You are on page 1of 81

Arithmetic of Quadratic Forms

1 Foundation
Throughout this section, F always denotes a field of characteristic different from 2.

1.1 Quadratic Forms and Quadratic Spaces


An (n-ary) quadratic form over F is a polynomial f in n variables x1 , . . . , xn over F that is
homogeneous of degree 2. In general, f takes the form
n
X
f (x1 , . . . , xn ) = bij xi xj , bij ∈ F.
i,j=1

To render the coefficients symmetric, it is customary to rewrite f as


n
X
f (x1 , . . . , xn ) = aij xi xj ,
i,j=1

where aij = (bij + bji )/2. In this way, f determines a symmetric matrix (aij ), which we
shall denote by Af . In terms of matrix multiplication, we have

f (x) = xt Af x

where x = (x1 , . . . , xn )t .
Let a be an element in F . We say that a is represented by f if the equation

(∗) f (x) = a

has a solution in F n . The representation problem of quadratic forms is to determine, in an


effective manner, the set of elements of F that are represented by a particular quadratic
form over F . We shall discuss the case when F is a field of arithmetic interest, for instance,
the field of complex numbers C, the field of real numbers R, a finite field F, and the field of
rational numbers Q. The representation problem for quadratic forms over any one of these
fields has a very satisfactory solution. At the end, we shall discuss the solubility of equation
(∗) over a subring R of F , that is, the problem of finding solution of (∗) in Rn . The most
interesting and difficult case is when R is the ring of integers Z for which there is still a lot
of questions left unanswered.
Let f and g be two n-ary quadratic forms. We say that f and g are equivalent, written
f∼ = g, if there exists an invertible matrix C ∈ GLn (F ) such that f (x) = g(Cx). This is the
same as saying that there is an invertible homogeneous linear substitution of the variables
x1 , . . . , xn which takes the form g to the form f . Since

g(Cx) = (Cx)t Ag (Cx) = xt (C t Ag C)x,

1
the condition f (x) = g(Cx) is equivalent to the matrix equality

Af = C t Ag C.

Thus equivalence of forms amounts to congruence of the associated symmetric matrices. As


expected, equivalence of forms is indeed an equivalence relation. It is clear that equivalent
forms represent the same set of elements of F .

Example 1.1 Let g(x, y) be the binary form xy. If we make the substitution x 7→ x+y, y 7→
x − y, then g changes to

g(x + y, x − y) = (x + y)(x − y) = x2 − y 2 = f (x, y).

The matrix C in this case is 11 −1


1 . This can be verified by


     
1 0 1 1 0 1/2 1 1
Af = = = C t Ag C.
0 −1 1 −1 1/2 0 1 −1

Any quadratic form f gives rise to a map Qf : F n −→ F defined by Qf (x) = xt Af x. We


shall refer to Qf as the quadratic map defined by f . The notion of equivalence of quadratic
forms, f ∼ = g, amounts to the existence of a linear automorphism C of F n (that is, an
invertible matrix in GLn (F )) such that Qf (x) = Qg (Cx) for all x ∈ F n . Note that the
quadratic map Qf determines the quadratic form f uniquely. For, suppose that Qf = Qg
as maps from F n to F . Let e1 , . . . , en be the standard basis for F n . Then for any i, we
have
(Af )ii = Qf (ei ) = Qg (ei ) = (Ag )ii .
For i 6= j, we have
Qf (ei + ej ) = Qf (ei ) + Qf (ej ) + 2(Af )ij ,
and a similar equation for Qg (ei +ej ). Therefore, (Af )ij = (Ag )ij ; thus Af = Ag and f = g.
The quadratic map Qf satisfies the following properties:

(1) For any a ∈ F and x ∈ F n , Qf (ax) = a2 Qf (x).

(2) The function Bf (x, y) = 21 [Qf (x + y) − Qf (x) − Qf (y)] is a symmetric bilinear form
on F n . That is, Bf is a function from F n × F n to F satisfying

(i) Bf (x, y) = Bf (y, x) for all x, y ∈ F n , and


(ii) Bf (ax + by, z) = aBf (x, z) + bBf (y, z) for all a, b ∈ F and x, y, z ∈ F n .

Note that the quadratic map Qf can be recaptured by the symmetric bilinear form Bf ,
since
Qf (x) = Bf (x, x), ∀x ∈ F n .
This motivates the following geometric approach to the notion of a quadratic form. Let
V be an n-dimensional vector space over F equipped with a symmetric bilinear form B :

