chinese Remainder Theorem. In its classical
form, formulated in the 3rd century AD by Sunzi (not to be confused with Sun Tzu, the author
of “The art of war”) it is
Theorem 5.30. Let n1 , . . . , nk be positive, pairwise coprime integers. Then for any integers
a1 , . . . , ak , there exists an integer a such that
a ≡ ai (mod ni ), ∀i = 1, . . . , k.
There exists a unique such integer with 0 ≤ a < n = n1 . . . nk .
This theorem can be also formulated as a statement that the map
Z/nZ → Z/n1 Z × . . . × Z/nk Z, a + nZ 7→ (a + n1 Z, . . . , a + nk Z)
is an isomorphism (of rings or of abelian groups).
Proof. The above map is injective. Indeed, if [a] = a + nZ is mapped to zero, then ni | a for
Q
all i, hence n = i ni | a as the elements ni are pairwise coprime. This implies that [a] = 0.
Injectivity of the map implies bijectivity as the groups on both sides have the same number of
elements.
We can generalize the above result to arbitrary PID.
Theorem 5.31. Let A be a PID and n1 , . . . , nk be pairwise coprime elements (this means that
gcd(ni , nj ) = 1 for i 6= j). Then the map
A/(n) → A/(n1 ) × . . . × A/(nk ), a + (n) 7→ (a + (n1 ), . . . , a + (nk )),
is an isomorphism of rings (or A-modules), where n = n1 . . . nk .
Remark 5.32. In particular, we see that for any coprime p, q ∈ A, we have an isomorphism
A/(pq) ' A/(p) ⊕ A/(q) of A-modules.
This statement, in its own right, can be generalized as follows
Theorem 5.33. Let A be a commutative ring and I1 , . . . , Ik be ideals of A that are pairwise
coprime: Ii + Ij = A, i 6= j. Then the map
φ : A/I → A/I1 × . . . × A/Ik , a + I 7→ (a + I1 , . . . , a + Ik ),
T
is an isomorphism of rings (or A-modules), where I = i Ii . Moreover I = I1 I2 . . . Ik .
Proof. The kernel of the map
A → A/I1 × . . . × A/Ik
T
is i Ii = I. This implies injectivity of φ. Let us prove surjectivity. For any i 6= j, we can find
eij ∈ Ii and eji ∈ Ij such that 1 = eij + eji . Then, for any i, we have
Y Y
1= (eij + eji ) ∈ Ii + eji .
j6=i j6=i
Q Q
Let ei ∈ Ii and fi = j6=i eji ∈ j6=i Ij be such that 1 = ei + fi . Given elements [ai ] ∈ A/Ii for
P
all i, we claim that a = j fj aj satisfies a ≡ ai (mod Ii ) for all i. Indeed,
X
a − ai = a − (ei + fi )ai = fj aj − ei ai ∈ Ii
j6=i
48 SERGEY MOZGOVOY
as ei ∈ Ii and fj ∈ Ii for j 6= i. Therefore a ≡ ai (mod Ii ) for all i and the map φ is surjective.
T Q Q T
We also have to show that i Ii = i Ii . Inclusion i Ii ⊆ i Ii is trivial. On the other hand,
T
let a ∈ i Ii . Then by induction a ∈ J = I1 . . . Ik−1 . Therefore
Y
a = a(ek + fk ) ∈ JIk + Ik J = Ii
i
Q T Q
as ek ∈ Ik , fk ∈ j6=k Ij = J and a ∈ Ik . This proves that i Ii ⊆ i Ii .
FIELDS, RINGS AND MODULES 49
5.5. Modules over PID. Consider the following two fundamental results of linear algebra and
the theory of finite abelian groups. In linear algebra one proves that every square matrix over C
is conjugate to its Jordan canonical form which is a direct sum of Jordan blocks
λ 1 0 ...... 0
0 λ 1 . . . . . . 0
0 0 λ . . . . . . 0
Jn,λ = λ ∈ C, n ≥ 1.
. . . . . . . . . . . . . . . . . . .
