Quadratic Residues

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Finite elds and quadratic residues

by Bruno Joyal
A. Fields of prime order
Theorem 1 If p is any prime number, then F = Z/pZ is a eld (of order p).
Proof: pZ is a maximal ideal of Z. By the ring isomorphism theorems the
result follows.
From now on F

will denote the nonzero elements of F.


Theorem 2 F

is a cyclic group of order p 1.


Proof: Obviously F

is a group of order p 1; we must show that it is


cyclic.
Let n be the exponent of F

, i.e. the least positive integer n such that


x
n
= 1 for all x F

. We know that n p 1 (because for any group G and


any x G, x
|G|
= 1). Then the polynomial x
n
1 has p 1 distinct roots in F,
which implies n p 1 (here we use the fact that F is a eld; in an arbitrary
ring R, a polynomial of degree n could have more than n roots - for example
take x
2
1 over Z/8Z, which has 4 roots). Hence n = p 1, F

is cyclic and
has (p 1) generators (also called primitive roots mod p).
Primitive roots are mysterious objects about which many questions remain
unanswered. There are even upper bound statements about the least positive
primitive root mod p which are equivalent to the Riemann hypothesis.
Theorem 3 Suppose we have an irreducible polynomial p(x) F[X] with deg p =
n. Then there exists a eld of order p
n
.
Proof: The polynomial ring F[X] is an integral domain (and even an Eu-
clidean domain, which is even stronger). Since p(x) is irreducible, (p(x)) is a
maximal ideal and F[X]/(p(x)) is a eld, by the isomorphism theorems for rings.
Writing for a root of p(x), we have the isomorphism F[X]/(p(x))

= F[],
which is a vector space of dimension n over F having 1, , ...,
n1
as a basis;
counting elements we thus nd p
n
of them.
1
Setting q = p
n
, we will from now on write F
q
for a eld with q elements (we
should in fact say the eld, but we havent yet showed that it is unique up to
isomorphism). We will say that F
q
is of characteristic p (the characteristic of a
eld is the additive order of 1).
The same argument as in the proof of Theorem 2 yields that F

q
is cyclic of
order q 1; therefore a primitive root satises x
q1
1 = 0, and q 1 is the
least positive integer having this property.
The Frobenius Automorphism
The map : F
q
F
q
given by x x
p
is called the Frobenius automorphism.
Theorem 4 The map satises the following properties:
a) It is a bijection;
b) It is a eld automorphism : that is, we have
(xy) = (x)(y)
(x + y) = (x) + (y)
c)
n
= 1
F
q
F
q
and
j
,= 1
F
q
F
q
for 0 < j < n.
Proof: To see a), note that q 1 and p are relatively prime; if G is a cyclic
group of order m, and (m, s) = 1 then x x
s
is an automorphism of G. So the
restriction of to F

q
is a bijection; moreover (0) = 0 and we are done.
The rst part of b) is immediate. To see the second part, expand (x + y)
p
and notice that the binomial coecients
_
p
j
_
=
p!
j!(pj)!
are all divisible by p for
1 j p 1.
The third part is left as an exercise.
Consider now the set S of the roots of p(x) = x
q
x (as a polynomial over
F
p
). Suppose , S. Then ()
q
=
q

q
= = 0, so that
S. Moreover by Theorem 4, (+)
q
(+) =
n
(+) (+) = 0,
so that also + S. Finally (
1
)
q

1
= (
q

1
)
q
= 1
q1
= 0.
Therefore S is closed under addition, multiplication and taking inverses! Thus
S is a eld. Examining the polynomial x
q
x, we nd that it is separable,
i.e. that all of its roots are distinct. Therefore [S[ = q = p
n
. Moreover every
element of a eld of order q is a root of p(x); therefore any eld of q elements is
a splitting eld for the polynomial p(x). Since splitting elds are unique up to
isomorphism, F
q
is unique up to isomorphism.
2
B. Squares in nite elds
The study of quadratic residues goes back a long way in the history of math-
ematics. Euler was the rst to have important ideas about this subject and he
furnished, with proof, several important theorems; moreover he conjectured
what is known as the Quadratic Reciprocity Law, which is without a doubt one
of the most beautiful and far-reaching theorems in elementary mathemat-
ics. Many great mathematicians such as Legendre and Lagrange tried without
success to prove the reciprocity law in the following decades. The study of
quadratic residues reached a high point when Gauss, at 21 years old, proved
the reciprocity law in his Disquisitiones Arithmeticae. He gave birth, in the
same book, to the theory of quadratic elds, seemingly without consciously
realizing that it is closely related to the reciprocity law. (To be more precise,
Gauss created the theory of binary quadratic forms, but his theory is completely
equivalent to the modern theory of quadratic elds).
Today, hundreds of proofs of the quadratic reciprocity law are known. Gauss
himself gave no less than six dierent proofs of the theorem in his lifetime.
Theorem 5 If p = 2, then every element of F

