Ireland Rosen Chapters 4 5 6 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

Solutions to Ireland, Rosen “A Classical Introduction to

Modern Number Theory”


Richard Ganaye

August 21, 2022

Chapter 4

Ex. 4.1 Show that 2 is a primitive root modulo 29.

Proof. Let p = 29 : p − 1 = 22 × 7.
24 = 16 6≡ 1[29]
214 = 47 = 4 × 163 = 64 × 256 ≡ 6 × (−34) = −204 ≡ 86 = 3 × 29 − 1 ≡ −1[29]
228 ≡ 1[29] and 2d 6≡ 1 if d | 28, d < 28, hence 2 is a primitive element modulo 29.

Ex. 4.2 Compute all primitive roots for p = 11, 13, 17, and 19.

Proof. • p = 11. Then p − 1 = 10 = 2 × 5.


22 = 4 6≡ 1 (mod 11), and 25 = 32 ≡ −1 6≡ 1 (mod 11), so 2 is a primitive element
modulo 11.
The other primitive elements modulo 11 are congruent to the powers 2i , i ∧ 10 =
1, 1 ≤ i < 10, namely 2, 23 , 27 , 29 .
27 ≡ 7 (mod 11), 29 ≡ 6 (mod 11), so
{2, 8, 7, 6} is the set of the generators of U (Z/11Z).
Similarly :
• p = 13 : {2, 6, 11, 7} is the set of the generators of U (Z/13Z).
• p = 17 : {3, 10, 5, 11, 14, 7, 12, 6} is the set of the generators of U (Z/17Z).
• p = 19 : {2, 13, 14, 15, 3, 10} is the set of the generators of U (Z/19Z).
I obtain these results with the direct orders in S.A.G.E. :
p = 19; Fp = GF(p); a = Fp.multiplicative_generator()
print([a^k for k in range(1,p) if gcd(k,p-1) == 1])

Ex. 4.3 Suppose that a is a primitive root modulo pn , p an odd prime. Show that a is
a primitive root modulo p.

Proof. Suppose that a is a primitive root modulo pn : then a is a generator of U (Z/pn Z).
If a was not a primitive root modulo p, a is not a generator of U (Z/pZ), so there
exists b ∈ Z, b∧p = 1 such that ak 6≡ b (mod p) for all k ∈ Z. A fortiori ak 6≡ b (mod pn ),
and b ∧ pn = 1, so b ∈ U (Z/pn Z) and b 6∈ hai in U (Z/pn Z), in contradiction with the
hypothesis. So a is a primitive root modulo p.
(the reasoning on the orders of a, modulo p and modulo pn , is possible, but not so
easy.)

1
Ex. 4.4 Consider a prime p of the form 4t + 1. Show that a is a primitive root modulo
p iff −a is a primitive root modulo p.

Proof. Solution 1.
Suppose that a is a primitive root modulo p. As p − 1 is even, (−a)p−1 = ap−1 ≡ 1
(mod p).
If (−a)n ≡ 1 (mod p), with n ∈ N, then an ≡ (−1)n (mod p).
Therefore a2n ≡ 1 (mod p). As a is a primitive root modulo p, p − 1 | 2n, 2t | n, so
n is even.
Hence an ≡ 1 (mod p), and p − 1 | n. So the least n ∈ N∗ such that (−a)n ≡ 1
(mod p) is p − 1 : the order of −a modulo p is p − 1, −a is a primitive root modulo p.
Conversely, if −a is a primitive root modulo p, we apply the previous result at −a to
to obtain that −(−a) = a is a primitive root.

Solution 2.
Let p − 1 = 2a0 pa11 · · · pakk the decomposition of p − 1 in prime factors.
As pi is odd for i = 1, 2, · · · k, (p − 1)/pi is even, and a is primitive, so

(−a)(p−1)/pi = a(p−1)/pi 6≡ 1 (mod p),


(−a)(p−1)/2 = (−a)2k = a2k = a(p−1)/2 6≡ 1 (mod p).

So the order of a is p − 1 modulo p (see Ex. 4.8) : a is a primitive element modulo p.

Ex. 4.5 Consider a prime p of the form 4t + 3. Show that a is a primitive root modulo
p iff −a has order (p − 1)/2.

Proof. Let a a primitive root modulo p.


As ap−1 ≡ 1 (mod p), p | (a(p−1)/2 − 1)(a(p−1)/2 + 1), so p | a(p−1)/2 − 1 or p |
a (p−1)/2 + 1. As a is a primitive root modulo p, a(p−1)/2 6≡ 1 (mod p), so

a(p−1)/2 ≡ −1 (mod p).

Hence (−a)(p−1)/2 = (−1)2t+1 a(p−1)/2 ≡ (−1) × (−1) = 1 (mod p).


Suppose that (−a)n ≡ 1 (mod p), with n ∈ N.
Then a2n = (−a)2n ≡ 1 (mod p), so p − 1 | 2n, p−1
2 | n.
So −a has order (p − 1)/2 modulo p.

Conversely, suppose that −a has order (p − 1)/2 = 2t + 1 modulo p. Let 2, p1 , . . . pk


the prime factors of p − 1, where pi are odd.
a(p−1)/2 = a2t+1 = −(−a)2t+1 = −(−a)(p−1)/2 ≡ −1, so a(p−1)/2 6≡ 1 (mod 2).
As p − 1 is even, (p − 1)/pi is even, so
a(p−1)/pi = (−a)(p−1)/pi 6≡ 1 (mod p) (since −a has order p − 1).
So the order of a is p − 1 (see Ex. 4.8) : a is a primitive root modulo p.

n
Ex. 4.6 If p = 22 + 1 is a Fermat prime, show that 3 is a primitive root modulo p.

Proof. Solution 1 (with quadratic reciprocity).


Write p = 2k + 1, with k = 2n .
We suppose that n > 0, so k ≥ 2, p ≥ 5. As p is prime, 3p−1 ≡ 1 (mod p).
k
In other words, 32 ≡ 1 (mod p) : the order of 3 is a divisor of 2k , a power of 2.

2
k−1
 k−1 2
3 has order 2k modulo p iff 32 6≡ 1 (mod p). As 32 ≡ 1 (mod p), where p is
k−1
prime, this is equivalent to 32 ≡ −1 (mod p), which remains to prove.
2k−1 (p−1)/2 3

3 =3 ≡ p (mod p).
As the result is true for p = 5, we can suppose n ≥ 2. From the law of quadratic
reciprocity :   
3 p k−1
= (−1)(p−1)/2 = (−1)2 = 1.
p 3
So p3 = p3
 

n n
p = 22 + 1 ≡ (−1)2 + 1 (mod 3)
≡ 2 ≡ −1 (mod 3),
3
 p
so p = 3 = −1, that is to say
k−1
32 ≡ −1 (mod p).
n n
The order of 3 modulo p = 22 + 1 is p − 1 = 22 : 3 is a primitive root modulo p.
(On the other hand, if 3 is of order p − 1 modulo p, then p is prime, so
n 2n −1
Fn = 22 + 1 is prime ⇐⇒ 3(Fn −1)/2 = 32 ≡ −1 (mod Fn ).)

Solution 2 (without quadratic reciprocity, with the hint of chapter 4).


n−1
As above, if we suppose that 3 is not a primitive root modulo p, then 32 ≡ 1
(p−1)/2 2n−1
(mod p), so n ≥ 2, and (−3) =3 ≡ 1 (mod p), so −3 is a square modulo p :
there exists a ∈ Z such that −3 ≡ a2 (mod p).
As 2 ∧ √
p = 1, there exists u ∈ Z such that 2u ≡ −1 + a (mod p) (u is similar to
ω = −1+i2
3
∈ C). Then

8u3 ≡ (−1 + a)3


≡ −1 + 3a − 3a2 + a3
≡ −1 + 3a + 9 − 3a
≡8 (mod p)

As p ∧ 2 = p ∧ 8 = 1, u3 ≡ 1 (mod p). Moreover, if u ≡ 1 (mod 3), then a ≡ 3 (mod p),


−3 ≡ 9 (mod p), p | 12, so p = 2 or p = 3, in contradiction with p ≥ 5. So the order of u
modulo p is 3 : (Z/pZ)∗ contains an element u of order 3. So 3 | p − 1, p ≡ 1 (mod 3),
n
but p ≡ (−1)2 + 1 ≡ 2 ≡ −1 (mod 3) : this is a contradiction, so 3 is a primitive root
n
modulo p = 22 + 1.

Ex. 4.7 Suppose that p is a prime of the form 8t + 3 and that q = (p − 1)/2 is also a
prime. Show that 2 is a primitive root modulo p.

Proof. The first examples of such couples (q, p) are (5, 11), (29, 59), (41, 83), (53, 107), (89, 179).
p = 2q + 1 = 8t + 3 and p, q are prime numbers.
From Fermat’s little theorem, 2p−1 ≡ 1 (mod p), so 22q ≡ 1 (mod p).
The order of 2 modulo p divides 2q : to prove that the order of 2 is 2q = p − 1, it is
suffisant to prove
22 6≡ 1 (mod p), 2q 6≡ 1 (mod p).

3
If 22 ≡ 1 (mod p), then p | 3, p = 3 and q = 1 : q is not a prime, so 22 6≡ 1 (mod p).
If 2q = 2(p−1)/2 ≡ 1 (mod p), then 2 is a square modulo p (prop. 4.2.1) : there exists
a ∈ Z such that 2 ≡ a2 (mod p).
From the complementary case of the law of quadratic reciprocity (see next chapter,
prop. 5.1.3), 2 is a square modulo p iff
 
2 2
1= = (−1)(p −1)/8 .
p
2
Yet p ≡ 3 (mod 8), so p2 ≡ 9 (mod 16), p2 = (−1)(p −1)/8 = −1, so 2 is not a square


modulo p. This is a contradiction, so 2q 6≡ 1 (mod p) : 2 is a primitive root modulo


p.

Ex. 4.8 Let p be an odd prime. Show that a is a primitive root modulo p iff a(p−1)/q 6≡ 1
(mod p) for all prime divisors q of p − 1.
Proof. • If a is a primitive root, then ak 6≡ 1 for all k, 1 ≤ k < p − 1, so a(p−1)/q 6≡ 1
(mod p) for all prime divisors q of p − 1.
• In the other direction, suppose a(p−1)/q 6≡ 1 (mod p) for all prime divisors q of p−1.
Let δ the order of a, and p − 1 = q1a1 q2a2 · · · qkak the decomposition of p − 1 in prime
factors. As δ | p − 1, δ = q1b1 pb22 · · · qkbk , with bi ≤ ai , i = 1, 2, . . . , k. If bi < ai for some
index i, then δ | (p − 1)/qi , so a(p−1)/qi ≡ 1 (mod p), which is in contradiction with the
hypothesis. Thus bi = ai for all i, and δ = q − 1 : a is a primitive root modulo p.

Ex. 4.9 Show that the product of all the primitive roots modulo p is congruent to
(−1)φ(p−1) modulo p.
Proof. Here we suppose p prime, p > 2. Let g a primitive root modulo p. U (Z/pZ) is
cyclic, generated by g:
U (Z/pZ) = {1, g, g 2 , . . . , g p−2 }, g p−1 = 1.
g k is a primitive element iff k ∧ (p − 1) = 1, so the product of primitive elements in
U (Z/pZ) is Y
P = gk .
k∧(p−1)=1
1≤k<p−1

gS ,
P
so P = where S = k.
k∧(p−1)=1
1≤k<p−1
From Ex. 2.22, we know that for n ≥ 2,
X 1
k = nφ(n).
2
k∧n=1
1≤k<n

k = 12 (p − 1)φ(p − 1).
P
So S =
k∧(p−1)=1
1≤k<p−1
As p > 2, p − 1 is even. (g (p−1)/2 )2 = g p−1 = 1, and g (p−1)/2 6= 1. As Z/pZ is a field,
g (p−1)/2 = −1.
Thus P = (−1)φ(p−1) : so the product P of all the primitive roots modulo p is such
that
P ≡ (−1)φ(p−1) (mod p).

4
Ex. 4.10 Show that the sum of all the primitive roots modulo p is congruent to µ(p−1)
modulo p.

Proof. Notation : Fp = Z/pZ is the field with p elements, |x| the multiplicative order of
an element x ∈ F∗p , N∗ = {1, 2, 3, . . .}.
Let 
 N∗ → Fp P
ψ: n 7→ ψ(n) = d
d∈F∗ ,|d|=n

p

ψ(n) is the sum of the elements with order n in F∗p .


So ψ(n) = 0 if n - p − 1, and
S = ψ(p − 1) is the sought sum of all the primitive roots modulo p.
We compute for all n ∈ N∗ X
f (n) = ψ(d).
d|n

f (n) is the sum of elements whose order divides n, in other worlds the sum of the
roots of xn − 1. This sum is, up to the sign, the coefficient of xn−1 , so is null, except in
the case n = 1, where the sum of the unique root 1 of x − 1 is 1. So

f (1) = 1, ∀n > 1, f (n) = 0,

(f = χ{1} is the characteristic function of {1}).


From the Möbius inversion formula, for all n ∈ N∗ , ψ(n) = d|m µ n
P 
d f (d), so
X p − 1
ψ(p − 1) = µ f (d) = µ(p − 1).
d
d|p−1

Conclusion : X
S= d = µ(p − 1) :
d∈F∗p ,|d|=p−1

the sum of all the primitive roots modulo p is congruent to µ(p − 1) modulo p.

Ex. 4.11 Prove that 1k + 2k + ... + (p − 1)k ≡ 0 (mod p) if p − 1 - k, and −1 (mod p)


if p − 1 | k.

Proof. Let Sk = 1k + 2k + · · · + (p − 1)k .


Let g a primitive root modulo p : g a generator of F∗p .
As (1, g, g 2 , . . . , g p−2 ) is a permutation of (1, 2, . . . , p − 1),
k k k
Sk = 1 + 2 + · · · + p − 1
p−2
(
X
ki
p − 1 = −1 if p − 1 | k
= g = g (p−1)k −1
g k −1
= 0 if p − 1 - k
i=0

since p − 1 | k ⇐⇒ g k = 1.
Conclusion :

1k + 2k + · · · + (p − 1)k ≡ 0 (mod p) if p − 1 - k
k k k
1 + 2 + · · · + (p − 1) ≡ −1 (mod p) if p − 1 | k

5
Ex. 4.12 Use the existence of a primitive root to give another proof of Wilson’s theorem
(p − 1)! ≡ −1 (mod p).

Proof. As the result is trivial if p = 2, we suppose that p is an odd prime.


Let g a primitive root modulo p : g a generator of F∗p .
As (g (p−1)/2 )2 = g p−1 = 1, and g (p−1)/2 6= 1 in the field F∗p , then g (p−1)/2 = −1, and
(1, g, g 2 , . . . , g p−2 ) is a permutation of (1, 2, . . . , p − 1), so
p−2
Y
(p − 1)! = gk
k=0
Pp−2
k
=g k=0

= g (p−2)(p−1)/2
 p−2
= g (p−1)/2
= (−1)p−2
= −1.

Hence (p − 1)! ≡ −1 (mod p) for each prime p.

Ex. 4.13 Let G be a finite cyclic group and g ∈ G a generator. Show that all the other
generators are of the form g k , where (k, n) = 1, n being the order of G.

Proof. Suppose G = hgi, with Card G = n, so the order of g is n.


Let x another generator of G, then x = g k , and g = xl , k, l ∈ Z, so g = g kl , g kl−1 =
e : n | kl − 1, then kl − 1 = qn, q ∈ Z, so n ∧ k = 1.
Conversely, if u ∧ k = 1, there exist u, v ∈ Z such that un + vk = 1, so g = g un+vk =
(g )u (g k )v = xv ∈ hxi, so G ⊂ hxi, G = hxi : x is a generator of G.
n

Conclusion : if g is a generator of G, all the other generators are the elements g k ,


where k ∧ n = 1, n = |G|.

Ex. 4.14 Let A be a finite abelian group and a, b ∈ A elements of order m and n,
respectively. If (m, n) = 1, prove that ab has order mn.

Proof. Suppose |a| = m, |b| = n, m ∧ n = 1.