0 0 0 . . . λ 1
0 0 0 ... 0 λ
The fundamental theorem of finitely generated abelian groups states that every such group is
isomorphic to a direct sum
Zk ⊕ Z/(pn1 1 ) ⊕ · · · ⊕ Z/(pnr r ),
where k ≥ 0, pi ∈ Z are prime numbers and ni ≥ 1. In particular, every finite abelian group is
isomorphic to
Z/(pn1 1 ) ⊕ · · · ⊕ Z/(pnr r ).
We will see that these two results are essentially equivalent, if seen from an appropriate point of
view, and then we will give a general unified proof.
Given a vector space V = K n over a field K and a linear operator A ∈ EndK (V ) = Mn (K),
we can equip V with a structure of a K[x]-module (we denote it by VA )
f · v = f (A)(v), f ∈ K[x], v ∈ V.
Conversely, if V is a K[x]-module, then it is a K-vector space and we can define
A ∈ EndK (V ), A(v) = x · v, v ∈ V.
Example 5.34. Consider the K[x]-module V = K[x]/(x − λ)n , where λ ∈ K and n ≥ 1.
Choose the basis (x − λ)n−1 , . . . , (x − λ)2 , (x − λ), 1 of V . In this basis we have
A(x − λ)k = x(x − λ)k = (x − λ)k+1 + λ(x − λ)k
for k < n − 1 and A(x − λ)n−1 ≡ λ(x − λ)n−1 (mod (x − λ)n ). Therefore the matrix of A in
this basis is exactly the Jordan block Jn,λ .
Let W be another K-vector space, B ∈ EndK (W ) and WB be the corresponding K[x]-
module. Then an isomorphism g : VA → WB of K[x]-modules can be identified with a K-linear
isomorphism g : V → W such that gA = Bg:
gA(v) = g(x · v) = x · g(v) = Bg(v), v ∈ V.
This means that B = gAg −1 and the corresponding matrices are conjugate if V = W = K n .
The statement that a matrix A ∈ Mn (K) (with K = C) is conjugate to a direct sum of Jordan
blocks Jn1 ,λ1 , . . . , Jnr ,λr can be translated now to the statement that the K[x]-module VA is
isomorphic to a direct sum of K[x]-modules
K[x]/(x − λ1 )n1 ⊕ · · · ⊕ K[x]/(x − λr )nr .
The polynomials (x − λ)n are powers of irreducible polynomials x − λ ∈ K[x] and these are the
only irreducible (or prime) elements in K[x] (up to a unit) if K = C or K is algebraically closed.
In the same way for abelian groups we had summands Z/(pn ), where p is prime. This shows
that both statements are essentially equivalent, with the first statement being about modules
over K[x] and the second statement about modules over Z (that is, abelian groups). Both of
them follow from the general result we will prove next.
50 SERGEY MOZGOVOY
Definition 5.35. A module M over a ring A is called finitely generated if there exists a finite
family (m1 , . . . , mk ) of elements in M (called generators of M ) such that every element of M
can be written in the form ki=1 ai mi for some ai ∈ A.
P
Theorem 5.36. Let A be a PID. Then evry finitely generated A-module is isomorphic to a
direct sum
Ak ⊕ A/(pn1 1 ) ⊕ · · · ⊕ A/(pnr r ),
where k ≥ 0, pi ∈ A are prime and ni ≥ 1. The modules A and A/(pn ) are indecomposable.
Proof. Let M be a finitely generated A-module with generators x1 , . . . , xm . Then there is a
surjective homomorphism ψ : Am → M (where Am 3 ei 7→ xi ). The kernel Ker ψ ⊆ Am is
finitely generated (see below). Therefore there is a surjective homomorphism φ : An → Ker ψ
and we obtain a sequence of maps
φ ψ
An →
− Am − → M,
where Im φ = Ker ψ so that
Coker φ := Am / Im φ = Am / Ker ψ ' M.