q
is a square. If p is an odd
prime then there are as many squares in F

q
as there are non-squares. (In
group-theoretic terms, if we denote F
2
q
the subgroup of nonzero squares in F
q
,
then F

q
/F
2
q

= 1)
Proof: If p = 2 then x x
2
is just the Frobenius automorphism, which
is a bijection; hence every element is a square. If p is odd, F

q
is cyclic of even
order and the result follows.
From now on we will assume p > 2.
Theorem 6
F
2
p
=
_
1
2
, 2
2
, ...,
_
p 1
2
_
2
_
.
Proof: The squares are repeated backwards after
_
p1
2
_
2
; indeed (pa)
2
=
p
2
2pa + a
2
= a
2
. Since we know there are
p1
2
squares, the set above must
contain all of them, exactly once.
3
Denition: The Legendre Symbol is dened as
_
x
p
_
=
_
1 if x = y
2
for some y F

p
1 otherwise
With this notation, for example, the statement If p is an odd prime then
there are as many squares in F

p
as there are non-squares can be stated as

xF

p
_
x
p
_
= 0.
Now we begin the fun.
Theorem 7 (Eulers Criterion) For x F

p
,
_
x
p
_
= x
p1
2
Proof: Clearly we must have x
p1
2
= 1, because (x
p1
2
)
2
= 1 and the only
elements of F

p
which square to 1 are 1. If x = y
2
then x
p1
2
= y
p1
= 1. Now
suppose 1 = x
p1
2
. Let be a primitive root and set x =
k
. Then
k(p1)
2
= 1
implies k must be even because is of order p 1. Therefore x = (
k
2
)
2
is a
square.
As a corollary, we have the First supplement to the Quadratic Reciprocity
Law:
_
1
p
_
= (1)
p1
2
=
_
1 if p 1 mod 4
1 if p 1 mod 4
For example, 29 = 4 7 + 1 and we nd 12
2
= 144 1 mod 29.
Another important corollary (which we could have deduced directly from
F

q
/F
2
q

= 1) is:
Theorem 8 For x, y F

p
,
_
xy
p
_
=
_
x
p
__
y
p
_
Proof:
(xy)
p1
2
= x
p1
2
y
p1
2
.
4
That was short! In words, the product of two residues is a residue, the product
of two non-residues is a residue, and the product of a residue and a non-residue
is a non-residue.
Now suppose we have a set S such that F

p
is the disjoint union of S and
S. Notice that every x F

p
can be written uniquely as (x)x, with x S and
(x) = 1. (In other words S is a set of coset representatives for the quotient
group F

p
/1.)
Theorem 9 (Gauss Lemma) . For y F

p
,
_
y
p
_
=

sS
(sy).
Proof: The map S S : s sy is a bijection, because if we had s
1
y = s
2
y
then we would have s
1
y = s
2
y s
1
= s
2
, contradicting the denition of S.
Therefore

sS
s =

sS
sy =

sS
(sy)sy = y
|S|
_

sS
(sy)
__

sS
s
_
Dividing out by

sS
s, and noting that [S[ =
p1
2
, we have
y
p1
2
=

sS
(sy)
which by Eulers criterion is equal to
_
y
p
_
.
We have the following corollary :
Theorem 10 (Gauss second supplement to the quadratic reciprocity law) .
_
2
p
_
= (1)
p
2
1
8
=
_
1 if p 1 mod 8
1 if p 3 mod 8
Proof: By the previous theorem, taking S = 1, 2, ...,
p1
2
, we know that
_
2
p
_
= (1)
v
where v is the number of integers
1 x
p 1
2
such that
p 1
2
< 2x < p.
We leave to the reader the task to count how many such integers there are
depending on the residue class of p (mod 8), and to conclude that the theorem
holds.
5
The next theorem was used by a young mathematician named Yegor Ivanovich
Zolotarev, in his proof of the quadratic reciprocity law. His proof is based on
permutations and is truly remarkable. Unfortunately Zolotarev died in a train
accident when he was only 31 years old. I am proud to say that this proof of
his lemma is mine (meaning : I havent seen it elsewhere!), and hopefully it is
nice enough not to dishonor Zolotarev.
Recall that the signature (or sign) of a permutation of a nite set is given
by (1)
t
where t is the number of pairs of elements which are inverted by the
permutation. Inverted obviously depends on which order we are using, but
luckily the total number of inversions is completely independent of the chosen
order.
Theorem 11 (Zolotarevs lemma) . For each a F