• If (ab)k = e, then ak = b−k , so akn = b−kn = (bn )−k = e, so m | kn, with m ∧ n = 1,
so m | k.
Similarly, bkm = a−km = (am )−k = e, so n | km, n ∧ m = 1 : n | k.
As n | k, m | k, n ∧ m = 1, nm | k.
• Conversely, if nm | k, k = qnm, q ∈ Z, so (ab)k = ak bk = (am )qn (bn )qm = e.

∀k ∈ Z, (ab)k = e ⇐⇒ nm | k.

So |ab| = nm.

Ex. 4.15 Let K be a field and G ⊂ K ∗ a finite subgroup of the multiplicative group of
K. Extend the arguments used in the proof of Theorem 4.1 to show that G is cyclic.

Solution 1.

6
Proof. Let n = |G|. From Lagrange’s theorem, an = 1 for all a ∈ G, so the polynomial
xn − 1 ∈ K[x] has exactly n roots in G, and so

∀x ∈ K, x ∈ G ⇐⇒ xn = 1.

If d | n, the polynomial xd − 1 ∈ K[x] has exactly d roots in K otherwise xn − 1 =


(xd − 1)g(x), g(x) ∈ K[x], and deg(g) = n − d has at most n − d roots, so xn − 1 would
have less than n roots in K. As xd0 = 1 ⇒ xn0 = 1, all these roots are in G : xd − 1 has d
roots in G. P
Let ψ(d) the number of elements in G of order d ( ψ(d)
P = 0 if d - n). Then c|d ψ(c) =
d. Applying the Möbius inversion theorem, ψ(d) = c|d µ(c)d/c = φ(d) (Prop. 2.2.5),
in particular, ψ(n) = φ(n) ≥ 1. This proves the existence of an element of order n in G,
so G is cyclic.
(variation : ψ(d) = 0 if there exists no element
P of order d,Pand ψ(d) = φ(d) otherwise
: see Ex.4.13. So ψ(d) ≤ φ(d) for all d | n. As d|n ψ(d) = d|n φ(d) = n, ψ(d) = φ(d)
for all d | n. So there exists in G an element of order n, and G is cyclic.)

Solution 2.

Proof. Let n = |G| = pa11 · · · pakk . From Lagrange’s theorem, y n = 1 for all y ∈ G.
p(x) = xn/p1 − 1 ∈ K[x] has at most n/p1 < n roots in K ∗ , a fortiori in G, so there
exists a ∈ G such that an/p1 6= 1.
a1 a2 ak a1 a1 −1
p p
Let c1 = an/p1 = ap2 ···pk . Then c11 = 1 and c11 = an/p1 6= 1, so |c1 | = pa11 .
Similarly, there exist c2 , . . . , ck with respective orders |ci | = pai i .
From exercise 4.14, we obtain by induction that c = c1 · · · ck has order pa11 · · · pakk = n,
so G is cyclic.

Ex. 4.16 Calculate the solutions to x3 ≡ 1 (mod 19) and x4 ≡ 1 (mod 17).

Proof. Here we note a the class of x in Z/pZ.


Let a ∈ F19 . Then

a3 − 1 = 0 ⇐⇒ a − 1 = 0 or a2 + a + 1 = 0.

a2 + a + 1 = 0 ⇐⇒ (a + 10) − 99 = 0
⇐⇒ (a + 10)2 − 4 = 0
⇐⇒ (a + 8)(a + 12) = 0

So, for all x ∈ Z,

x3 ≡ 1 (mod 19) ⇐⇒ x ≡ 1, 7, 11 (mod 19).

Let a ∈ F17 .

a4 = 1 ⇐⇒ a2 = 1 or a2 = −1 = 42
⇐⇒ a = ±1 or a = ±4

So, for all x ∈ Z,

x4 ≡ 1 (mod 17) ⇐⇒ x ≡ −1, 1, −4, 4 (mod 17).

7
Alternatively, we can take primitives roots modulo 19 and 17.
2 is a primitive root modulo 19, Let a = 2k ∈ F19 .

a3 = 1 ⇐⇒ 23k = 1
⇐⇒ 18 | 3k
⇐⇒ 6 | k
⇐⇒ a = 1, 26 = 7, 212 = 11

3 is a primitive root modulo 17. Let a = 3k ∈ F17 .

a4 = 1 ⇐⇒ 34k = 1
⇐⇒ 16 | 4k
⇐⇒ 4 | k
⇐⇒ a = 1, 34 = −4, 38 = −1, 312 = 4

Ex. 4.17 Use the fact that 2 is a primitive root modulo 29 to find the seven solutions
to x7 ≡ 1 (mod 29).

Proof. Let x ∈ Z, then x ≡ 2k (mod 29), k ∈ N.

x7 ≡ 1 (mod 29) ⇐⇒ 27k ≡ 1 (mod 29)


⇐⇒ 28 | 7k
⇐⇒ 4 | k

So the group cyclic S of the roots of x7 − 1 in F29 are

S = {1, 24 , 28 , 212 , 216 , 220 , 224 },

S = {1, 16, 24, 7, 25, 23, 20}.

Ex. 4.18 Solve the congruence 1 + x + · · · + x6 ≡ 0 (mod 29).

Proof. As (1 + x + · · · + x6 )(1 − x) = 1 − x7 ,

x7 ≡ 1

6 (mod 29)
1 + x + ··· + x ≡ 0 (mod 29) ⇐⇒
x 6≡ 1 (mod 29)

From Ex. 4.17, the solutions are congruent to 24 , 28 , 212 , 216 , 220 , 224 modulo 29.

Ex. 4.19 Determine the numbers a such that x3 ≡ a (mod p) is solvable for p =
7, 11, 13.

Proof. (a) If p = 7, then 3 | p − 1, d = 3 ∧ (p − 1) = 3. From Prop. 4.2.1,

∃x ∈ Z, a ≡ x3 (mod 7) ⇐⇒ a ≡ 0 (mod 7) or a(p−1)/3 = a2 ≡ 1 (mod 7).

So the numbers a such that x3 ≡ a (mod 7) is solvable are congruent at 0, 1, −1


modulo 7.

8
(b) If p = 11, then d = 3 ∧ (p − 1) = 1. With the same proposition,

∃x ∈ Z, a ≡ x3 (mod 11) ⇐⇒ a ≡ 0 (mod 11) or ap−1 = a6 ≡ 1 (mod 11).

So all integers a are cube modulo 11, in only one way.


For an alternative proof, the application
 ∗
F11 → F∗11
f:
x 7→ x3

f is a bijection. Indeed,
• f is a group homomorphism,
• x3 = 1 ⇒ (x3 )7 = 1 ⇒ x = 1 so ker(f ) = {1},
• f : F∗11 → F∗11 is injective and F∗11 is finite, so f is bijective.
In F11 , 0 = 03 , 1 = 13 , 2 = 73 , 3 = 93 , 4 = 53 , 5 = 33 , 6 = 83 , 7 = 63 , 8 = 23 , 9 =
43 , 10 = 103 .

(c) If p = 13, then 3 | p − 1, 3 ∧ (p − 1) = 3, so

∃x ∈ Z, a ≡ x3 (mod 13) ⇐⇒ a ≡ 0 (mod 13) or a(p−1)/3 = a4 ≡ 1 (mod 13)


⇐⇒ a ≡ 0, 1, −1, 5, −5 (mod 13)

(5 ≡ 83 (mod 13).)

Ex. 4.20 Let p be a prime, and d a divisor of p − 1. Show that dth powers form a
subgroup of U (Z/pZ) of order (p − 1)/d. Calculate this subgroup for p = 11, d = 5, for
p = 17, d = 4, and for p = 19, d = 6.

Proof. Here p is a prime number, and d | p − 1. Let


 ∗
Fp → F∗p
f:
x → xd

Then f is a group homomorphism, and im(f ) is the set of dth powers, and consequently
is a subgroup of U (Fp ) = F∗p . ker(f ) is the group of the roots of xd − 1. As d | p − 1, the
polynomial xd − 1 has exactly d roots (Prop. 4.1.2), so | ker(f )| = d.
As im(f ) ' F∗p / ker(f ),

|im(f )| = |F∗p |/| ker(f )| = (p − 1)/d.

So there exist exactly (p − 1)/d dth powers in (Z/pZ)∗ .

From Prop. 4.2.1, as d | p − 1, d ∧ (p − 1) = d, for all x ∈ F∗p ,

x ∈ im(f ) ⇐⇒ x(p−1)/d = 1.
So the group of dth powers is the group of the roots of x(p−1)/d − 1.
• If p = 11, d = 5, im(f ) = {1, −1}.
• If p = 17, d = 4, x ∈ im(f ) ⇐⇒ x4 = 1 : im(f ) = {1, −1, 4, −4}.
• If p = 19, d = 6, x ∈ im(f ) ⇐⇒ x3 = 1 : im(f ) = {1, 7, 72 = 11},
where 7 ≡ 26 (mod 19).

9
Ex. 4.21 If g is a primitive root modulo p, and d|p − 1, show that g (p−1)/d has order
d. Show also that a is a dth power iff a ≡ g kd (mod p) for some k. Do Exercises 16-20
making use of those observations.

Proof. Let x = g (p−1)/d ∈ F∗p , where g is a primitive root modulo p. For all k ∈ Z,
p−1
xk = 1 ⇐⇒ g k =1
d

p−1
⇐⇒ p − 1 | k
d
⇐⇒ d | k

So the order of g (p−1)/d is d.


• If a = g kd , then a = xd , where x = g k , so a is a dth power.
• If a 6= 0 is a dth power, a = xd , x ∈ F∗p . As x ∈ hgi, x = g k , so a = g kd .
So, if a 6≡ 0 (mod p), a is a dth power iff a ≡ g kd (mod p) for some k.
By example (Ex. 4.20), 2 is a primitive root modulo 19, so the 6th powers modulo
19 are 20 = 1, 26 = 7, 212 = 11.

Ex. 4.22 If a has order 3 modulo p, show that 1 + a has order 6.

Proof. If a has order 3 modulo p, then 0 ≡ a3 − 1 = (a − 1)(a2 + a + 1) (mod p), with


a 6≡ 1 (mod p), so a2 + a + 1 ≡ 0 (mod p). Thus

(1 + a)3 ≡ 1 + 3a + 3a2 + a3
≡ 1 + 3a + 3(−1 − a) + 1
≡ −1 (mod p)

So (1 + a)6 ≡ 1 (mod p).


(1 + a)2 ≡ 1 + 2a + a2 = 1 + 2a + (−1 − a) ≡ a 6≡ 1 (mod p).
So (1 + a)6 ≡ 1, (1 + a)2 6≡ 1, (1 + a)3 6≡ 1 (mod p), so the order of 1 + a divides 6,
but doesn’t divides 2 or 3, so 1 + a has order 6 modulo p.

Ex. 4.23 Show that x2 ≡ −1 (mod p) has a solution iff p ≡ 1 (mod 4), and that
x4 ≡ −1 (mod p) has a solution iff p ≡ 1 (mod 8).

Proof. If x2 ≡ −1 (mod p), then x has order 4 in F∗p , hence from Lagrange’s theorem,
4 | p − 1.
Conversely, suppose 4 | p − 1, so p = 4k + 1, k ∈ N∗ . From proposition 4.2.1, as
2 | p − 1, −1 is a square modulo p iff (−1)(p−1)/2 ≡ 1 (mod p), which is true because
(−1)(p−1)/2 = (−1)2k = 1.
If x4 ≡ −1 (mod p), then x8 = 1 ∈ F∗p , and x4 6= 1, so x has order 8 in F∗p , so 8 | p − 1.
Conversely, if p ≡ 1 (mod 8), p = 8K + 1, K ∈ N∗ . From Prop.4.2.1, as 4 | p − 1,
there exists x ∈ Z such that −1 = x4 iff (−1)(p−1)/4 ≡ 1 (mod 8), which is true because
(−1)(p−1)/4 = (−1)2K = 1.
Conclusion :

∃x ∈ Z, x4 ≡ −1 (mod p) ⇐⇒ p ≡ 1 (mod 8).

10
Ex. 4.24 Show that axm + by n ≡ c (mod p) has the same number of solutions as
0 0
axm + by n ≡ c (mod p), where m0 = (m, p − 1) and n0 = (n, p − 1).

Proof. If a∧b - c, the two equations have no solution. So we can suppose a∧b | c, and after
division by δ = a ∧ b, we obtain an equation a0 xm + b0 y n = c0 , a0 = a/δ, b0 = bδ, c0 = cδ,
and a0 ∧ b0 = 1. So it remains to prove that axm + by n ≡ c (mod p) has the same number
0 0
of solutions as axm + by n ≡ c (mod p) when a ∧ b = 1.
In this case the equation au + bv = c has solutions. Let N the number of solutions
(x, y) of the equation a xm + b y n = c,N 0 the number of solutions (x, y) of the equation
0 0
a xm + b y n = c. Then

N = Card{(x, y) ∈ Fp × Fp | a xm + b y n = c}
X
= Card{(x, y) ∈ Fp × Fp | xm = u, y n = v}
au+bv=c
X
= Card{x ∈ Fp |xm = u} × Card{y ∈ Fp |y n = v}.
au+bv=c

The same is true for N 0 , so it is suffisiant to prove that


0
Card{x ∈ Fp | xm = u} = Card{x ∈ Fp | xm = u},

where m0 = m ∧ (p − 1), and a similar equality for the equation y n = v.


Let g a generator of F∗p . Write u = g r , r ∈ N.

∃x ∈ Fp , xm = u ⇐⇒ ∃k ∈ Z, g mk = g r
⇐⇒ ∃k ∈ Z, p − 1 | mk − r
⇐⇒ ∃k ∈ Z, ∃l ∈ Z, r = mk + l(p − 1)
⇐⇒ m ∧ (p − 1) | r

So
{x ∈ Fp | xm = u} =
6 ∅ ⇐⇒ m ∧ (p − 1) | r,
and similarly
0
6 ∅ ⇐⇒ m0 ∧ (p − 1) | r.
{x ∈ Fp | xm = u} =
Since m0 ∧ (p − 1) = (m ∧ (p − 1)) ∧ (p − 1) = m ∧ (p − 1), these two conditions are
equivalent, so these two sets are empty for the same values of u.
Let u is such that {x ∈ Fp | xm = u} = 6 ∅, and x0 a fixed solution of xm = u.
Write x = g , x0 = g 0 . Let d = m ∧ (p − 1)(= m0 ).
k k

xm = u ⇐⇒ xm = x0 m
⇐⇒ g mk = g mk0
⇐⇒ p − 1 | m(k − k0 )
p−1 m
⇐⇒ | (k − k0 )
d d
p−1
⇐⇒ | k − k0
d
p−1
⇐⇒ ∃j ∈ Z, k = k0 + j
d

11
p−1 p−1 p−1
As g is a primitive root modulo p, the distinct solutions are x0 , x0 g d , . . . , x0 g k d , . . . x0 g (d−1) d ,
so in this case
Card{x ∈ Fp | xm = u} = d = m ∧ (p − 1).
As m0 ∧ (p − 1) = m ∧ (p − 1),
0
Card{x ∈ Fp | xm = u} = Card{x ∈ Fp | xm = u}.
0 0
So N = N 0 : axm + by n ≡ c (mod p) has the same number of solutions as axm + by n ≡ c
(mod p), where m0 = (m, p − 1) and n0 = (n, p − 1).

Ex. 4.25 Prove Propositions 4.2.2 and 4.2.4.


Proposition 4.2.2. Suppose that a is odd, e ≥ 3, and consider the congruence
xn ≡ a (mod 2e ). If n is odd, a solution always exists and it is unique.
e−2
If n is even, a solution exists iff a ≡ 1 (mod 4), a2 /d ≡ 1 (mod 2e ), where d =
(n, 2e−2 ). When a solution exists there are exactly 2d solutions.

Proof. We suppose that a is odd and e ≥ 3.


From Theorem 2’, we know that {(−1)a 5b | 0 ≤ a ≤ 1, 0 ≤ b ≤ 2e−2 } constitutes a
reduced residue system modulo 2e , so we can write

a ≡ (−1)s 5t (mod 2e ), 0 ≤ s ≤ 1, 0 ≤ t ≤ 2e−2 ,


x ≡ (−1)y 5z (mod 2e ), 0 ≤ y ≤ 1, 0 ≤ z ≤ 2e−2 .