We can represent φ as an m × n matrix with entries in A. By changing the bases of Am and An
we can put this matrix to the form, called a Smith normal form,
a1 0 0 . . . . . 0
0 a2 0 . . . . . 0
0 0 . . . . . . . . 0
. . . . . . . . . . ar . . . .
0 0 . . . . . . 0 0
0 0 ...... 0 0
with nonzero a1 | a2 | · · · | ar and r ≤ m, n. The elements ai are unique (up to a unit). The
algorithm is somewhat involved for general PID, but in the case of Euclidean domains one can
use the usual Gaussian elimination on rows and columns to get the required form. I omit the
details. We obtain then
M ' Am / Im φ = A/(a1 ) ⊕ · · · ⊕ A/(ar ) ⊕ Am−r .
If a ∈ A is nonzero and a = i pki i is a factorization into irreducible (distinct) factors, then
Q
5.6. Noetherian modules. In the proof of the previous theorem we used the fact that a
submodule of An is finitely generated if A is a PID. Let us discuss this in more detail.
Definition 5.37. Let A be a ring.
(1) An A-module M is called Noetherian if every submodule of M is finitely generated.
(2) The ring A is called Noetherian if it is Noetherian as a (left) A-module over itself.
Example 5.38. A PID A is Noetherian. Indeed, every submodule of A is an ideal, hence a
principal ideal, generated by one element.
Lemma 5.39. Let A be a ring and M be an A-module. The following conditions are equivalent
(1) Every submodule of M is finitely generated.
(2) Every increasing chain of submodules
M1 ⊆ M2 ⊆ . . . ⊆ M
stabilizes, that is, Mn = Mn+1 = . . . for n 0.
Proof. 1 =⇒ 2. Consider an increasing chain
M1 ⊆ M2 ⊆ . . . ⊆ M
and let N = ∪n≥1 Mn ⊆ M . Then N is a submodule of M and by assumption it is finitely
generates. Let x1 , . . . , xk be generators of N . Then xi ∈ Mni for some ni ≥ 1. Taking
n = maxi ni , we obtain xi ∈ Mn for all i, hence N ⊆ Mn and Mn = Mn+1 = . . . .
2 =⇒ 1. Let N ⊆ M be a submodule and let x0 = 0 ∈ N . Assuming that elements x0 , . . . , xk
in N are constructed, let Mk ⊆ N be the module generated by them. If Mk = N then N is
finitely generated and we are done. If Mk 6= N , let xk+1 ∈ N \Mk and continue the procedure.
In this way we obtain a chain of modules
M1 ⊆ M2 ⊆ . . . ⊆ N ⊆ M
with Mk 6= Mk+1 for all k ≥ 1. A contradiction.
Lemma 5.40. Let M be an A-module and N ⊆ M be a submodule. Then M is Noetherian
⇐⇒ N and M/N are Noetherian.
Proof. Let M be Noetherian. If L ⊆ N is a submodule, then L ⊆ M , hence L is finitely
generated and N is Noetherian. Let L ⊆ M/N be a submodule and let π : M → M/N be the
projection. The module L0 = π −1 (L) ⊆ M is finitely generated, hence also L = π(L0 ) is finitely
generated and M/N is Noetherian.
Assume that N and M/N are Noetherian and let L ⊆ M . Then L ∩ N ⊆ N is finitely
generated and L/(L ∩ N ) ' (L + N )/N ⊆ M/N is finitely generated. This implies that L is
also finitely generated.
Corollary 5.41. If M, N are Noetherian A-modules, then M ⊕ N is also Noetherian.
Proof. Let M 0 = M ⊕ N . Then N ⊆ M 0 and M 0 /N ' M are Noetherian. We conclude that
M 0 is Noetherian.
Corollary 5.42. If A is a Noetherian ring, then the module An is Noetherian for all n ≥ 1.
Proof. The module An is a direct sum of copies of A.
Corollary 5.43. If A is a PID, then every submodule of An is finitely generated.
Proof. We have seen that A is Noetherian. Therefore An is also Noetherian. This means that
every submodule of An is finitely generated.