p
, consider the permutation
of F

p
given by t
a
: x ax. Then
sgn (t
a
) =
_
a
p
_
.
Proof: Consider
=

1j<ip1
(i j)
. Then
= sgn (t
a
)

1j<ip1
(t
a
(i) t
a
(j)) = sgn (t
a
)

1j<ip1
(ai aj)
Since there are p(p 1)/2 terms in the product, we have
= sgn (t
a
) a
p(p1)
2

1j<ip1
(i j) = sgn (t
a
) a
p(p1)
2

Considering that a
p
= a, and dividing out by , we have
sgn (t
a
) = a
p1
2
=
_
a
p
_
.
The following theorem was used by Gauss in a few of his proofs of the
Quadratic Reciprocity Law.
Theorem 12 (Gauss) If = e
2i
p
, then
_
a
p
_
=
1 +
a
+
2
2
a
+ ... +
(p1)
2
a
1 + +
2
2
+ ... +
(p1)
2
6
Proof: If a is a square mod p, then the top summands are just the bottom
summands rearranged. Otherwise the top summands (except 1) are all
m
where
m runs through the
p1
2
non-squares, each of them taken twice. Letting N, D
be the numerator and denominator, we have 0 = 2 +2 +... +2
p1
= N +D;
hence N = D and we are done.
Theorem 13 (The present author) If = e
2i
p
, then
_
a
p
_
=

1j<kp1

aj

ak

k
Proof: By Zolotarevs lemma,

1j<kp1

aj

ak
=
_
a
p
_

1j<kp1

k
because we can act on the exponents of exactly like we would on the residue
classes mod p.
(Those with a bit of knowledge of algebraic number theory may be able to
relate the above expression to the discriminant of the cyclotomic eld Q[
p
]
where
p
is a primitive p-th root of unity.)
Unfortunately I was not able to use the above theorem to prove the Quadratic
Reciprocity Law. Perhaps the reader can.
Before we proceed with the more serious theorem, we take a bit of time to
look at some possible applications of what weve done so far.
Theorem 14
__
p 1
2
_
!
_
2

_
1
p
_
mod p
Proof: The roots of the polynomial x
p1
2
1 are exactly the squares of F

p
.
By Theorem 6, and using the fact that F
p
[X] is a unique factorization domain,
we must have
x
p1
2
1 =
_
x 1
2
_ _
x 2
2
_
. . .
_
x
_
p 1
2
_
2
_
Equating the constant terms, we have
1 = (1)
p1
2
__
p 1
2
_
!
_
2
and the result follows from Eulers criterion.
7
In very much the same way, we can easily prove Wilsons theorem : (p1)!
1 mod p, by using the polynomial x
p1
1 instead.
Theorem 15 (Gauss) Let R be the set of primitive roots of F
p
. Then

rR
r =
1, unless p = 3.
Proof: We know that F

= Z/(p 1)Z, where one of the possible isomor-


phisms is given by : k where k (Z/(p 1)Z)

. If is a primitive root
then the other primitive roots are
a
where a ranges through (Z/(p 1)Z)

.
Thus

rR
r
_
=

rR
(r) =

(a,p1)=1
a.
Now take b (Z/(p 1)Z)

, b ,= 1 (which is possible when p > 3); we have


b

(a,p1)=1
a =

(a,p1)=1
ba =

(a,p1)=1
a
which implies

rR
r
_
=

(a,p1)=1
a = 0
which nally implies

rR
r = 1.
As a nice application of Eulers criterion, we prove a well-known theorem of
Fermat. Fermat annouced this result without proof, as he usually did; the rst
to publish a proof was Euler himself, many years after Fermats claim, but his
proof was extremely long and involved. The following proof is certainly the best
possible (the book proof, in Erdos words)!
Theorem 16 (Euler-Fermat) Let p be a prime of the form 4n + 1; then p =
a
2
+b
2
is solvable in integers a, b and the solution is unique up to rearrangement.
Proof: We know by Eulers criterion that
_
1
p
_
= 1; hence we can nd
n N such that n
2
= mp 1. Then mp = n
2
+ 1 = (n i)(n + i). Since p
divides mp, but divides neither ni, n+i, and since Z[i] is a unique factorization
domain (and even a norm-Euclidean domain), p must not be prime in Z[i]. Thus
p admits a factorization p = (a + bi)(c + di) with a, b, c, d Z