For all x ∈ Z,

xn ≡ a (mod 2e ) ⇐⇒ (−1)ny 5nz ≡ (−1)s 5t (mod 2e )

Then (−1)ny ≡ (−1)s (mod 4), ny ≡ s (mod 2), (−1)ny = (−1)s , so 5nz ≡ 5t (mod 2e ).
Conversely, if ny ≡ s (mod 2) and 5nz ≡ 5t (mod 2e ), then xn ≡ a (mod 2e ), so
 
ny ≡ s (mod 2) ny ≡ s (mod 2)
xn ≡ a (mod 2e ) ⇐⇒ ⇐⇒
5nz ≡ 5t (mod 2e ) nz ≡ t (mod 2e−2 )

since the order of 5 modulo 2e is 2e−2 .


• Suppose that n is an odd integer. Then
 
ny ≡ s (mod 2) y ≡ s (mod 2)
⇐⇒
nz ≡ t (mod 2e−2 ) z ≡ n0 t (mod 2e−2 )

where n0 is an inverse of n modulo 2e−2 : nn0 ≡ 1 (mod 2e−2 ).


So xn ≡ a (mod 2e ) has an unique solution modulo 2e .
• Suppose
 that n is an even integer.
ny ≡ s (mod 2)
Then implies s ≡ 0 (mod 2) and d = n ∧ 2e−2 | t.
nz ≡ t (mod 2e−2 )
Then a ≡ (−1)s 5t ≡ 5t (mod 2e ), so a ≡ 1 (mod 4).
2e−2
 e−2  t
d
Hence a d ≡ 52 ≡ 1 (mod 2e ), since 5 has order 2e−2 , and d | t.
So, if n is even, and d = n ∧ 2e−2 ,
(
a ≡ 1 (mod 4),
∃x ∈ Z, xn ≡ a (mod 2e ) ⇒ 2e−2
a d ≡ 1 (mod 2e ).

12
(
a ≡ 1 (mod 4),
Conversely, suppose that 2e−2
a ≡ 1 (mod 2e ).
d

Then a ≡ (−1)s 5t (mod 2e ) implies a ≡ (−1)s (mod 4), so s is even, and a ≡ 5t


(mod 2e ).
2e−2 e−2
Therefore 5t d ≡ 1 (mod 2e ), which implies 2e−2 | t 2 d , so d | t.

ny ≡ s (mod 2)
∃x ∈ Z, xn ≡ a (mod 2e ) ⇐⇒ ∃y ∈ Z, ∃z ∈ Z,
nz ≡ t (mod 2e−2 )
⇐⇒ ∃z ∈ Z, nz ≡ t (mod 2e−2 ) (since n, s even)
e−2
⇐⇒ ∃z ∈ Z, 2 | nz − t
n 2e−2 t
⇐⇒ ∃z ∈ Z, z− |
d d d
2e−2 n t
⇐⇒ ∃z ∈ Z, ∃q ∈ Z, q +z =
d d d
e−2 2e−2
As 2 d ∧ nd = 1, there exists a solution (q, z0 ) of this last equation, where 0 ≤ z0 < d ,
and so x0 = 5z0 is a particular solution of xn ≡ a (mod 2e ), therefore
(
n e
a ≡ 1 (mod 4)
∃x ∈ Z, x ≡ a (mod 2 ) ⇐⇒ 2e−2
a d ≡ 1 (mod 2e )

If there exists a particular solution x0 ≡ (−1)y0 5z0 , then

xn ≡ a (mod 2e ) ⇐⇒ xn ≡ xn0 (mod 2e )



ny ≡ ny0 (mod 2)
⇐⇒
nz ≡ nz0 (mod 2e−2 )
⇐⇒ n(z − z0 ) ≡ 0 (mod 2e−2 ) (since n even)
2e−2 n
⇐⇒ | (z − z0 )
d d
2 e−2 2e−2 n
⇐⇒ | z − z0 , (since ∧ = 1)
d d d
2 e−2
⇐⇒ ∃k ∈ Z, z = z0 + k
d
As the order of 5 modulo 2e is 2e−2 , the solutions of xn ≡ a (mod 2e ) are
2e−2
xk = (−1)y 5z0 +k d , 0 ≤ y < 2, 0 ≤ k < d,

so there are exactly 2d solutions modulo 2e .

Proposition 4.2.4. Let 2l be the highest power of 2 dividing n. Suppose that a is


odd and that xn ≡ a (mod 22l+1 ) is solvable. Then xn ≡ a (mod 2e ) is solvable for all
e ≥ 2l + 1 (and consequently for all e ≥ 1). Moreover, all these congruences have the
same number of solutions.

Proof. We suppose that a is odd, and that xn ≡ a (mod 22l+1 ) is solvable. l is such that
n = 2l n0 , where n0 is an odd integer.

13
Let the induction hypothesis be, for a fixed integer m ≥ 2l + 1,
∃x0 ∈ Z, xn0 ≡ a (mod 2m ).
Let x1 = x0 + b2m−l : we show that for an appropriate choice of b ∈ {0, 1}, xn1 ≡ a
(mod 2m+1 ).
xn1 = xn0 + nb2m−l xn−1
0 + 22m−2l A, A ∈ Z.
Since m ≥ 2l + 1, 2m − 2l ≥ m + 1, so
xn1 ≡ xn0 + nb2m−l xn−1
0 (mod 2m+1 ).

xn1 ≡ a (mod 2m+1 ) ⇐⇒ (xn0 − a) + n0 bxn−1 0 2m ≡ 0 (mod 2m+1 )


xn − a
⇐⇒ 0 m + n0 bxn−1 0 ≡ 0 (mod 2)
2
As a is odd, and xn0 ≡ a (mod 2m ), m ≥ 1, x0 is odd, and n0 is odd, so there exists an
xn −a
unique b ∈ {0, 1} such that 02m + n0 bx0n−1 ≡ 0 (mod 2). So there exists x1 ∈ Z such
that xb1 ≡ a (mod 2m+1 ), and the induction is completed. Therefore, xn ≡ a (mod 2e )
is solvable for all e ≥ 2l + 1, and consequently for all e ≥ 1.

From the Proposition 4.2.2., with the hypothesis e ≥ 3, we know that the number of
solutions of the solvable equation xn ≡ a (mod 2e ), e ≥ 2l + 1, is 1 if n is odd, 2(n ∧ 2e−2 )
if n is even.
If n is even, l ≥ 1, e ≥ 2l + 1 ≥ 3. Since e ≥ 2l + 1, and n = 2l n0 for an odd n0 ,
l ≤ e−1 e−2 = n0 2l ∧ 2e−2 = 2l , and the number of solutions is 2l+1 ,
2 ≤ e − 2, so n ∧ 2
independent of e ≥ 2l + 1.
Conclusion : under the hypothesis xn ≡ a (mod 22l+1 ), where l = ord2 (n), then
x ≡ a (mod 2e ) is solvable for all e ≥ 1, and all these congruences have the same
n

number of solutions for e ≥ 2l + 1, e ≥ 3.

Chapter 5
5 3 6 −1
   
Ex. 5.1 Use Gauss’ lemma to determine 7 , 11 , 13 , p .
Proof. • a = 5, p = 7.
The array of values of the least residues modulo p = 7, for 1 ≤ k ≤ (p − 1)/2.
k mod 7 1 2 3
5k mod 7 −2 3 1
So the number of negative least residues is µ = 1, and 57 = (−1)µ = −1.


• a = 3, p = 11.
k mod 11 1 2 3 4 5
3k mod 11 3 −5 −2 1 4
3
= (−1)µ = 1.

So µ = 2, 11
• a = 6, p = 13.
k mod 13 1 2 3 4 5 6
6k mod 13 6 −1 5 −2 4 −3
6
= (−1)µ = −1.

So µ = 3, 13
• If a = −1, and p is an odd prime, the values of the least residues of −k modulo p
for k = 1, 2, . . . , (p − 1)/2are −k, all negative. So the number of negative least residues
is µ = (p − 1)/2, and −1 p = (−1)
(p−1)/2 .

14
Ex. 5.2 Show that the number of solutions to x2 ≡ a (mod p) is equal to 1 + (a/p).

Proof. Let N the number of solutions of x2 ≡ a (mod p).


• If ap = 0, then p | a, a ≡ 0 (mod p), so the unique solution of x2 ≡ a = 0 is x ≡ 0
(mod p), so N = 1 = 1 + ap .


• If ap = −1, then N = 0 = 1 + ap .
 

• If ap = 1, then x2 ≡ a (mod p) has a solution x0 , and x2 ≡ a (mod p) ⇐⇒ x2 ≡




x20 (mod p) ≡ p | (x − x0 )(x + x0 ) ≡ x ≡ ±x0 (mod p), so N = 2 = 1 + ap .




Ex. 5.3 Suppose p 6 | a. Show that the number of solutions to ax2 + bx + c ≡ 0 (mod p)
is equal to 1 + ((b2 − 4ac)/p).

Proof. Here p is an odd prime number, and p - a. Let N be the number of solutions of
ax2 + bx + c ≡ 0 (mod p)
For x ∈ Fp = Z/pZ,
 
2 2 b c
ax + bx + c = a x + x +
a a
 2 2
!
b b − 4ac
=a x+ −
2a 4a2
 2
b ∆
Let ∆ = b2 − 4ac. Then N is the number of solutions of x + 2a − 4a2
= 0 in Fp . As
in Ex.5.2, N = 1 if ∆ = 0, N = 0 if ∆ is not a square in F∗p , otherwise ∆ = δ 2 , δ ∈ F∗p ,
and the solutions are x = (−b ± δ)/2a, so N = 2. In the three cases, N = 1 + ∆

p .

Pp−1
Ex. 5.4 Prove that a=1 (a/p) = 0.

Proof. Here p is an odd prime (the result is false if p = 2). In the interval [1, p − 1],
there exist (p − 1)/2 residues, and (p − 1)/2 nonresidues (Prop. 5.1.2., Corollary 1), so
P p−1
a=1 (a/p) = 0.

p−1
a b
P  
Proof. As an alternative proof, let S = p , and b a nonresidue modulo p : p = −1
a=1
(such a b exists if p 6= 2). As a 7→ ab is a bijection from F∗p to itself,
  p−1   p−1  
b X ab X c
S= = = S,
p p p
a=1 c=1

so −S = S, S = 0.

Ex. 5.5 Prove that p−1


P
x=1 ((ax + b)/p) = 0 provided that p - a.
There is a mistake
Pp−1 in the sentence : we must read
Prove that x=0 ((ax + b)/p) = 0 provided that p - a.
For instance,
5−1        
X x+1 2 3 4
= + + = −1 6= 0.
5 5 5 5
x=1

15
0

Proof. From exercise 5.3, as p = 0, we know that

X  x Xp−1   p−1  
x X x
= = = 0.
p p p
x∈Fp x=0 x=1

x0
(This sum is well defined, since xp depends only of x : x ≡ x0 (mod p) ⇒ x
  
p = p .)

Fp → Fp
As a 6= 0 in Fp , f : is a bijection. Thus
x 7→ ax + b

p−1   X f (x)
X ax + b
=
p p
x=0 x∈Fp
X y 
= (y = f (x))
p
y∈Fp

=0

Ex. 5.6 Show that the number of solutions to x2 − y 2 ≡ a (mod p) is given by:
p−1   2 
X y +a
1+ .
p
y=0

Proof. Let S = {(x, y) ∈ F2p | x2 − y 2 = a}. From Ex.5.2,


X
|S| = Card {x ∈ Fp | x2 = y 2 + a}
y∈Fp
p−1   2 
X y +a
= 1+ .
p
y=0

Ex. 5.7 By calculating directly show that the number of solutions to x2 − y 2 ≡ a


(mod p) is p − 1 if p - a, and 2p − 1 if p | a. (Hint. Use the change of variables
u = x + y, v = x − y.)

Proof. Let S = {(x, y) ∈ F2p | x2 − y 2 = a}, and T = {(u, v) ∈ F2p | u v = a}. Then

S → T
f : is well defined (if (x, y) ∈ S, (x − y)(x + y) = a, so
(x, y) 7→ (x + y, x − y)
(x + y, x − y) ∈ T ). Moreover f is a bijection, with inverse (u, v) 7→ ((u + v)/2, (u − v)/2),
so |S| = |T |.
We compute |T |.
• Suppose that p - a, so a 6= 0. For v = 0, there is no solution, and for each v 6= 0,
we obtain the unique solution (a v −1 , v), so there exist p − 1 solutions.
• Suppose that p | a. The solutions of uv = 0 are (0, 0), (0, v) for each v 6= 0, (u, 0)
for each v = 0, that is to say N = 1 + (p − 1) + (p − 1) = 2p − 1 solutions.

16
Conclusion :
Card {(x, y) ∈ F2p | x2 − y 2 = a} = p − 1 if p - a
= 2p − 1 if p | a

Ex. 5.8 Combining the results of Ex. 5.6 and 5.7 show that:
p−1  2  (
X y +a −1 if p - a
=
p p − 1 if p | a
y=0

Proof. Let S = {(x, y) ∈ F2p | x2 − y 2 = a}.


p−1
P 2 
We obtain in Ex 5.6, |S| = 1 + y p+a , and in Ex. 5.7. , |S| = p − 1 if p - a,
y=0
|S| = 2p − 1 if p | a.
So
p−1  2  (
X y +a −1 if p - a
|S| − p = =
p p − 1 if p | a
y=0

Ex. 5.9 Prove that 12 32 · · · (p − 2)2 ≡ (−1)(p+1)/2 (mod p) using Wilson’s theorem.

Proof. Here p is an odd prime.


From Wilson’s theorem, as k(p − k) ≡ −k 2 (mod p) for k = 1, 2, . . . , p − 1,
−1 ≡ (p − 1)!
     
p−1 p+1
≡ 1 × 2 × ··· × k × ··· × × × · · · × (p − k) · · · × (p − 2) × (p − 1)
2 2
(p−1)/2
Y
≡ k(p − k)
k=1
(p−1)/2
Y
(p−1)/2
≡ (−1) k2
k=1
  2
(p−1)/2 p−1
≡ (−1) ! (mod p)
2
So  2

p−1
! ≡ (−1)(p+1)/2 (mod p).
2
Moreover, from Wilson’ theorem and Fermat’s little theorem,
12 22 32 · · · (p − 1)2 = [(p − 1)!]2 ≡ 1 (mod p)
  2   2
2 2 2 p−1 2
 p−1 p−1
2 4 · · · (p − 1) = 2 ! ≡ ! (mod p)
2 2

17
Thus   2
p−1
12 32 · · · (p − 2)2 ! ≡1 (mod p),
2
which gives
12 32 · · · (p − 2)2 ≡ (−1)(p+1)/2 (mod p).

Ex. 5.10 Let r1 , r2 , . . . , r(p−1)/2 be the quadratic residues between 1 and p. Show that
their product is congruent to 1 (mod p) if p ≡ 3 (mod 4), and to −1 if p ≡ 1 (mod 4).

Proof. We proved in Ex. 5.9 that


  2
p−1
! ≡ (−1)(p+1)/2 (mod p).
2

{1, 2, . . . , (p − 1)/2} →
7 {r1 , r2 , . . . , r(p−1)/2 }
The application f : is a bijection, so
x 7 → x2

(p−1)/2   2
Y p−1
ri ≡ ! (mod p),
2
i=1

so
(p−1)/2
Y
ri ≡ (−1)(p+1)/2 (mod p).
i=1

That is to say, the product of the quadratic residues between 1 and p is congruent to 1
(mod p) if p ≡ 3 (mod 4), and to −1 if p ≡ 1 (mod 4).

Ex. 5.11 Suppose that p ≡ 3 (mod 4), and that q = 2p + 1 is also prime. Prove that
2p − 1 is not prime. (Hint : Use the quadratic character of 2 to show that q | 2p − 1)
One must assume that p > 3.

Proof. The result is false if p = 3, so we must suppose p > 3.


p = 4k + 3 for an integer k, so q = 2p + 1 = 8k + 7 ≡ −1 (mod 8). Thus
 
2 2
= (−1)(q −1)/8 = 1.
q

So 2(q−1)/2 ≡ 1 (mod q), 2p ≡ 1 (mod q), so q | 2p − 1.