52 SERGEY MOZGOVOY
Remark A.2. An element ξ = e2πik/n is an n-th primitive root of unity if and only if gcd(k, n) =
1. Indeed, if d = gcd(k, n) > 1, then ξ n/d = e2πik/d = 1 as k/d ∈ Z. This implies that ord ξ < n.
Conversely, if gcd(k, n) = 1 and ξ m = 1 for some 1 ≤ m < n, then n | km =⇒ n | m, a
contradiction.
Remark A.3. Let ξ be a d-th primitive root of unity. Then ξ n = 1 if and only if d | n. Indeed,
ξ = e2πik/d with gcd(k, d) = 1. Therefore 1 = ξ n = e2πikn/d =⇒ d | kn =⇒ d | n. Every n-th
root of unity is a primitive root for a unique d | n.
Let Y
Φd (x) = (x − ξ).
ord ξ=d
We obtain by induction (and Gauss lemma) that Φd (x) ∈ Z[x]. This polynomial is called a d-th
cyclotomic polynomial.
Example A.4. We have Φ1 (x) = x − 1. For any prime p, we have d | p if and only if d = 1 or
d = p. This implies
xp − 1 = (x − 1)Φp (x),
hence
Φp (x) = xp−1 + · · · + x + 1.
We have
x4 − 1 = Φ1 (x)Φ2 (x)Φ4 (x),
hence Φ4 (x) = x2 + 1.
x6 − 1 = Φ1 (x)Φ2 (x)Φ3 (x)Φ6 (x),
hence
(x3 − 1)(x3 + 1) x3 + 1
Φ6 (x) = = = x2 − x + 1.
(x3 − 1)Φ2 (x) x+1
Theorem A.5 (Kronecker). The polynomial Φd (x) is irreducible for every d ≥ 1.
d
X d
X
p
f (x) ≡ fip xip ≡ fi xip = f (xp ) (mod p),
i=0 i=0
p
where we used a ≡ a (mod p) for a ∈ Z (Fermat’s little theorem).
By assumption (p, n) = 1 =⇒ ξ p is a primitive n-th root of 1 =⇒ Φn (ξ p ) = 0. Let g | Φn
be an irreducible polynomial such that g(ξ p ) = 0. If f = g, then we are done, hence we assume
that f =6 g. As f and g are irreducible and divide Φn (and xn − 1), we obtain that f g | xn − 1.
As ξ is a root of g(xp ), we obtain that f (x) divides g(xp ). Therefore, modulo p, f¯(x) ∈ Fp [x]
divides ḡ(xp ) = ḡ(x)p . Hence f¯(x), ḡ(x) have a non-trivial common factor h̄(x). This implies
that h̄2 divides f¯ḡ and xn − 1 in Fp [x]. But xn − 1 ∈ Fp [x] does not have multiple factors (if
a polynomial has a multiple factor, then this polynomial and its derivative have a non-trivial
common factor; however xn − 1 and its derivative nxn−1 are coprime in Fp [x]). A contradiction.
FIELDS, RINGS AND MODULES 55
where the product runs over all prime divisors of n (φ is called Euler’s totient function). Then,
for every a ∈ Z coprime with n, we have
aφ(n) ≡ 1 (mod n).
Proof. Let Z× n denote the multiplicative group of invertible elements in Zn . Let us show that
Qr ki
the number of elements in Z×n equals φ(n). If n = p
i=1 i (where pi are prime and pi 6= pj for
Qr
i 6= j), then Zn ' i=1 Zpki by the Chinese remainder theorem. An element in Zn is invertible
i
if and only if the corresponding components in Zpki are invertible. The only non-invertible
i
elements in Zpk (where p is prime) are multiples of p and there are pk−1 of them. Therefore
the number of invertible elements in Zpk equals pk − pk−1 . This implies that the number of
invertible elements in Zn equals
Y r r
Y Y
ki ki −1
(p − p )= pki (1 − 1/pi ) = n (1 − 1/p) = φ(n).
i=1 i=1 p|n