. Taking the
norm on both sides we have p
2
= (a
2
+ b
2
)(c
2
+ d
2
); this is in Z, so necessarily
p = a
2
+ b
2
= c
2
+ d
2
.
8
For example, 2141 is prime and indeed we have 2141 = 46
2
+ 5
2
.
As an exercise, and taking for granted that Z[

2] is a unique factorization
domain, the reader may wish to prove that primes of the form 8n+1 and 8n+3
can be written as a
2
+ 2b
2
. (Hint : show that in both cases
_
2
p
_
= 1).
Finally, two easy theorem to relax our brain before the big work:
Theorem 17 In any nite eld F
p
with p > 5, every nonzero element is the
sum of two nonzero squares.
Proof: Two cases: rst, suppose x is not a square. Consider the
p+1
2
elements
_
x, x 1
2
, x 2
2
, ..., x
_
p 1
2
_
2
_
.
None of them is zero since x is not a square; moreover all of them are distinct.
Since there are only
p1
2
non-squares in F

p
, by the pigeonhole principle one of
the above elements is a square, and we know that it is not x; hence for some
a ,= 0 we have x a
2
= b
2
and we are done.
The second case, with x a square, is a bit more tricky. First, notice that if
any one of the squares of F

p
can be written as a sum of two nonzero squares,
then all of them can : indeed suppose we have y
2
= a
2
+ b
2
. Then for any
choice of c we have also (cy)
2
= (ca)
2
+ (cb)
2
, and (cy)
2
will run through all
the elements of F
2
p
. Thus we need only nd a single solution (x, y, z) to the
equation x
2
= y
2
+z
2
. To do this, nd m, n F

p
such that
_
m
n
_
2
,= 1 (this is
why we need p > 5 : we need enough elements to choose from!). Then by the
well-known formula for pythagorean triples,
(m
2
+ n
2
)
2
= (m
2
n
2
)
2
+ (2mn)
2
and we know that neither m
2
n
2
nor m
2
+ n
2
are zero or else we would
have
_
m
n
_
2
,= 1; hence we have our desired solution.
Theorem 18 In any nite eld F
p
with p > 3, every nonzero element is the
sum of n nonzero squares, for any positive integer n > 1.
Proof: By induction on n. We case n = 2 was settled above; now suppose
we have proven it up to n 1. Write x = a
2
+b
2
. By the induction hypothesis,
we can write b
2
= x
2
1
+ ... + x
2
n1
and we have x = a
2
+ x
2
1
+ ... + x
2
n1
.
9
And now, the holy grail, for which we give three proofs.
Theorem 19 (Conjectured by Euler, proved by : Gauss (6) - Cauchy (2) -
Jacobi - Dirichlet (2) - Lebesgue (2) - Eisenstein (5) - Kummer - Liouville -
Dedekind (3) - Kronecker (7) - etc...)
The Quadratic Reciprocity Law : if p, q are distinct odd primes, then
_
p
q
__
q
p
_
= (1)
p1
2
q1
2
First proof (Gauss): This is a modernized version of one of Gauss proofs,
which remains today one of the most beautiful and penetrating. All calculations
take place in = F
q
, an algebraic closure of F
q
. We let denote a root of
x
p
1 dierent from 1.
We dene the Gauss sums as the function : F

p
given by
a

xF

p
_
x
p
_

ax
.
First, note that we have
_
a
p
_
(a) =

xF

p
_
ax
p
_

ax
=

yF

p
_
y
p
_

y
= (1)
because y = ax runs through F
p
as x does.
Moreover,
(1)
2
=
_
_

xF

p
_
x
p
_

x
_
_
_
_

yF

p
_
y
p
_

y
_
_
=

(x,y)F

p
F

p
_
xy
p
_

x+y
Now we can write x = yt, and t will run through F

p
. This yields
(1)
2
=

(t,y)F

p
F

p
_
t
p
_

(1+t)y
=

tF

p
_
t
p
_
_
_

yF

(1+t)y
_
_
Now if 1 +t ,= 0, then (1 +t)y runs through F

p
as y does; hence in that case

yF

(1+t)y
= +
2
+ ... +
p1
= 1.
When 1 + t = 0,

yF

(1+t)y
=

yF

p
1 = p 1.
10
Therefore
(1)
2
=

tF

p
_
t
p
_
_
_

yF

(1+t)y
_
_
=
_
1
p
_
(p 1)