Moreover, as p > 3, q = 2p + 1 < 2p − 1
(2p + 1 < 2p − 1 ⇐⇒ 2p < 2p − 2 ⇐⇒ p + 1 < 2p−1 .
4 + 1 < 24−1 and for all k ≥ 4, k + 1 < 2k−1 implies k + 2 < 2k−1 + 1 ≤ 2k , and
4 + 1 < 24−1 , so by induction k + 1 < 2k−1 for all k > 3.
So q | 2p − 1 with 1 < q < 2p − 1 : 2p − 1 is composite.
Conclusion : if p ≡ 3 (mod 4), p > 3 is prime, and q = 2p + 1 is also prime, then
p
2 − 1 is not a prime.
For instance, le Mersenne’s number 211 −1 = 2047 is not a prime : 2047 = 23×89.

18
Ex. 5.12 Let f (x) ∈ Z[x]. We say that a prime p divides f (x) if there’s an integer n
such that p | f (n). Describe the prime divisors of x2 + 1 and x2 − 2.
2 2
 p divides x + 1 iff there exists a ∈ Z such that −1 ≡ a (mod p), iff p = 2 or
Proof.
−1
p = 1 iff p = 2 or p ≡ 1 (mod 4).
p divides x2 − 2 iff there exists a ∈ Z such that 2 ≡ a2 (mod p), iff p = 2 or p2 = 1


iff p = 2 or p ≡ ±1 (mod 8).

Ex. 5.13 Show that any prime divisor of x4 − x2 + 1 is congruent to 1 modulo 12.

Proof. • As a6 + 1 = (a2 + 1)(a4 − a2 + 1), p | a4 − a2 + 1 implies p | a6 + 1, so −1



p =1
and p ≡ 1 (mod 4).
• p | 4a4 − 4a2 + 4 = (2a − 1)2 + 3, so −3

p = 1.
−3
 p p
As −3 ≡ 1 (mod 4), p = 3 , so 3 = 1, thus p ≡ 1 (mod 3).
4 | p − 1 and 3 | p − 1, thus 12 | p − 1 :

p≡1 (mod 12).

Ex. 5.14 Use the fact that U (Z/pZ) is cyclic to give a direct proof that (−3/p) = 1
when p ≡ 1 (mod 3).[Hint : There is a ρ in U (Z/pZ) of order 3. Show that (2ρ + 1)2 =
−3.]

Proof. Suppose that p ≡ 1 (mod 3). Let g a generator of F∗p . Then g has order p − 1,
thus ρ = g (p−1)/3 has order 3. As ρ3 = 1, ρ 6= 1, then ρ2 + ρ + 1 = 0.

(2ρ + 1)2 = 4ρ2 + 4ρ + 1


= 4(ρ2 + ρ + 1) − 3
= −3.
−3

Thus p = 1.
−3
F∗p

The converse is also true for an odd prime p : if p = 1, then there exists a ∈
−1+a 2
such that −3 = a2 . ρ = 2 has order 3. Indeed ρ = 1+a4−2a = −2−2a
2
4 = −1−a
2 , so

−1 + a −1 − a
1 + ρ + ρ2 = 1 + +
2 2
=0

so ρ 6= 1, ρ3 = 1. The group F∗p contains an element of order 3, thus from Lagrange’s


theorem 3 | p − 1 : p ≡ 1 (mod 3).

Ex. 5.15 If p ≡ 1 (mod 5), show directly that (5/p) = 1 by the method of Ex. 5.14.
[Hint : Let ρ be an element of U (Z/pZ)) of order 5. Show that (ρ+ρ4 )2 +(ρ+ρ4 )−1 = 0,
etc.]

19
Proof. Let g a generator of F∗p . g has order p − 1, thus ρ = g (p−1)/5 has order 5.
Let
α = ρ + ρ4


β = ρ2 + ρ3
As 0 = ρ5 − 1 = (ρ − 1)(1 + ρ + ρ2 + ρ3 + ρ4 ) and ρ 6= 1, then 1 + ρ + ρ2 + ρ3 + ρ4 = 0,
thus

α + β = −1
αβ = ρ3 + ρ4 + ρ + ρ2 = −1

So α, β are the roots in Fp of x2 + x − 1 : α2 + α − 1 = 0.


Thus 4α2 + 4α − 4 = (2α + 1)2 − 5 = 0 : 5 is a square in F∗p and 5

p = 1.

Ex. 5.16 Using quadratic reciprocity find the primes for which 7 is quadratic residue.
Do the same for 15.

Proof. 7 is a quadratic residue for 2 and for the odd primes such that p7 = 1.


From the law of quadratic reciprocity,


   
7 (p−1)/2 p
= 1 ⇐⇒ (−1) =1
p 7

iff either p ≡ 1 (mod 4) and p7 = 1, or p ≡ −1 (mod 4) and p7 = −1.


 

In the first case , p ≡ 1 (mod 4), p ≡ 1, 4, 2 (mod 7), which gives p ≡ 1, −3, 9
(mod 28).
In the second case, p ≡ −1 (mod 4), p ≡ −1, −4, −2 (mod 7), which gives p ≡
−1, 3, −9 (mod 28).
Conclusion : the primes for which 7 is a quadratic residue are 2 and the odd primes
p such that  
7
= 1 ⇐⇒ p ≡ ±1, ±3, ±9 (mod 28).
p

15

15 is a quadratic residue for 2 and for the odd primes such that p = 1.
         
15 3 5 3 5
= 1 ⇐⇒ = = 1 or = = −1
p p p p p
From the examples of theorem 2, we know that
   
3 3
= 1 ⇐⇒ p ≡ 1, −1 (mod 12), = −1 ⇐⇒ p ≡ 5, −5 (mod 12),
p p
   
5 5
= 1 ⇐⇒ p ≡ 1, −1 (mod 5), = −1 ⇐⇒ p ≡ 2, −2 (mod 5).
p p
As 5 ∧ 12 = 1, there exist 8 cases, all possible, which give
 
15
= 1 ⇐⇒ p ≡ ±1, ±7, ±11, ±17 (mod 60).
p
For instance, the primes 2, 7, 11, 17, 43, 53, 59, 61, 67, 137, . . . are suitable.

20
Ex. 5.17 Supply the details to the proof of Proposition 5.2.1 and to the corollary to
the lemma following it.
Proposition 5.2.1

(a) (a1 /b) = (a2 /b) if a1 ≡ a2 (mod b).

(b) (a1 a2 /b) = (a1 /b)(a2 /b).

(c) (a/b1 b2 ) = (a/b1 )(a/b2 ).

Proof. (a) Let b = p1 p2 · · · pm , where the pi are not necessarily


Q distinct
Q primes. For
each prime pi , (a1 , pi ) = (a2 , pi ) (Prop. 5.1.2 (c)), so i (a1 , pi ) = i (a2 , pi ), thus
(a1 /b) = (a2 /b).

(b) From Prop. 5.1.2(b),


Q Q Q Q
(a1 a2 /b) = i (a1 a2 /pi ) = i (a1 /pi )(a2 /pi ) = i (a1 /pi ) i (a2 /pi ) = (a1 /b)(a2 /b).

(c) Let b1 = p1 p2 · · · pm , b2 = q1 q2 · · · ql . Then b1 b2 = p1 p2 · · · pm q1 q2 · · · ql = m+l


Q
i=1 ri ,
where ri = pi for i = 1, . . . , m, ri = qi−m for i = m + 1, . . . , m + l. Then
(a/b1 b2 ) = m+l
Q Qm Ql
i=1 (a/ri ) = i=1 (a/pi ) j=1 (a/qi ) = (a/b1 )(a/b2 ).

Lemma. Let r and s be odd integers. Then

(a) (rs − 1)/2 ≡ ((r − 1)/2) + ((s − 1)/2) (mod 2).

(b) (r2 s2 − 1)/8 ≡ ((r2 − 1)/8) + ((s2 − 1)/8) (mod 2).

(Proof in the book.)

Corollary. Let r1 , r2 , . . . , rm be odd integers. Then


Pm
(a) i=1 (ri − 1)/2 ≡ (r1 r2 · · · rm − 1)/2 (mod 2).
Pm 2 2 2 2
(b) i=1 (ri − 1)/8 ≡ (r1 r2 · · · rm − 1)/8 (mod 2).

Proof. Let P(m) the proposition defined by


m
X
P(m) ⇐⇒ (ri − 1)/2 ≡ (r1 r2 · · · rm − 1)/2 (mod 2).
i=1

Then P(1) ⇐⇒ (r1 − 1)/2 ≡ (r1 − 1)/2 (mod 2) is true, and P(2) is part (a) of the
lemma. If we make the induction hypothesis P(m), then
m+1
X m
X
(ri − 1)/2 = (ri − 1)/2 + (rm+1 − 1)/2
i=1 i=1
≡ (r1 r2 · · · rm − 1)/2 + (rm+1 − 1)/2 (mod 2)
≡ (r1 r2 · · · rm rm+1 − 1)/2 (mod 2),

where the last congruence is a consequence of the part (a) of the Lemma : the induction
is completed, and P(m) is true for all m ≥ 1.
The proof of part (b) is similar.

21
Ex. 5.18 Let D be a square-free integer that is also odd and positive. Show that there
is an integer b prime to D such that (b/D) = −1.

Proof. Let D = p1 p2 · · · pk , where the pi are distinct odd primes.


Let s a nonresidue modulo pk . From Chinese remainder theorem, as pi ∧ pj = 1 if
i 6= j, there exists an integer b such that

b≡1 (mod p1 ), b ≡ 1 (mod p2 ), · · · , b ≡ 1 (mod pk−1 ), b ≡ s (mod pk ).

Then (b/pi ) = 1, i = 1, 2, . . . k − 1, (b/pk ) = −1, so b ∧ pi = 1 for all i = 1, 2, . . . , k. Then


b ∧ D = b ∧ p1 · · · pk = 1 , and
  Yk    
b b b
= = = −1.
D pi pk
i=1

P
Ex. 5.19 Let D be as in Exercise 18. Show that (a/D) = 0, where the sum is over
a reduced residue system modulo D. Conclude that exactly one half of the elements in
U (Z/DZ) satisfy (a/D) = 1.

Proof. Let b such that (b/D) = −1 and b ∧ D = 1 : the existence of b comes from Ex
5.18. P
Let S = a∈A (a/D), where A is reduced residue system modulo D. As two reduced
system modulo D represent the same elements in U (Z/DZ), the sum is independent of
the reduced residue system A : we can write
X
S= (a/D).
a∈U (Z/DZ)

As b ∧ D = 1, we know from Ex. 3.6 that B = bA = {ba | a ∈ A} is also a reduced


system modulo D. In other words, the application U (Z/DZ) → U (Z/DZ), a 7→ ab is a
bijection, so
       X c
b X b a X ba
S= = = =S (c = ab).
D D D D D
a∈U (Z/DZ) a∈U (Z/DZ) c∈U (Z/DZ)

As (b/D) = −1, −S = S, so S = 0.
Since (a/D) = ±1, one half of the elements in U (Z/DZ) satisfy (a/D) = 1, and one
half of the elements in U (Z/DZ) satisfy (a/D) = −1.

Ex. 5.20 (continuation) Let a1 , a2 , . . . , aφ(D)/2 be integers between 1 and D such that
(ai , D) = 1 and (ai /D) = 1. Prove that D is a quadratic residue modulo a prime p 6 | D,
p ≡ 1 (mod 4) iff p ≡ ai (mod D) for some i.

Proof. From Ex. 5.19 we know that there exist exactly φ(D)/2 integers ai between 1
and D such that ai ∧ D = 1 and (ai /D) = 1. So {a1 , . . . , aφ(D)/2 } is the set of all
a ∈ U (Z/DZ) such that (a/D) = 1.
Let D = p1 p2 · · · pk , with distinct pi , and p a prime number, p ≡ 1 (mod 4), p 6∈
{p1 , . . . , pk } (so p = 4k + 1, k ∈ N).

22
(⇐) Suppose that p ≡ ai for some i, 1 ≤ i ≤ φ(D)/2, then (p/D) = (ai /D) = 1, so
(Prop. 5.2.2)
       
D p−1 D−1 p 2k( D−1 ) p p
= (−1) 2 2 = (−1) 2 = = 1.
p D D D
(⇒) Suppose that D is a quadratic residue modulo p. Then (D/p) = 1, so
   
p p−1 D−1 D
= (−1) 2 2 = 1.
D p
Thus p ∈ {a1 , . . . , aφ(D)/2 } since {a1 , . . . , aφ(D)/2 } is the set of all a ∈ U (Z/DZ) such
that (a/D) = 1. Consequently p ≡ ai (mod D) for some i.

Ex. 5.21 Apply the method of Ex. 5.19 and 5.20 to find those primes for which 21 is
a quadratic residue.

Proof. Let D = 21 = 3 × 7 (D is positive, odd and square-free). We first search the


φ(D)/2 = 6 integers a, 1 ≤ a ≤ 21, such that (a/D) = 1.
         
a a a a a
= 1 ⇐⇒ = = 1 or = = −1.
21 3 7 3 7
The first case is equivalent to a ≡ 1 (mod 3), a ≡ 1, 2, 4 (mod 7), that is a ≡ 1, 16, 4
(mod 21).
The second case gives a ≡ −1 (mod 3), a ≡ −1, −2, −4 (mod 7), that is a ≡ −1, −16, −4
(mod 21), or equivalently a ≡ 20, 5, 17 (mod 21).
So A = {1, 4, 5, 16, 17, 20} is the set of the integers a such that 1 ≤ a ≤ 21, (a/D) = 1.
As (21/3) = (21/7) = 0, 21 is not a quadratic residue modulo 3 or 7.
• p ≡ 1 (mod 4).
From Ex.5.20, we know that D = 21 is a quadratic residue modulo an odd prime p,
p 6= 3, p 6= 7, p ≡ 1 (mod 4), iff p ≡ a (mod D) for some a ∈ A.
• p ≡ −1 (mod 4).
 p p−1 D−1
As D = 21 ≡ 1 (mod 4), D p D = (−1)
2 2 = 1, so the same reasoning as in Ex.
5.20 show that D is a quadratic residue modulo 21 iff p ≡ a, a ∈ A.
Conclusion : 21 is a quadratic residue for 2, and for the primes p such that
p ≡ 1, 4, 5, 16, 17, 20 (mod 21).

Ex. 5.22 Use the Jacobi symbol to determine (113/997), (215/761), (514/1093), and
(401/757).

Proof. (113/997) = (997/113) = (93/113) = (113/93) = (20/93) = (22 /93)(5/93) =


(5/93) = (93/5) = (3/5) = (5/3) = (2/3) = −1.
(215/761) = (761/215) = (116/215) = (22 /215)(29/215) = (29/215) = (215/29) =
(12/29) = (22 /29)(3/29) = (3/29) = (29/3) = (2/3) = −1.
(514/1093) = (2/1093)(257/1093) = −(257/1093) = −(1093/57) = −(65/257) =
−(257/65) = −(62/65) = −(2/65)(31/65) = −(31/65) = −(65/31) = −(3/31) =
(31/3) = (1/3) = 1.
(401/757) = (757/401) = (356/401) = (401/89) = (45/89) = (89/45) = (44/45) =
(22 /45)(11/45)
= (11/45) = (45/11) = (1/11) = 1.

23
Ex. 5.23 Suppose that p ≡ 1 (mod 4). Show that there exist integers s and t such
that pt = 1 + s2 . Conclude that p is not a prime in Z[i]. Remember that Z[i] has unique
factorization.
p−1
Proof. As p ≡ 1 (mod 4), then −1

p = (−1)
2 = 1 : −1 is a square modulo p.
So −1 ≡ s (mod p), s ∈ Z : there exist s ∈ Z, t ∈ Z such that pt = 1 + s2 .
2

In Z[i], p|(s + i)(s − i).


If p was a prime in Z[i], then p | s + i ou p | s − i.
This implies s ± i = (a + bi)p, (a, b) ∈ Z2 , thus ±1 = bp, p | 1 : it’s impossible.
Conclusion : if p ≡ 1 (mod 4), p is not a prime in Z[i].