1=tF

p
_
t
p
_
= p
_
1
p
_
.
And now the nal blow. Note that when we distribute the exponent q inside
the parentheses, we use the fact that has characteristic q.
(1)
q
=
_
_

xF

p
_
x
p
_

x
_
_
q
=

xF

p
_
x
p
_

qx
= (q) =
_
q
p
_
(1).
Therefore
(1)
q1
=
_
q
p
_
= ((1)
2
)
q1
2
= p
q1
2
_
1
p
_
q1
2
=
_
p
q
_
(1)
p1
2
q1
2
which completes the proof of the Quadratic Reciprocity Law.
Before proceeding with Eisensteins proof, we need two useful lemmas :
Lemma 1: If m is an odd positive integer, then
(m1)/2

j=1
sin
2
2j
m
=
m
2
m1
To see this, set z = e
2i
m
. Then
(m1)/2

j=1
sin
2
2j
m
= (1)
m1
2

1jm1
z
j
z
j
2i
=
(1)
m1
2
(2i)
m1

1jm1
(z
j
z
j
) =
1
2
m1

1jm1
(z
j
z
j
)
=
z
1+2+...+(m1)
2
m1

1jm1
(1 z
2j
) =
z
m(m1)/2
2
m1

1jm1
(1 z
j
),
because ez
2j
: 1 j m1 = z
j
: 1 j m1 (z being of odd
order).
11
Notice that z
m(m1)/2
= 1, and nally that z
j
: 1 j m 1 are the
roots of
x
m
1
x 1
= x
m1
+ ... + x + 1.
So the last product is just this polynomial evaluated at x = 1; which yields
(m1)/2

j=1
sin
2
2j
m
=
m
2
m1
as desired.
Lemma 2: If m is an odd positive integer, then
sinmx
sinx
= (4)
m1
2
(m1)/2

j=1
_
sin
2
x sin
2
2j
m
_
.
To see that this is true, rst notice that when m is odd,
sin mx
sin x
is indeed
a polynomial of degree m in sinx, as can be seen by inspecting sinmx =
(
_
1 sin
2
x + i sinx)
m
. Its zeroes are the zeroes of sinmx, i.e. mx = j for
some j
_
1, 2, ...
m1
2
_
; moreover each of them occurs exactly once. Finally
the constant term (4)
m1
2
= (1)
m1
2
2
m1
can be obtained by comparing the
coecient of x in the Taylor series expansions of sinmx and of
sinx(4)
m1
2
(m1)/2

j=1
_
sin
2
x sin
2
2j
m
_
.
In the rst case it is equal to m, and in the second case it is equal to
(4)
m1
2
(m1)/2

j=1
sin
2
2j
m
= 2
m1
(m1)/2

j=1
sin
2
2j
m
= m
by Lemma 1.
Second proof (Eisenstein): This proof is beautiful. Eisenstein was able
to use similar ideas to prove the cubic and quartic reciprocity laws as well.
As in Gauss lemma, we let S and T be sets of representatives for F

p
/F
2
p
and
F

q
/F
2
q
, respectively; so we can write x F

p
uniquely as
s
(x)x with x S, and
y F

q
uniquely as
t
(y)y with y T.
12
Then
sin
_
2
p
qx
_
= sin
_
2
p

s
(qx)qx
_
=
s
(qx) sin
_
2
p
qx
_
Then by Gauss lemma,
_
q
p
_
=

xS

p
(qx) =

xS
sin
_
2
p
qx
_
sin
_
2
p
qx
_ =

xS
sin
_
2
p
qx
_
sin
_
2
p
x
_ .
where the last step is justied because x qx is a bijection of S onto itself
(so we have merely rearranged the bottom terms).
Now we apply Lemma 2, with m = q, to each term of the product, to get :
_
q
p
_
=