Ex. 5.24 If p ≡ 1 (mod 4), show that p is a sum of two squares, i.e. p = a2 + b2 with
a, b ∈ Z.(Hint : p = αβ, with α and β being non units in Z[i]. Remember that Z[i] has
unique factorisation.)

Proof. Z[i] is a principal ideal domain, thus p prime is in Z[i] iff p is irreducible in Z[i].
If p ≡ 1 (mod 4), p is not a prime from Ex.5.23, so it is not irreducible :
p = αβ, α, β ∈ Z[i], N (α) > 1, N (β) > 1 (where N (a + bi) = a2 + b2 is the complex
norm).
N (p) = p2 = N (u)N (v), 1 < N (u) < p2
Thus N (u) = p, that is p = a2 + b2 , where u = a + bi.
Conclusion : if p is prime in N, p ≡ 1 (mod 4), then p = a2 + b2 , a, b ∈ Z, p is a sum
of two squares.

Ex. 5.25 An integer is called a biquadratic residue modulo p if it is congruent to a


fourth power. Using the identity x4 + 4 = ((x + 1)2 + 1)((x − 1)2 + 1) show that −4 is a
biquadratic residue modulo p iff p ≡ 1 (mod 4).

Proof. x4 + 4 = (x4 + 4x2 + 4) − 4x2 = (x2 + 2)2 − 4x2 = (x2 + 2 − 2x)(x2 + 2 + 2x), so

x4 + 4 = ((x − 1)2 + 1)((x + 1)2 + 1).


If −4 ≡ x4 [p], then p | (x + 1)2 + 1 or p | (x − 1)2 + 1
In the two cases, −1 is a quadratic residue modulo p, thus −1

p = 1 : p ≡ 1 [4].
−1
Conversely, if p ≡ 1 [4], p = 1, then it exists an integer a such that −1 ≡ a2 [p].


Let x = a − 1. Then p | (x + 1)2 + 1, thus p | x4 + 4 : −4 is a biquadratic residue


modulo p.
Conclusion :
∃x ∈ Z, x4 ≡ −4 [p] ⇐⇒ p ≡ 1 [4].

Ex. 5.26 This exercise and Ex. 5.27 and 5.28 give Dirichlet’s beautiful proof that 2
is a biquadratic residue modulo p iff p can be written in the form A2 + 64B 2 , where
A, B ∈ Z. Suppose that p ≡ 1 (mod 4). Then p = a2 + b2 by Ex. 5.24. Take a to be odd.
Prove the following statements:

(a) (a/p) = 1.
2 −1)/8
(b) ((a + b)/p) = (−1)((a+b) .

24
(c) (a + b)2 ≡ 2ab (mod p)

(d) (a + b)(p−1)/2 ≡ (2ab)(p−1)/4 (mod p).

Proof. Let p a prime number, p ≡ 1 [4] : p = 4k + 1, k ∈ N∗ .


Then p = a2 + b2 (Ex. 5.24).
As a, b are not of the same parity, up to exchange a and b, we will suppose that a is
odd (then b is even).
(a)  
a p−1
≡ a 2 = a2k [p].
p
Using the law of quadratic reciprocity for Jacobi’s symbol (Proposition 5.2.2), where
a, p are odd numbers :
     
a p p−1 a−1 p
= (−1) 2 2 = ,
p a a
since p ≡ 1 [4].
If a = p1 p2 · · · pl is the decomposition of a in prime factors, with not necessarily
distinct primes , then
      
p p p p
= ··· .
a p1 p2 pl
Since p = a2 + b2 , p ≡ b2 [pi ], thus ppi = 1 for all i.


   
a p
= = 1.
p a

(b) a + b is odd, and p ≡ 1 [4], thus


     2    
a+b p 2 p 2 2p
= = = .
p a+b a+b a+b a+b
2p
If a + b = q1 q2 · · · ql , as 2p = (a + b)2 + (a − b)2 , 2p ≡ (a − b)2 [qi ], thus qi = 1.
     
2p 2p 2p
= ··· = 1.
a+b q1 ql
(a+b)2 −1
2

Moreover a+b = (−1) 8 , so
 
a+b (a+b)2 −1
= (−1) 8 .
p

(c) (a + b)2 = a2 + b2 + 2ab = p + 2ab ≡ 2ab [p]


p−1 p−1
(d)[(a + b)2 ] 4 ≡ (2ab) 4 [p], thus
p−1 p−1
(a + b) 2 ≡ (2ab) 4 [p].

25
Ex. 5.27 Suppose that f is such that b ≡ af (mod p). Show that f 2 ≡ −1 (mod p),
and that 2(p−1)/4 ≡ f ab/2 (mod p).

Proof. Let f such as b ≡ af [p].


This is equivalent to f = ba−1 in F∗p .
2 2
As a2 = −b , f = −1 : f 2 ≡ −1 [p].
We deduce from Ex. 5.26 (d) and (b) that
 
p−1 p−1 a+b
(2ab) 4 ≡ (a + b) 2 =
p
(a+b)2 −1
≡ (−1) 8

(a+b)2 −1
≡ (f 2 ) 8

(a+b)2 −1 a2 +b2 −1+2ab


≡f 4 =f 4

p−1 ab
≡f 4 f 2 (mod p)
p−1
a

Since a 2 = p = 1 from Ex. 5.26(a)), then
p−1 p−1 p−1 p−1 p−1
(ab) 4 ≡ (a2 f ) 4 ≡a 2 f 4 ≡f 4 [p],

so p−1 p−1 p−1


ab
2 4 f 4 ≡f 2 f 4 [p].
p−1
As f 4 6≡ 0 [p],
p−1 ab
2 4 ≡f 2 [p].

Ex. 5.28 Show that x4 ≡ 2 (mod p) has a solution for p ≡ 1 (mod 4) iff p is of the
form A2 + 64B 2 .

Proof. If p ≡ 1 [4] and if there exists x ∈ Z such that x4 ≡ 2 [p], then


p−1
2 4 ≡ xp−1 ≡ 1 [p].

From Ex. 5.27, where p = a2 + b2 , a odd, we know that


ab p−1
f 2 ≡2 4 ≡ 1 [p].

Since f 2 ≡ −1 [p], the order of f modulo p is 4, thus 4 | ab


2 , so 8 | ab.
As a is odd, 8|b, then p = A2 + 64B 2 (with A = a, B = b/8).

Conversely, if p = A2 + 64B 2 , then p ≡ A2 ≡ 1 [4].


Let a = A, b = 8B. Then
p−1 ab
2 4 ≡f 2 ≡ f 4AB ≡ (−1)2AB ≡ 1 [p].
p−1
As 2 4 ≡ 1 [p], x4 ≡ 2 [p] has a solution in Z (Prop. 4.2.1) : 2 is a biquadratic
residue modulo p.
Conclusion :

26
∃A ∈ Z, ∃B ∈ Z , p = A2 + 64B 2 ⇐⇒ (p ≡ 1 [4] and ∃x ∈ Z, x4 ≡ 2 [p]).

Note : the equation x4 ≡ 2 [p] has also solutions if p ≡ −1 [8].


p−1
Indeed, the equation x4 ≡ 2 [p] has a solution in Z iff 2 d = 1, where d = 4∧(p−1) =
p−1
2, thus iff 2 2 ≡ 1 [p], which is true since p2 = 1.


For instance, 84 ≡ 2 (mod 23), with 23 ≡ −1 (mod 8).

Ex. 5.29 Let (RR) be the number of pairs (n, n + 1) in the set 1, 2, 3, . . . , p − 1 such
that n and n + 1 are both quadratic residues modulo p. Let (N R) be the number of pairs
(n, n + 1) in the set 1, 2, 3, . . . , p − 1 such that n is a quadratic nonresidue and n + 1 is a
quadratic residue. Similarly, define (RN ) and (N N ). Determine the sums (RR)+(RN ),
(N R) + (N N ), (RR) + (N R), and (RN ) + (N N ).

Proof. Let E the set of pairs (n, n + 1) ∈ N2 , 1 ≤ n ≤ p − 2 : |E| = p − 2.


Write RR the set of pairs (n, n + 1) such that n and n + 1 are both a quadratic
residues, and (RR) = |RR| its cardinality, and similar definitions for RN, N R, N N .
As E = RR ∪ RN ∪ N R ∪ N N (disjoint union),

(RR) + (RN ) + (N R) + (N N ) = |E| = p − 2.

• RR ∪ RN is the set of pairs (n, n + 1) in E such that n is a residue. Its cardinality


is the number of residues in [1, p − 2], so the number of residues in [1, p − 1], minus s,
where s = 1 if p − 1 is a residue, s = 0 otherwise. In both cases p ≡ 1, 3 (mod 4),
p−1
1+(−1) 2
s= 2 , and the total number of residues is (p − 1)/2, so
p−1
p−1 p − 1 1 + (−1) 2 1 p−1
(RR) + (RN ) = −s= − = (p − 2 − (−1) 2 ).
2 2 2 2
• Similarly, (N R) + (N N ) is the number of nonresidues in [1, p − 1], minus t, where
p−1
1−(−1) 2
t = 1 if p − 1 is a nonresidue, t = 0 otherwise : t = 2 , so

1 p−1
(N R) + (N N ) = (p − 2 + (−1) 2 )
2
(the sum of these two results is indeed p − 2 = |E|).
• As 1 is a residue, (RR) + (N R) is the number of residues in [1, p − 1], minus 1 :
p−1
(RR) + (N R) = − 1.
2
• (RN ) + (N N ) is the number of nonresidues in [2, p − 1], equal to the number of
residues in [1, p − 1] :
p−1
(RN ) + (N N ) = .
2

27
Ex. 5.30 Show that (RR) + (N N ) − (RN ) − (N R) = p−1
P
n=1 (n(n + 1)/p). Evaluate
this sum and show that it is equal to −1. (Hint : The result of Exercise 8 is useful.)

Proof. Let χ be the characteristic function of RR ∪ N N : if 1 ≤ n ≤ p − 1, χ(n) = 1 if


n, n + 1 are both residues, or if n, n + 1 are both non residues. Then
   
1 n n+1
χ(n) = 1+
2 p p

(if χ(n) = 1, np n+1 = 1, and np n+1


   
p p = −1 otherwise.)
Similarly, let χ0 the characteristic function of RN ∪ N R : χ0 (n) = 1 if exactly one of
the integer n, n + 1 is a residue, 0 otherwise. Then
   
0 1 n n+1
χ (n) = 1− .
2 p p

As each integer n between 1 and p − 1 brings the contribution 1 if n ∈ RR ∪ N N , and


−1 if n ∈ RN ∪ N R, then
p−1
X
(RR) + (N N ) − (RN ) − (N R) = (χ(n) − χ0 (n))
n=1
p−1      
1 X n(n + 1) n(n + 1)
= 1+ − 1−
2 p p
n=1
p−1  
X n(n + 1)
=
p
n=1

To evaluate this sum S, note that 4n(n + 1) = (2n + 1)2 − 1, so


p−1  p−1  p−1 
(2n + 1)2 − 1
  
X n(n + 1) X 4n(n + 1) X
S= = = .
p p p
n=1 n=1 n=1

This sum can be written S = n∈F∗p ((2n + 1)2 − 1)/p) = n∈Fp ((2n + 1)2 − 1)/p), since
P P

(0/p) = 0. As f : Fp → Fp , n 7→ (2n + 1) is a bijection (2 is invertible in F∗p ),

X (2n + 1)2 − 1 X y 2 − 1
= (y = 2n + 1).
p p
n∈Fp y∈Fp

As p - 1, the evaluation of this last sum is given in Exercise 5.8 : S = −1, so


p−1  
X n(n + 1)
(RR) + (N N ) − (RN ) − (N R) = = −1.
p
n=1

Ex. 5.31 Use the results of Exercises 29 and 30 to show that (RR) = 14 (p − 4 − ε),
where ε = (−1)(p−1)/2

28
Proof. To summarize the results of the Ex. 5.29 and 5.30,

(a)(RR) + (RN ) + (N R) + (N N ) = p − 2
(b)(RR) + (N N ) − (RN ) − (N R) = −1

and
1 p−1

(c)(RR) + (RN ) = p − 2 − (−1) 2
2
p−1
(d)(RR) + (N R) = −1
2

The sum of (a) and (b) gives


p−3
(e)(RR) + (N N ) = .
2
The sum of (c),(d),(e) gives (using (a))
p−1
p−2 p−1 p−3 (−1) 2
2(RR) + p − 2 = + + −1− ,
2 2 2 2
so p−1 p−1
p−1 p−3 p−2 (−1) 2 p (−1) 2
2(RR) = + − −1− = −2− ,
2 2 2 2 2 2
1 p−1
(RR) = (p − 4 − ε), where ε = (−1) 2 .
4

Q(p−1)/2
Ex. 5.32 If p is an odd prime, show that (2/p) = j=1 2 cos(2πj/p). Use this to
give another proof to Proposition 5.1.3.

Proof. Let p an odd prime number, and ζ = e2iπ/p : then ζ p = 1.


Let
p−1
Y p−1
Y
P = (ζ j + ζ −j ) = 2 cos(2πj/p).
j=0 j=0

p−1
Y
P = ζ 0 ζ −1 · · · ζ −(p−1) (ζ 2j + 1)
j=0
p−1
Y
= (ζ p )−(p−1)/2 (ζ 2j + 1)
j=0
p−1
Y
= (ζ 2j + 1)
j=0

As ζ j depends only of the class j ∈ Fp , this product can be written


Y Y
P = (ζ 2j + 1) = (ζ k + 1) (k = 2j),
j∈Fp k∈Fp

29
since f : Fp → Fp , x 7→ 2x is a bijection. So
p−1
Y
P = (ζ k + 1).
k=0

Since ζ 0 = 1, ζ, · · · , ζ p−1 are the roots of the polynomial f (x) = xQ


p −1, then 1+ζ 0 , · · · , 1+

ζ p−1 are the roots of g(x) = (x − 1)p − 1 = f (x − 1), so g(x) = p−1 k


k=0 (x − (1 + ζ )).
p−1
As g(0) = (−1)p − 1 = −2 = (−1 − ζ 0 ) · · · (−1 − ζ p−1 ) = − k=0 (ζ k + 1), we obtain
Q

p−1
Y p−1
Y
P = 2 cos(2πj/p) = (ζ k + 1) = 2,
j=0 k=0

so
p−1
Y
2 cos(2πj/p) = 1.
j=1

p−1
Y
1= 2 cos(2πj/p)
j=1
(p−1)/2 p−1
Y Y
= 2 cos(2πj/p) 2 cos(2πj/p)
j=1 j=(p+1)/2
(p−1)/2 (p−1)/2
Y Y
= 2 cos(2πj/p) 2 cos(2π − 2πk/p) (k = p − j)
j=1 k=1

As cos(2π − α) = cos(α),
 2
(p−1)/2 (p−1)/2
Y Y
1= 2 cos(2πj/p) , so 2 cos(2πj/p) = ±1
j=1 j=1

Case 1 : if 1 ≤ j ≤ p/4, 0 ≤ 2πj/p < π/2, so cos(2πj/p) > 0,


case 2 : if p/4 < j ≤ (p − 1)/2, π/2 < 2πj/p < π, so cos(2πj/p) < 0.
In the first case, 2 ≤ 2j ≤ (p − 1)/2 : the least residue of 2j is positive. In the second
case p/2 < 2j ≤ p − 1 : the least residue of 2j is negative.
Let µ the number of negative least residues of the integer 2j, 1 ≤ j ≤ (p − 1)/2 : we
know from Gauss’ Lemma that (2/p) = (−1)µ . As µ is also the number of j, 1 ≤ j ≤
(p − 1)/2 such that cos(2πj/p) < 0,
(p−1)/2  
Y 2 µ
2 cos(2πj/p) = (−1) = .
p
j=1

If p ≡ 1 [8] , p = 8q + 1, q ∈ N. For 1 ≤ j ≤ (p − 1)/2,

cos(2πj/p) < 0 ⇐⇒ p/4 ≤ j ≤ (p − 1)/2 ⇐⇒ 2q + 1 ≤ j ≤ 4q,


so µ = 2q and (2/p) = (−1)µ = 1.