xS
(4)
q1
2

tT
_
sin
2
2s
p
sin
2
2t
q
_
= (4)
p1
2
q1
2

xS

tT
_
sin
2
2s
p
sin
2
2t
q
_
.
because [S[ =
p1
2
.
Reversing the roles of p and q we get
_
p
q
_
= (4)
p1
2
q1
2

tT

xS
_
sin
2
2t
q
sin
2
2s
p
_
.
so nally, noticing that each term of the product has simply changed sign,
we have
_
q
p
_
_
p
q
_ =

xS

tT
(1) = (1)
p1
2
q1
2
which is the Quadratic Reciprocity Law.
13
Third proof : This proof is by G. Rousseau and is a modernized version
of Gauss fth proof. It is remarkable for depending only on general group-
theoretic considerations. The reader may nd this third proof more illuminating
than the above two, which seem to yield this dicult and beautiful theorem in
an almost magical fashion.
We consider the product of the elements of the group
G
pq
= (Z/pqZ)

/1
in two dierent ways; once directly, and once using the Chinese remainder the-
orem. Writing U = (1, 1), (1, 1) and G
(p,q)
=
_
F

p
F

q
_
/U, we clearly
have, by the Chinese remainder theorem,
G
pq

= G
(p,q)
.
We will denote : G
pq
G
(p,q)
the obvious isomorphism.
First, note that a set of coset representatives for G
pq
in (Z/pqZ)

may be
taken to be
S =
_
1 n
pq 1
2
: (n, pq) = 1
_
,
in very much the same way that we had chosen 1, 2, . . . , (p 1)/2 as a set of
coset representatives for F

p
/1. Now, we let
A =
p1
_
j=1
j
p1
_
j=1
(p + j) ...
p1
_
j=1
__
q 1
2
1
_
p + j
_
p1
2
_
j=1
_
q 1
2
p + j
_
so that A contains exactly the integers coprime to p in the range 1 n
pq1
2
. The last union is from j = 1 to j = (p 1)/2, because we dont want
to go any farther than
pq1
2
. Now notice that we have accidentally included in
A all the multiples of q lying between 1 and
pq1
2
. Indeed we have only been
careful not to include in A the multiples of p, but the multiples of p and the
multiples of q are disjoint between 1 and pq 1. Hence
S = A
_
q, 2q, . . . ,
p 1
2
q
_
.
Be sure you are convinced of this. It helps to notice that the largest multiple
of q less than
pq1
2
is
_
pq 1
2q
_
q =
p 1
2
q.
So now we can write
14

sS
s =

p1
j=1
j

p1
j=1
(p + j)...

p1
j=1
__
q1
2
1
_
p + j
_
p1
2
j=1
_
q1
2
p + j
_
q 2q ...
p1
2
q
Reducing mod p, we nd that

sS
s
(p 1)!
(q1)/2
q
(p1)/2
(1)
(q1)/2
_
q
p
_
mod p
by Wilsons theorem and Eulers criterion. Using the same reasoning but
inverting the roles of p and q, we get
() =
_

sS
s1
_
=
_
(1)
(q1)/2
_
q
p
_
, (1)
(p1)/2
_
p
q
__
U.
Now we forget about this for a minute and take the product directly in
G
(p,q)
. A set of coset representatives for G
(p,q)
in F

p
F

q
is given by
S

=
_
(i, j) : 1 i j, 1 j
q 1
2
_
.
(To see this, note that all of these elements are incongruent mod U and that
there are exactly the right number of them.) Taking the product we have
() =

sS

sU =
_
(p 1)!
(q1)/2
,
_
q 1
2
!
_
p1
_
U
=
_
_
(1)
(q1)/2
,
_
_
q 1
2
!
_
2
_
p1
2
_
_
U
=
_
(1)
(q1)/2
, (1)
p1
2
(1)
p1
2
q1
2
_
U
where the last step is justied by Theorem 14. So nally, comparing our two
expressions for (), we have
(1, 1)U =
__
q
p
_
,
_
p
q
_
(1)
p1
2
q1
2
_
U
from which we simply read o the Quadratic Reciprocity Law.
15
REFERENCES AND SUGGESTED READING
1. J.-P. SERRE : Cours darithmetique, Presses Universitaires de France
(1970)
2. G. ROUSSEAU : On the Quadratic Reciprocity Law, J. Austral. Math.
Soc. (Series A) 51 (1991) 423-425
16

You might also like