30
If p ≡ −1 [8], p = 8q − 1, q ∈ N∗ .

cos(2πj/p) < 0 ⇐⇒ p/4 ≤ j ≤ (p − 1)/2 ⇐⇒ 2q ≤ j ≤ 4q − 1,

so µ = 2q and (2/p) = (−1)µ = 1.


If p ≡ 3 [8], p = 8q + 3, q ∈ N.

cos(2πj/p) < 0 ⇐⇒ p/4 ≤ j ≤ (p − 1)/2 ⇐⇒ 2q + 1 ≤ j ≤ 4q + 1,

so µ = 2q + 1 and (2/p) = (−1)µ = 1.


If p ≡ −3 [8], p = 8q − 3, q ∈ N∗ ,

cos(2πj/p) < 0 ⇐⇒ p/4 ≤ j ≤ (p − 1)/2 ⇐⇒ 2q ≤ j ≤ 4q − 2,

so µ = 2q − 1 and (2/p) = (−1)µ = 1.

Ex. 5.33 Use Proposition 5.3.2 to derive the quadratic character of −1.

Proof. Let f (z) = e2πiz − e−2πiz . If p is an odd prime, a ∈ Z, and p - a, we know from
Prop. 5.3.2 that
(p−1)/2     (p−1)/2  
Y la a Y l
f = f .
p p p
l=1 l=1

For a = −1, as f (−z) = −f (z),


  (p−1)/2   (p−1)/2  
−1 Y l Y −l
f = f
p p p
l=1 l=1
(p−1)/2  
Y l
= (−1)(p−1)/2 f
p
l=1

(z) = 0 ⇐⇒ e4πiz = 1 ⇐⇒ 4πiz = 2kiπ, k ∈ Z ⇐⇒ z = k/2, k ∈ Z, so, if


Moreover f 
l ∈ Z, f pl = 0 ⇐⇒ l/p = k/2, k ∈ Z ⇐⇒ p | 2l ⇐⇒ p | l. For 1 ≤ l < p, this is
Q(p−1)/2  l 
impossible, so l=1 f p 6= 0. Consequently,
 
−1
= (−1)(p−1)/2 .
p

Ex. 5.34 If p is an odd prime distinct from 3, show that


  (p−1)/2
Y   
3 2 2πj
= 3 − 4 sin .
p p
j=1

31
Proof. Let p an odd prime number, p 6= 3 and ζ = e2iπ/p .
2
ζ j − ζ −j
  
2 2πj
3 − 4 sin =3−4
p 2i
= 3 + ζ 2j + ζ −2j − 2
= 1 + ζ 2j + ζ −2j
 
4πj
= 1 + 2 cos
p

(As cos(2α) = 1 − 2 sin2 α, so 3 − 4 sin2 α = 1 + 2 cos α.)


Let
p−1
Y   Y  
2 2πj 2 2πj
P = 3 − 4 sin = 3 − 4 sin .
p p
j=1 ∗ j∈Fp

Then
Y
1 + ζ 2j + ζ −2j

P =
j∈F∗p
Y  
= 1 + ζ k + ζ −k (k = 2j)
k∈F∗p

since f : F∗p → F∗p , j 7→ 2j is a bijection. So

p−1
Y  
P = ζ −k 1 + ζ k + ζ 2k
k=1
p−1 Qp−1
Y
−k (1 − ζ 3k )
= ζ Qk=1
p−1 k
k=1 k=1 (1 − ζ )

= (ζ p )−(p−1)/2 = 1. Moreover, p−1


Qp−1 −k
Q 3k
Qp−1 k
k=1 ζ k=1 (1 − ζ ) = k=1 (1 − ζ ), since k 7→ 3k is

a bijection in Fp , so P = 1, and consequently

p−1
Y  
2 2πj
1= 3 − 4 sin
p
j=1
(p−1)/2    p−1   
Y
2 2πj Y
2 2πj
= 3 − 4 sin 3 − 4 sin
p p
j=1 j=(p+1)/2
(p−1)/2   (p−1)/2

Y   
2πj 2 2π(k − j)
Y
2
= 3 − 4 sin 3 − 4 sin (k = p − j)
p p
j=1 k=1
 2
 
(p−1)/2  
Y 2πj
= 3 − 4 sin2 
p
j=1

Q(p−1)/2   
So j=1 3 − 4 sin2 2πj
p = ±1.
Let ν the number of negative factors in this product.

32
If 1 ≤ j ≤ (p − 1)/2, then 0 < 4πj/p < 2π.

4πj 4πj 2π
1 + 2 cos < 0 ⇐⇒ cos < cos
p p 3
2π 4πj 4π
⇐⇒ < <
3 p 3
p p
⇐⇒ <j<
6 3
p
⇐⇒ < 3j < p
2
Let µ the number of integers j, 1 ≤ j ≤ (p − 1)/2 such that their least remainder is
negative. Since 3 ≤ 3j ≤ 3(p − 1)/2 and 3j 6= p/2, these j are such that p2 < 3j < p, so
µ = ν. Therefore
(p−1)/2     
Y
2 2πj ν µ 3
3 − 4 sin = (−1) = (−1) = .
p p
j=1

Ex. 5.35 Use the preceding exercise to show that 3 is a square modulo p iff p is
congruent to 1 or −1 modulo 12.

Proof. We know from Ex. 5.34 that ν = Card {j ∈ [1, (p − 1)/2] | p/2 ≤ 3j < p} = µ.
Therefore ν is the number of j such that p/6 ≤ j < p/3, so ν = bp/3c − bp/6c.
If p = 12k + 1, ν = bp/3c − bp/6c = 4k − 2k = 2k : (3/p) = (−1)ν = 1
If p = 12k + 5, ν = bp/3c − bp/6c = 4k + 1 − 2k = 2k + 1 : (3/p) = (−1)ν = −1
If p = 12k − 5, ν = bp/3c − bp/6c = 4k − 2 − (2k − 1) = 2k − 1 : (3/p) = (−1)ν = −1
If p = 12k − 1, ν = bp/3c − bp/6c = 4k − 1 − (2k − 1) = 2k : (3/p) = (−1)ν = −1
3 is a square modulo p (p 6= 2, p 6= 3) iff p is congruent to 1 or −1 modulo 12.

Ex. 5.36 Show that part (c) of Proposition 5.2.2 is true if a is negative and b is positive
(both still odd).

As said by Adam Michalik, the Jacobi symbol ab only defined for positive b, so the


question, which concerns ab ab , a < 0 makes no sense.


 

To give sense to this question, we must substitute the Kronecker symbol to the Jacobi
symbol. The Kronecker symbol (not defined in Ireland-Rosen) is the usual extension
of Jacobi symbol (see for instance [Henri Cohen] A course in computational algebraic
number theory, [Henri Cohen] Number theory (vol. 1), or [Harvey Cohn] Advanced
number theory).
We define Kronecker (or Kronecker-Jacobi) symbol ab for any a and b in Z in the


following way.

(1) If b = 0, then a0 = 1 if a = ±1, and a0 = 0 otherwise.


 

Q
(2) For b 6= 0, write b = p, where the p are not necessarily distinct primes (including
2), or p = −1 to take care of the sign. Then we set
  Y 
a a
= ,
b p

33
a

where p is the Legendre symbol defined above for p > 2, and where we define
  (
a 0 if a is even
= 2
2 (−1)(a −1)/8 if a is odd,

and also   (
a 1 if a ≥ 0
=
−1 −1 if a < 0

Proof. Suppose a < 0, b > 0, both odd. Let a = −A, A > 0, A = p1 p2 · · · pk , where the
pi are not necessarily distinct primes. Then
     
a b −A b
=
b a b −A
      
−A −1 A (b−1)/2 A
= = (−1)
b b b b
        
b b b b b
= ··· =
−A −1 p1 pk A

so, from Prop. 5.2.2, as A, b are odd and positive,


     
a b b−1 A b
= (−1) 2
b a b A
b−1 A−1 b−1
= (−1) 2 (−1) 2 2

b−1
[1+ −a−1 ]
= (−1) 2 2

b−1 1−a
= (−1) 2 2

b−1 a−1
= (−1) 2 2

So the law of quadratic reciprocity remains valid for the Kronecker symbol when a is
negative (b > 0, a, b both odd).

Ex. 5.37 Show that if a is negative, then p ≡ q (mod 4a), p - a implies (a/p) = (a/q).

Proof. Write a = −A, A > 0. As p ≡ q (mod 4a), we know from Prop. 5.3.3. (b) that
(A/p) = (A/q).
Moreover,
     
a −A (p−1)/2 A
= = (−1)
p p p
     
a −A A
= = (−1(q−1)/2
q q q

As p ≡ q (mod 4a), p = q + 4ak, k ∈ Z, so

(−1)(p−1)/2 = (−1)(q+4ak−1)/2 = (−1)(q−1)/2 ,

so (a/p) = (a/q).

34
Ex. 5.38 Let p be an odd prime. Derive the quadratic character of 2 modulo p by
verifying the following steps, involving the Jacobi symbol:
         
2 8−p p 8 2
= = = = .
p p p−8 p−8 p−8
Generalize the argument to show that
   
a a
= , a > 0, p - a.
p p − 4a

(As in Ex. 5.36, since 8 − p or p − 8 is negative, we interpret (a/b) as the Kronecker


symbol : see definition in Ex. 5.36.)

Proof. As (22 /p) = 1 and 8 − p ≡ 8 (mod p),


   2      
2 2 2 8 8−p
= = = .
p p p p p

As p and 8 − p are odd numbers and p > 0, from the extension of the law of quadratic
reciprocity to a < 0 proved in Ex. 5.36, we obtain
   
8−p 7−p p−1 p
= (−1) 2 2 .
p 8−p
Moreover

(7 − p)(p − 1) ≡ (−1 − p)(p − 1) = 1 − p2 (mod 8)

As p = 2k + 1 is odd, p2 = 4k 2 + 4k + 1 = 8 k(k+1)
2 + 1 ≡ 1 (mod 8), so (7 − p)(p − 1) ≡ 0
(mod 8) and 7−p 2
p−1
2 is even, so
   
8−p p
= .
p 8−p
p  p  p  p  p 
As p > 0, −1 = 1, thus 8−p = −1 p−8 = p−8 (with the same argument, this is
also true for the 3 odd primes such that 8 − p > 0), so
   
8−p p
= .
p p−8

p  8 8 2
, and since 8 = 22 × 2, p−8
  
As p ≡ 8 (mod p − 8), p−8 = p−8 = p−8 . We have
proved for all odd primes p that
         
2 8−p p 8 2
= = = = .
p p p−8 p−8 p−8

The preceding arguments remain valid if we replace the odd prime p by any odd
positive integer. So with an immediate induction, we see that for all k ∈ N,
   
2 2
= .
p p − 8k

35
So the quadratic character
 of 2 modulo p depends only of the class of p modulo 8.
If p ≡ 1 (mod 8), p2 = 21 = 1.
If p ≡ −1 (mod 8), p2 = −1 2
 
= 1.
2 2
 
If p ≡ ±3 (mod 8), p = ±3 = −1.

Generalization : let a > 0 and p an odd positive integer such that p ∧ a = 1 (not
necessarily prime).
       
a 4ap 4a − p 4a−p−1 p−1 p
= = = (−1) 2 2 .
p p p 4a − p
(4a − p − 1)(p − 1) = 4a(p − 1) + 1 − p2 ≡ 0 (mod 8), so
   
a p
= .
p 4a − p
p 
As −1 = 1,
   
p p
= .
4a − p p − 4a
Since p ≡ 4a (mod p) − 4a, and 4 is a square,
     
p 4a a
≡ = .
p − 4a p − 4a p − 4a

We have proved
         
a 4a − p p 4a a
= = = = .
p p p − 4a p − 4a p − 4a

By induction, for all k ≥ 0, ap = p−4a


a
, so ap depends only of the class of p modulo
  

4a.

Chapter 6
√ √
Ex. 6.1 Show that 2 + 3 is an algebraic integer.
√ √ √
Proof. Let x = 2√+ 3. Then x2 = 5 + 2 6.
(x2 − 5)2 = (2 6)2 = 24, so x4 − 10x2 + 1 = 0 : x is an algebraic integer.

Ex. 6.2 Let α be an algebraic number. Show that there’s an integer n such that nα is
an algebraic integer.
(0 is a valid answer to this sentence ! More seriously, we search a positive integer n.)

Proof. Let α an algebraic number. By definition, there exist a0 , a1 , · · · , an ∈ Z, an 6= 0,


such that
an αn + an−1 αn−1 + · · · + ak αk + · · · + a0 = 0.
(Up to multiply this equation by −1, we can suppose that an > 0).
Multiplying by an−1
n , we obtain

ann αn + an−1
n an−1 α
n−1
+ · · · + ann−1 ak αk + · · · + ann−1 a0 = 0.

36
So
(an α)n + an−1 (an α)n−1 + · · · + ann−k−1 ak (an α)k + · · · + an−1
n a0 = 0.
n−1
Soit p(x) = xn + an−k−1 ak xk . Then p(x) ∈ Z[x], p(x) is monic, and p(an α) = 0.
P
n
k=0
So an α is an algebraic integer, with m = an ∈ N∗ .
Conclusion : if α is an algebraic number, there exists an integer m > 0 such that mα
is an algebraic integer.

Ex. 6.3 If α and β are algebraic integers, prove that any solution to f (x) = x2 + αx +
β = 0 is an algebraic integer. Generalize this result.

Proof. Let γ a root of x2 + αx + β, where α, β verify :

αn + r1 αn−1 + · · · + rn = 0, ri ∈ Z,

β m + s1 β m−1 + · · · + sm = 0, sj ∈ Z.
Let V the set of linear combinations with integer coefficients of

αi β j γ k , 0 ≤ i < n, 0 ≤ j < m, 0 ≤ k < 2.

Then V if a finitely generated Z-module.


Moreover, for all δ ∈ V, γδ ∈ V . Indeed, every δ ∈ V is a linear combination with
coefficients in Z of αi β j , αi β j γ, and

γ(αi β j ) = αi β j γ ∈ V
γ(αi β j γ) = αi β j γ 2 = αi β j (−αγ − β) = −αi+1 β j γ − αi β j+1 ∈ V.
n−1
(if i + 1 = n, we replace αi+1 = αn by − rk αn−k , and a similar replacement if if
P
k=1
j + 1 = m.)
As for each x ∈ V , where V if a finitely generated Z-module, xγ ∈ V , so γ is an
algebraic integer (Proposition 6.1.4).
More generally, if γ n + α1 γ n−1 + · · · + αn = 0, where the αi are algebraic integers,
then x is an algebraic integer.

Ex. 6.4 A polynomial f (x) ∈ Z[x] is said to be primitive if the greatest common divisor
of its coefficients is 1. Prove that product of primitive polynomials is also primitive.

Solution 1

Proof. Let p(x) = ni=0 ai xi , q(x) = m j


P P
j=0 bj x two primitive polynomials, and p a prime
number. There exist a coefficient of p(x) (and of q(x)) not divisible by p. Let

i0 = min{i ∈ [0, n] | ai 6≡ 0 [p]}


j0 = min{j ∈ [0, m] | bj 6≡ 0 [p]}

Pn+m
ck xk . Then ck =
P
Let p(x)q(x) = k=0 ai bj , k = 0, . . . n + m. Then
i+j=k
X
ci0 +j0 = ai bj .
i+j=i0 +j0

37
• If i < i0 , then ai ≡ 0 (mod p).
• If i > i0 , then j < j0 and bj ≡ 0 (mod p).
In the two cases ai bj ≡ 0 (mod p), so ci0 +j0 ≡ ai0 bj0 (mod p), so cj0 6≡ 0 (mod p) :
as it’s true for all primes p, the polynomial p(x)q(x) is primitive.

Solution 2

Proof. Let

Z[x] → Fp [x]
ϕ:
p(x) = a0 + · · · + an xn 7→ p(x) = a0 + · · · + an xn ,

where ai is the class of ai in Fp . ϕ is a ring homomorphism.


As Fp [x] is an integrity domain, if p(x), q(x) are both primitive,

p(x) 6= 0, q(x) 6= 0 ⇒ p(x)q(x) = p(x) q(x) 6= 0.

As p(x)q(x) 6= 0 in all fields Fp , p(x)q(x) is a primitive polynomial.

Ex. 6.5 Let α be an algebraic integer and f (x) ∈ Q[x] be the monic polynomial of least
degree such that f (α) = 0. Use Exercise 6.4 to show that f (x) ∈ Z[x].

Proof. As α is an algebraic integer, there exists a monic polynomial h(x) ∈ Z[x] such
that h(α) = 0. As f (x) ∈ Q[x] is the minimal polynomial of α, and h(α) = 0, f (x)
divides h(x) in Q[x].
(quick reminder : h(x) = q(x)f (x) + r(x), q(x), r(x) ∈ Q[x], deg(r(x)) < deg(f (x))
or r(x) = 0. As r(α) = 0 and f (x) ∈ Q[x] is the monic polynomial of least degree such
that f (α) = 0, r = 0 so f (x) | h(x)).
So there exists g(x) ∈ Q[x] such that h(x) = f (x)g(x). As h(x), f (x) are both monic,
g(x) is also monic.
Let d ∈ Z, d 6= 0 such that df (x) = m i
P
i=0 ai x ∈ Z[x], and c = a1 ∧ a2 ∧ · · · ∧ am ,
c
ai = cbi , with b1 ∧ b2 ∧ · · · ∧ bm = 1, so f (x) = d f1 (x), with f1 is primitive. Similarly
g(x) = st g1 (x), s, t ∈ Z, g1 (x) primitive.
Pr So h(x) = cs u
dt f1 (x)f2 (x) = v f1 (x)f2 (x), where u∧v = 1. The polynomial f1 (x)f2 (x) =
k
k=0 ck x is primitive (Ex. 6.4). As vh(x)(x) = uf1 (x)f2 (x), v | uck , and u ∧ v = 1,
thus v | ck , k = 0, 1, . . . , r. As c1 ∧ · · · ck = 1, v | 1, so v = ±1. h(x) = uf1 (x)f2 (x)
is monic, thus u = 1, and f1 , f2 are monic. From f (x) = dc f1 (x) we deduce dc = 1 and
f (x) = f1 (x) ∈ Z[x].
Conclusion : if f (x) is the minimal polynomial of an algebraic integer α, f ∈ Z[x].

Ex. 6.6 Let x2 + mx + n ∈ Z[x] be irreducible, and α be a root. Show that Q[α] =
{r + sα : r, s ∈ Q} is a ring (in√fact, it is a field). Let m2 − 4n = D02 D, where D is
square-free. Show that Q[α] = Q[ D].

Proof. By definition, for all z ∈ C, z ∈ Q[α] ⇐⇒ ∃P ∈ Q[x], z = P (α).


The Euclidean division gives P = Q1 (x2 + mx + n) + R, Q1 , R ∈ Q[x], deg(R) < 2,
so R = rx + s, r, s ∈ Q. So z = Q1 (α)(α2 + mα + n) + rα + s = rα + s :

Q[α] = {z ∈ C | ∃r ∈ Q, ∃s ∈ Q, z = r + sα}.

38
• Q[α] ⊂ C, where (C, +, ×) is a field. 1 ∈ Q[α] (1 = P0 (α), where P0 is the constant
polynomial 1).
• Let β, γ ∈ Q[α] : β = P (α), γ = Q(α), where P, Q are in Q[x]. Then α − β =
P (α) − Q(α) = R(α), where R = P − Q ∈ Q[x], and αβ = P (α)Q(α) = S(α), where
S = P Q ∈ Q[x]. So α − β ∈ Q[α], αβ ∈ Q[α]. So Q[α] is a subring of (C, +, ×).
• Let β = P (α) ∈ Q[α], P ∈ Q[x] and β 6= 0. As β 6= 0, Q = x2 + mx + n - P .
Let D ∈ Q[x] such that D | P, D | Q. As Q is irreducible by hypothesis, D = λ or
D = λQ, λ ∈ C∗ (D is an associate of 1 or Q). If D = λQ, then Q | D, and D | P , so
Q | P . Since Q(α) = 0, this implies β = P (α) = 0, in contradiction with the definition
of β. So D = λ | 1. Therefore P ∧ Q = 1.
From Bézout’s theorem, there exist polynomials U, V ∈ Q[x] such that U P +V Q = 1.
As Q(α) = 0, U (α)P (α) = 1 and γ = U (α) ∈ Q[α] is such that γβ = 1. Therefore Q[α]
is a subfield of (C, +, ×) (and Q(α) = Q[α]).
As x2 + mx + n is irreducible, ∆ = m2 − 4n 6= 0 (if not, x2 + mx + n = (x +
m/2)2 − (m2 − 4n)/4 = (x + m/2)2 is not irreducible). So ∆ ∈ Z \ {0} can be written
∆ = m2 − 4n =√D02 D, where D is square-free (positive or negative), D 6= 0, D0 6= 0.
√ √ √
α = −m + ε ∆
, ε = ±1, so α = − m
+ εD 0 2
D
, thus α ∈ Q[ D] and Q[α] ⊂ Q[ D].
2 √2 2 √
As D0 6= 0, D = ε 2α+m D0 ∈ Q[α], so Q[ D] ⊂ Q[α)] :

Q[α] = Q[ D].

Ex.√ 6.7 (continuation) If√D ≡ 2, 3 (mod 4), show that all the algebraic integers in
Q[ D] have the form a√+ b D, where a, b ∈ Z. If D √ ≡ 1 (mod 4), show that all the
algebraic integers in Q[ D] have the form a + b((−1 + D)/2), where a, b ∈ Z.

Proof. (We write Z the ring of algebraic integers in C, and OK (or ZK ) the ring of
algebraic integers√in the field K.) √
If D = 1, Q[ D] = Q. If D 6= 1, as D is square-free, D is not a square, so D is
irrational. √ √ √
Let γ = r + s D ∈ Q[ D] (r, s ∈ Q) an algebraic integer of Q[ D] (D ∈ Z, D
square-free). (γ − r)2 = s2 D, so γ 2 − 2rγ + r2 − Ds2 = 0. γ is a root of

p(x) = x2 − 2rx + r2 − Ds2 .

If s = 0, then the minimal polynomial of γ is x − r. As r = γ is an algebraic integer


and r ∈ Q, then r ∈ Z. In this case r ∈ Z and s = 0.
If s 6= 0, γ 6∈ Q, so no polynom of degree d ≤ 1 has the root γ. Thus the minimal
polynomial of γ is p(x). From Exercise 6.5, p(x) ∈ Z[x], so (in the two cases s = 0, s 6= 0)

2r ∈ Z, r2 − Ds2 ∈ Z.

Conversely, if 2r ∈ Z, r2 − Ds2 ∈ Z, then p(x) ∈ Z[x] and p(γ) = 0, thus γ is an algebraic


integer.
If r, s ∈ Q, D 6= 1 square-free,

r + s D ∈ Z ⇐⇒ 2r ∈ Z, r2 − Ds2 ∈ Z.

Let γ = r + s D ∈ Z. We can write
a b
r= ,s = , a, b, d ∈ Z, d ≥ 1, d ∧ a ∧ b = 1.
d d

39
Then
2a a2 − Db2
n= ∈ Z, m= ∈ Z.
d d2
As D is square-free, D ≡
6 0 (mod 4).

• Case 1 : D ≡ 2, 3 (mod 4).


4Db2
n2 − 4m = d2
, so d | 2a, d2 | 4Db2 .
If 2 | d, 4 | a2 − Db2 , a2 ≡ Db2 (mod 4). As d ∧ a ∧ b = 1, and 2 | d, a or b is
odd, and a2 ≡ Db2 (mod 4), D 6≡ 0 (mod 4), implies that a and b are both odd.
Then a2 ≡ b2 ≡ 1 (mod 4), so D ≡ 1 (mod 4) : this is in contradiction with the
hypothesis D ≡ 2, 3 (mod 4). So d is an odd number.
Consequently, d | a, d2 | Dq 2 . If p ∈ N is a prime factor of d, p | d, p | a ,and
d ∧ a ∧ b = 1, so p - b, and since p2 | Db2 , p2 | D, in contradiction with D square-
free. So d√≥ 1 has no prime factor : d = 1 and r = a, s = b ∈ Z. Conversely, any
γ = a + b D, a, b ∈ Z is an algebraic integer, so
√ √
OQ[√D] = Z ∩ Q[ D] = {a + b D | a, b ∈ Z}.

• Case 2 : D ≡ 1 (mod 4).


Then r = n2 , n ∈ Z. Write s = uv , u ∧ v = 1, v ≥ 1.
2 2 2
m = r2 − Ds2 = n4 − D uv2 ∈ Z, 4D uv2 = n2 − 4m ∈ Z, so v 2 | 4Du2 . Since
u ∧ v = 1, u2 ∧ v 2 = 1, so v 2 | 4D. As D is square-free, v has no odd prime factor,
so v = 2k . Since D is odd, k ≤ 1 and v = 1 or v = 2. So r, s are both half-integers:
r = n/2, s = n0 /2, n, n0 ∈ Z.
4m = n2 − Dn02 , thus n2 ≡ n02 (mod 4), so n, n0 have the same parity. √ Let
0
a = n+n ∈ Z, b = n 0 ∈ Z. Then n = 2a − b, n0 = b and γ = n + n0 D =
2 2 2
b b
√  √ 
−1+ D
a− 2 + 2 D =a+b 2 .
√ 2
 √ 
Conversely, −1+2 D is a root of x2 + x + 1−D
4 ∈ Z[x], so every a + b −1+ D
2 is
an algebraic integer.
√ !
√ −1 + D
OQ[√D] = Z ∩ Q[ D] = {a + b | a, b ∈ Z}.
2

Ex. 6.8 Let ω = e2πi/3 , ω satisfies x3 − 1 = 0. Show that (2ω + 1)2 = −3, and use this
to determine (−3/p) by the method of section 2.

Proof. As ω 2 + ω + 1 = 0,(2ω + 1)2 = 4ω 2 + 4ω + 1 = −4 + 1 = −3. Let α = 2ω + 1, so


α2 = −3
 
−3
≡ (−3)(p−1)/2 (mod p)
p
≡ αp−1(mod p)
 
p −3
α = α.
p

40
From Prop. 6.1.6,

αp = (2ω + 1)p
≡ 2p ω p + 1 (mod p)
p
≡ 2ω + 1 (mod p)
−3

• If p ≡ 0 (mod 3), p = 0.

• If p ≡ 1 (mod 3), ω p = ω, so αp ≡ α (mod p).


−3 −3 −3
  2 2

p α ≡ α (mod p), thus p  α ≡ α (mod p), p 3 ≡ 3 (mod p). As p ∧ 3 = 1,
−3 −3 −3

p ≡ 1 (mod p). Since p = ±1, p = 1.

• If p ≡ −1 (mod 3),

αp ≡ 2ω p + 1 (mod p)
2
≡ 2ω + 1 = 2(−1 − ω) + 1 = −2ω − 1 = −α (mod p).

−3
α ≡ −α (mod p), thus −3 α ≡ −α2 (mod p), −3
  2 
p p p 3 ≡ −3 (mod p). As
−3 −3 −3
  
p ∧ 3 = 1, p ≡ −1 (mod p). Since p = ±1, p = −1.

Conclusion :
 
−3
p ≡ 0[3] ⇐⇒ =0
p
 
−3
p ≡ 1[3] ⇐⇒ =1
p
 
−3
p ≡ −1[3] ⇐⇒ = −1
p

p
In other words, −3

p = 3 .
Note : α = 2ω + 1 = ω − ω 2 = g, the quadratic Gauss sum for p = 3.

Ex. 6.9 Verify Proposition 6.3.2 explicitly for p = 3, 5, i.e., write out the Gauss sum
longhand and square.

• p=3. Let ω = e2iπ/3 . Let g = 2t=0 (t/3)ω t the quadratic Gauss sum. Then
P
Proof.
g = ω − ω2.
As 1 + ω + ω 2 = 0, g 2 = (ω − ω 2 )2 = ω 2 − 2ω 3 + ω 4 = ω 2 − 2 + ω = −3 :

g 2 = −3.

• p=5. Let ζ = e2iπ/5 .


g = 4t=0 (t/3)ζ t = ζ − ζ 2 − ζ 3 + ζ 4 .
P

Then g = α − β, where α = ζ + ζ 4 , β = ζ 2 + ζ 3 .
α + β = ζ + ζ 4 + ζ 2 + ζ 3 = −1.
αβ = ζ 3 + ζ 4 + ζ 6 + ζ 7 = ζ 3 + ζ 4 + ζ + ζ 2 = −1

41
So α, β are the two roots of x2 + x − 1.

g 2 = (α − β)2
= α2 + β 2 − 2αβ
= (α + β)2 − 4αβ
= (−1)2 − 4(−1)
= 5.

Note : here we know explicitely g :



if p = 3, g = ω − ω 2 = i 3.
√ √ √
If p = 5, g = α − β = (−1 + 5)/2 − (−1 − 5)/2 = 5.

Pp−1
Ex. 6.10 What is a=1 ga ?

Proof. From Prop. 6.3.1 and Lemma 2,


p−1 p−1  
X X a
ga = g1 = 0.
p
a=1 a=1

+ (t/p))ζ t in two ways, prove that g = t2 .


P P
Ex. 6.11 By evaluating t (1 tζ

Proof. For a ∈ Fp , Write N [x2 = a] the number of solutions of the equation x2 = a in


Fp . We know from Ex. 5.2 that N [x2 = a] = 1 + (a/p). So
p−1
2 2
X X
ζt = ζt
t=0 t∈Fp
X
= N [x2 = t]ζ t
t∈Fp
X  
t
= 1+ ζt
p
t∈Fp
X X t
t
= ζ + ζt
p
t∈Fp t∈Fp
p−1  
X t t
= ζ
p
t=0
=g

42
Ex. 6.12 Write ψa (t) = ζ at . Show that

(a) ψa (t) = ψa (−t) = ψ−a (t)


P
(b) (1/p) a ψa (t − s) = δ(t, s)

Proof. (a) Let a ∈ Z. As ζ = ζ −1 ,

ψa (t) = ζ at = ζ −at
= ζ a(−t) = ζ (−a)t
= ψa (−t) = ψ−a (t)

ψa (t) = ψa (−t) = ψ−a (t)

(b) From Corollary of Lemma 1 :


p−1 p−1
1X 1 X a(t−s)
ψa (t − s) = ζ = δ(t, s)
p p
a=0 a=0

1X
ψa (t − s) = δ(t, s).
p a

Ex. 6.13 Let f be a function from Z to the complex numbers. Suppose that p is a
prime and that f (n + p) = f (n) for all n ∈ Z. Let fˆ(a) = p−1 t f (t)ψ−a (t). Prove that
P

f (t) = a fˆ(a)ψa (t). This result is directly analogous to a result in the theory of Fourier
P
series.

Proof. Let fˆ(a) = p−1 t f (t)ψ−a (t). Then


P

p−1
X p−1
X p−1
X
−1
fˆ(a)ψa (t) = p f (s)ψ−a (s)ψa (t)
a=0 a=0 s=0
p−1
X p−1
X
−1
=p f (s) ψ−a (s)ψa (t)
s=0 a=0
p−1
X p−1
X
= p−1 f (s) ψa (t − s)
s=0 a=0
p−1
X
= f (s)δ(s, t)
s=0
= f (t)

Ex. 6.14 In Ex. 13 take f to be the Legendre symbol and show that fˆ(a) = p−1 g−a .
Pp−1 t  −at
Proof. Here f (a) = ap . Then fˆ(a) = p−1 t=0 = p−1 g−a .

p ζ

43
Ex. 6.15 Show that
m  
X t √
< p log p.

p


t=n
The inequality holds for the sum over any range.
Lemma. If 0 ≤ x ≤ π2 , sin x ≥ π2 x.
Proof. As − sin is a convex function on [0, π/2], the graph of sin is above any chord, and
the chord between the points (0, 0) and (π/2, 1) has equation y = (2/π)x, we conclude
that sin x ≥ π2 x for 0 ≤ x ≤ π/2.

m √ Pm
t t
P  
Proof. Let S = p g with n ≤ m. Then |S| = p p . As (t/p)g = gt ,
t=n t=n
n
X
S= gt
t=m
n Xp−1  
X s ts
= ζ
t=m s=0
p
p−1   n
X s ms X (t−m)s
= ζ ζ
p t=m
s=0
p−1   n−m
X s ms X us
= ζ ζ (u = t − m)
p
s=0 u=0
p−1  
X s ms ζ (n−m+1)s − 1
= ζ
p ζs − 1
s=1

Pn−m s
ζ us = n − m + 1 and

(since for s = 0, the sum u=0 p = 0). So
p−1   (n+1)s
X s ζ − ζ ms
S=
p ζs − 1
s=1
p−1  n+m+1 n−m+1 −n+m−1
X s ζ 2 s ζ 2 s−ζ 2 s
= s s −s
p ζ2 ζ2 −ζ 2
s=1
 
p−1 π

X s n+m sin (n − m + 1)s p
= ζ 2 s  
p sin s π
s=1 p

As sin(x) ≥ π2 x for x ∈ [0, π2 ], for all s, 1 ≤ s < p2 , 0 ≤ sπ π


p ≤ 2 , so
 
sin (n − m + 1)s π
≤ 1  = p
p

  (s = 1, 2, . . . , (p − 1)/2).

sin s πp
2
s π 2s
π p

Since ps ζ ts depends only of the class of s, we can replace in the preceding calculation


the values s = 1, 2, . . . , p − 1 by s = −(p − 1)/2, . . . , −1, 1, . . . , (p − 1)/2, so


   
(p−1)/2   π −1 π
X s n+m s sin (n − m + 1)s p X  
s n+m s sin (n − m + 1)s p
S= ζ 2   + ζ 2   .
p sin s π p sin s π
s=1 p s=−(p−1)/2 p

44
As sin is an odd function,
   
(p−1)/2  π (p−1)/2  π
X 
s n+m s sin (n − m + 1)s p X 
−s − n+m s sin (n − m + 1)s p
S= ζ 2   + ζ 2   .
p sin s πp p sin s πp
s=1 s=1

Thus
(p−1)/2 (p−1)/2
X p X 1
|S| ≤ 2 =p .
2s s
s=1 s=1
m √
t
P 
As S = p g and |g| = p,
t=n
(p−1)/2

m  
X t √ X 1
≤ p .

p s


t=n s=1
It remains to do a sufficient estimation of the harmonic sum. We prove by induction
that for all n ≥ 1,
1 1
1 + + · · · + ≤ log(2n + 1).
2 n
As 1 ≤ log(3), this proposition is true for n = 1. Suppose that is it true for n − 1 :
1 1
1 + + ··· + ≤ log(2n − 1).
2 n−1
Then
1 1 1
1 + + · · · + ≤ + log(2n − 1).
2 n n
If we prove that n1 + log(2n − 1) ≤ log(2n + 1), the induction is done.
Let u(x) = log(2x + 1) − log(2x − 1) − x1 , x > 12 .
2 2 1
u0 (x) = − + 2
2x + 1 2x − 1 x
−4 1
= 2 +
4x − 1 x2
−1
= <0
(4x − 1)x2
2
 
2x+1
As u(x) = log 2x−1 − x1 , lim u(x) = 0. Moreover u is a decreasing function, so for
x→+∞
all x > 1/2, u(x) > 0, and for all n ∈ N, n ≥ 1,
1
+ log(2n − 1) ≤ log(2n + 1).
n
We have proved by induction that for all n ≥ 1,
1 1
1 + + · · · + ≤ log(2n + 1).
2 n
If n = (p − 1)/2, where p is an odd prime (p ≥ 3),
(p−1)/2
X 1
≤ log p.
s
s=1
Conclusion :
m  
X t √
< p log p.

p


t=n

45
Ex. 6.16 Let α be an algebraic number with minimal polynomial f (x). Show that f (x)
does not have repeated roots in C.

Proof. Let γ a repeated root of f (x). Then f (γ) = f 0 (γ) = 0, so x − γ is a common


factor of f and f 0 . Thus f ∧ f 0 6= 1 (deg(f ∧ f 0 ) ≥ 1). Since f ∧ f 0 | f and f is irreducible
(with f , f ∧ f 0 monic), we conclude f ∧ f 0 = f , so f | f 0 . In C, this is impossible since
deg(f ) ≥ 1, so f 0 6= 0, and deg(f 0 ) < deg(f ). f (x) does not have repeated roots in C.

Ex. 6.17 Show that the minimal polynomial for 3 2 is x3 − 2.

Proof. Let f (x) = x3 − 2. Then f ( 3 2) = 0. If f (x) was not irreducible, then f (x) =
u(x)v(x), with 1 ≤ deg(u) ≤ deg(v) ≤ 2, deg(u) + deg(v) = deg(f ) = 3, so deg(u) =
1, deg(v) = 2.
Then f (x) = (ax + b)(cx2 + dx + e), a, b, c, d, e ∈ Q. Let w = −b/a. Then f (w) =
3
w − 2 = 0 and w ∈ Q, so there exist p, q ∈ Z, such that w = p/q, p ∧ q = 1.
Thus p3 = 2q 3 , so p3 is even, thefore p is even : p = 2p0 , p0 ∈ Z.
8p03 = 2q 3 , 4p03 = q 3 , so q 3 is even, which implies that q is even. Then 2 | p ∧ q = 1 :
this is a contradiction.
√ √
So f ( 3 2) = 0, and f is monic, irreducible : f is the minimal polynomial of 3 2 on
Q.

Ex. 6.18 Show that there exist algebraic numbers of arbitrarily high degree.

Proof. As 1 + x + · · · + xp−1 is irreducible on Q[x] (Prop. 6.4.1), the numbers ζp = e2iπ/p ,


with p prime number, are algebraic numbers of arbitrary large degree.

Ex. 6.19 Find the conjugates of cos(2π/5).

Proof. Let γ = cos(2π/5), ζ = e2iπ/5 and α = ζ + ζ 4 , β = ζ 2 + ζ 3 .


−1 4
Then γ = ζ+ζ2 = ζ+ζ 2 = 2.
α
4 2
α + β = ζ + ζ + ζ + ζ = −1. 3

αβ = ζ 3 + ζ 4 + ζ 6 + ζ 7 = ζ 3 + ζ 4 + ζ + ζ 2 = −1
So α, β are the two roots of x2 + x − 1 :
α2 + α − 1 = 0, so 4(α/2)2 + 2(α/2) − 1 = 0 : γ = α/2 is a root of

f (x) = 4x2 + 2x − 1.

As ∆ = 4 × 5, the two roots of f are irrational. deg(f ) = 2 and f has no root in Q,


so f (x) is irreducible in Q[x]. Therefore the minimal polynomial of γ = cos(2π/5) is
f (x) = 4x2 + 2x − 1. The other root of f is β/2 = (ζ 2 + ζ 3 )/2 = cos(4π/5).
Conclusion : the conjugates of γ = cos(2π/5) are γ = cos(2π/5) and cos(4π/5).

Ex. 6.20 Let F be a subfield of C which is a finite-dimensional vector space over Q of


degree n. Show that every element of F is algebraic of degree at most n.

Proof. Let α ∈ F , with dimQ F = n. Any subset of n + 1 vectors in F is linearly


dependent, so {1, α, α2 , · · · , αn } is linearly dependent.
Thus there exists (a0 , . . . , an ) ∈ Qn+1 , (a0 , . . . , an ) 6= (0, 0, . . . , 0) such that a0 +a1 α+
· · · + an αn = 0.
Let f (x) = a0 + a1 x + · · · + an xn . Then f (x) ∈ Q[x], f (x) 6= 0 and f (α) =
0, deg(f (x)) ≤ n. So every element of F is algebraic of degree at most n.

46
Ex. 6.21 Let f (x) = ∞
P n
P∞ n
n=0 an x /n! and g(x) = n=0 bn x /n! be power series with
an and bn integers. If p is a prime such Pthat p|ani for i = 0, . . . , p − 1, show that each
coefficient ct of the product f (x)g(x) = ∞ n=0 cn x for t = 0, . . . , p − 1 may be written in
the form p(A/B), p 6 |B.

Proof. Let k ∈ N, 0 ≤ k ≤ p − 1.
X ai bj
ck =
i! j!
i+j=k
k
X ai bk−i
=
i! (k − i)!
i=0
k  
1 X k
= ai bk−i
k! i
i=0

As k! ∧ p = 1, and ki=0 i ai bk−i ≡ 0 (mod p) for k = 0, 1, . . . , p − 1,


k
P 

ck = p(A/B), p ∧ B = 1.

Ex. 6.22 Show that the relation ε ≡ 1 (mod p) in Proposition 6.4.4 can also be
achieved by replacing x by 1 + t instead of ez .

Proof. (solution given by Mikomikon and A.Grounds (agrounds))


We know from the remark after Prop 6.4.3 that
(p−1)/2
Y
g(χ) = ε (ζ 2k−1 − ζ −(2k−1) ),
k=1

where ε = ±1. Let


p−1 (p−1)/2
X Y
j
f (x) = χ(j)x − ε (x2k−1 − xp−(2k−1) ).
j=1 k=1

Then f (0) = 0 and f (ζ) = 0, so (xp − 1) divides f (x). As f (x) ∈ Z[x] and xp − 1 ∈ Z[x]
is monic, f (x) = (xp − 1)h(x), h(x) ∈ Z[x]. If we replace x by 1 + t, we obtain

p−1 (p−1)/2  
X Y
j
f (1 + t) = χ(j)(1 + t) − ε (1 + t)2k−1 − (1 + t)p−(2k−1) .
j=1 k=1

We compute the coefficient of t(p−1)/2 in the polynomial f (1 + t) :


p−1 p−1 j  
X X X j i
χ(j)(1 + t)j = χ(j) t
i
j=1 j=1 i=1
p−1 X
p−1  
X j i
= χ(j) t
i
i=1 j=i

47
p−1 p−1
j
So the coefficient of t(p−1)/2 in χ(j)(1 + t)j is
P P 
χ(j) (p−1)/2 .
j=1 j=(p−1)/2

(p−1)/2 (p−1)/2
Y Y
2k−1 p−(2k−1)
(1 + (2k − 1)t) − (1 + (p − (2k − 1))t + t2 u(t)

((1 + t) − (1 + t) )=
k=1 k=1
(p−1)/2
Y
(4k − 2 − p)t + t2 v(t)

=
k=1
 
(p−1)/2
Y
= t(p−1)/2  (4k − 2 − p) + t(p+1)/2 w(t),
k=1

where u(t), v(t), w(t) are polynomials. So the coefficient of t(p−1)/2 in f (1 + t) is

p−1   (p−1)/2
X j Y
c(p−1)/2 = χ(j) −ε (4k − 2 − p).
(p − 1)/2
j=(p−1)/2 k=1

Furthermore,

f (1 + t) = ((1 + t)p − 1) h(1 + t)


" p   #
X p
= ti h(1 + t)
i
i=1
" p   #
X p ti
= i! h(1 + t)
i i!
i=1
" p #
X ti
= ai h(1 + t),
i!
i=0

p p!

where a0 = 0, ai = i! = (p−i)!
i , so p | ai , i = 0, . . . , p − 1 : the conditions of Ex.21 are
Pp−1 i
verified, so f (1 + t) = i=0 ci t is such that c(p−1)/2 = p(A/B), p - B. Equating these
two evaluations of c(p−1)/2 , we obtain

p−1   (p−1)/2
X j Y A
χ(j) −ε (4k − 2 − p) = p , p - B.
(p − 1)/2 B
j=(p−1)/2 k=1

Multiplying by B(p − 1)!/2, we obtain, as p - B,

p−1   (p−1)/2
(p − 1)! X j (p − 1)! Y
χ(j) ≡ε (4k − 2)
2 (p − 1)/2 2
j=(p−1)/2 k=1
(p−1)/2
Y
≡ ε(2 · 4 · 6 · · · (p − 1)) (2k − 1) ≡ ε(p − 1)!
k=1
≡ −ε (mod p)

48
To prove that ε = +1, it remains to prove
  p−1  
(p − 1) X j
! χ(j) ≡ −1 (mod p)
2 (p − 1)/2
j=(p−1)/2

j 
The factor of ((p − 1)/2)! cancels the denominator of (p−1)/2 , which leaves

p−1
X p−1
χ(j) · j(j − 1) · · · (j − + 1)
2
j=(p−1)/2
p−1
X p−1
= χ(j) · j(j − 1) · · · (j − + 1).
2
j=1

The equality is justified because all terms for j < p−1 2 are zero. Collecting powers of j,
this is
p−1 (p−1)/2 (p−1)/2 p−1
X X X X p−1
χ(j) ak j k ≡ ak j k+ 2 (mod p)
j=1 k=0 k=0 j=1

for some integers ak . It’s important to note that a(p−1)/2 = 1.


Now, we compute p−1 n
P
j=1 j mod p. Let S denote this sum and let g be a generator
of Z/p× . Then
p−1
X p−1
X
n n
g S= (gj) ≡ j n = S (mod p).
j=1 j=1

Congruence holds because gj also runs over a complete system of nonzero residues mod
p. If g n 6≡ 1, then S ≡ 0. If g n ≡ 1, then j n ≡ 1 for all j, hence S ≡ p − 1 ≡ −1.
Returning to the previous sum, the only nonzero term modulo p is k = p−1 2 , so

(p−1)/2 p−1
X X p−1
≡ ak j k+ 2 ≡ a(p−1)/2 · (−1) = −1 (mod p)
k=0 j=1

as desired.

Ex. 6.23 If f (x) = xn + a1 xn−1 + . . . + an , ai ∈ Z, and p is prime such that p | ai for


i = 1, . . . , n, and p2 - an , show that f (x) is irreducible over Q (Eisenstein’s irreducibility
criterion).

Lemma. If f ∈ Z[x], deg(f ) ≥ 1, is not irreducible in Q[x], then there exist g, h ∈


Z[x], deg(g) ≥ 1, deg(h) ≥ 1 such that f = gh.
n
ak xk , ak
P
Proof. (lemma) Suppose that f (x) = ∈ Z, is not irreducible in Q[x].
k=0
Then f (x) = f1 (x)f2 (x), with f1 , f2 ∈ Q[X],
and deg(f1 ) ≥ 1, deg(f2 ) ≥ 1. As in
Ex. 6.5, we can write f1 (x) = λp(x), f2 (x) = µq(x) where λ, µ ∈ Q, and p, q ∈ Z[X] are
primitive. Let ν = λµ ∈ Q : write ν = u/v, u ∧ v = 1, v ≥ 1. Then r(x) = p(x)q(x) =
n
ck xk is primitive (Ex. 6.4), and f (x) = uv r(x) = uv p(x)q(x).
P
k=0
As vf (x) = ur(x), v | uci , i = 0, 1, . . . , n, with u ∧ v = 1, so u | ci for all i. The
polynomial r being primitive, v | 1, so v = ε = ±1.

49
Let g(x) = εup(x), h(x) = q(x).Then g, h ∈ Z[x], deg(g) ≥ 1, deg(h) ≥ 1, and f = gh
is the product of two non constant polynomials in Z[x].

Proof. (Ex. 6.23)


Let 
Z[x] → Fp [x]
ϕ:
p(x) = a0 + · · · + an x 7→ p(x) = a0 + · · · + an xn ,
n

where ai is the class of ai in Fp . ϕ is a ring homomorphism.


We show that f (x) = g(x)h(x), g, h ∈ Z[x], deg(g) ≥ 1, deg(h) ≥ 1 is impossible.
Indeed in such a situation,
f (x) = xn = g(x)h(x).
As the only irreducible factor of xn is x, the unicity of the decomposition of a polynomial
in irreducible factors in Fp [x] gives

g(x) = λxi , h(x) = µxj , λ, µ ∈ Fp , i, j ∈ N.

As deg(g) ≤ deg(g), deg(h) ≤ deg(h) and deg(g) + deg(h) = n = deg(f ) + deg(g),


this implies that i = deg(f ) = deg(f ), j = deg(g) = deg(g), so i ≥ 1, j ≥ 1. Therefore
p | g(0), p | h(0), so p2 | an = g(0)h(0), which is in contradiction with the hypothesis.
From the lemma we deduce that f (x) is irreducible in Q[x].

50

You might also like