2
V ×V −→ F . The pair (V, B) is called a quadratic space, and associate with it is a quadratic
map Q = QB : V −→ F given by Q(v) = B(v, v) for all v ∈ V . As in (1) and (2) above, we
have Q(av) = a2 Q(v) for all a ∈ F and v ∈ V , and 2B(u, v) = Q(u + v) − Q(u) − Q(v) for
all u, v ∈ V . Therefore, Q and B determine each other and hence it is legitimate to write
(V, Q) to represent the quadratic space (V, B).
Now, suppose that v1 , . . . , vn is a basis for V . Then the quadratic space (V, B) gives
rise to a quadratic form X
f (x1 , . . . , xn ) = B(vi , vj )xi xj ,
i,j

with
Af = (B(vi , vj )).
If we identify V with F n via the basis v1 , . . . , vn that is, we identify each vector v =
x1 v1 + . . . xn vn with the column x = (x1 , . . . , xn )t , then Q = QB is precisely the quadratic
map Qf associated with the form f . An element a ∈ F is represented by the form f if and
only if a is represented by V , that is, there exists a vector v ∈ V such that Q(v) = a.
Now, let us choose another
P basis u1 , . . . , un for V , and let g be the resulting quadratic
form. Suppose that ui = k cki vk , then

(Ag )ij = B(ui , uj )


X X
= B( cki vk , c`j v` )
k `
X
= cki B(vk , v` )c`j
k,`
t
= (C Af C)ij

where C = (ck` ). Thus the quadratic space (V, B) determines uniquely an equivalence class
of quadratic forms, which we shall denote by [fB ].
If (V, B) and (V 0 , B 0 ) are two quadratic spaces, we say that they are isometric, written
(V, B) ∼= (V 0 , B 0 ), if there exists an isomorphism σ : V −→ V 0 such that
B 0 (σ(u), σ(v)) = B(u, v), ∀u, v ∈ V.

Such σ is called an isometry from V to V 0 . The set of all isometries from V to V itself form
a group O(V ) under the composition of functions. It is clear that

(V, B) ∼
= (V 0 , B 0 ) ⇐⇒ [fB ] = [fB 0 ].
Thus there is a one-to-one correspondence between the equivalence classes of n-ary quadratic
forms and the isometry classes of n-dimensional quadratic spaces. In this set of lecture notes,
we shall adopt the geometric language of quadratic spaces.
Let (V, B) be an n-dimensional quadratic space, and v1 , . . . , vn be a basis for V . Let A
be the symmetric matrix (B(vi , vj )). We call A a matrix for V and write

V ∼
= A.

3
So, V ∼=A∼ = C t AC for any C ∈ GLn (F ). We call V nondegenerate if det(A) is nonzero.
Otherwise, V is degenerate. For a nondegenerate space V , its discriminant of V , denoted
d(V ), is defined to be the square class det(A)F ×2 in F × /F ×2 . For convention, the zero
space is considered to be nondegenerate and its discriminant is defined to be F ×2 .
If W is a subspace of V , the map BW : W × W −→ F defined by BW (x, y) = B(x, y)
for all x, y ∈ W is a symmetric bilinear form on W . Thus (W, BW ) is also a quadratic
space. We say that W is a nondegenerate subspace of V if (W, BW ) is nondegenerate as a
quadratic space.
Let (V, B) be a quadratic space. Let V ∗ be the vector space of all homomorphisms from
V to F . It is called the dual space of V . The function B̂ : V −→ V ∗ defined by

B̂(x)(y) = B(x, y)

is clearly linear.
Suppose that E = {v1 , . . . , vn } is a basis for V . For each i, define vi∗ ∈ V ∗ by
(
∗ 1 if i = j,
vi (vj ) = δij =
0 otherwise.

Then E∗ = {v1∗ , . . . , vn∗ } is a basis for V ∗ .

Lemma 1.2 The matrix of B̂ with respect to the bases E and E∗ is the symmetric matrix
for V associated with E.
Proof. From B̂(ei )(ej ) = B(ei , ej ) follows
n
X
B̂(ei ) = B(ei , ej )e∗j .
j=1

Corollary 1.3 A quadratic space (V, B) is nondegenerate if and only if the map B̂ : V −→
V ∗ is an isomorphism.

1.2 Orthogonal Decomposition


Let (V, B) be a quadratic space. Two vectors x, y ∈ V are orthogonal if B(x, y) = 0. Two
subsets X and Y of V are said to be orthogonal if B(x, y) = 0 for all x ∈ X, y ∈ Y . With
each subset X of V , the orthogonal complement of X in V is the set

X ⊥ = {v ∈ V : B(v, x) = 0 for all x ∈ X}.

The orthogonal complement of V itself is called the radical of V , denoted by rad(V ) (but
not by V ⊥ ).

Lemma 1.4 Let X, Y be subsets of V . Then

4
(a) X ⊥ is a subspace of V .

(b) If X ⊆ Y , then X ⊥ ⊇ Y ⊥ .

(c) X ⊆ X ⊥⊥

(d) V is nondegenerate if and only if rad(V ) = 0.

Proof. Parts (a), (b) and (c) are direct consequences of the definition of orthogonal com-
plements. For (d), note that the kernel of the map B̂ is precisely rad(V ). Therefore,
dim(V ) = dim(rad(V )) + dim(Im(B̂)). By Corollary 1.3, V is nondegenerate if and only
if B̂ is an isomorphism. Thus V is nondegenerate if and only if dim(rad(V )) = 0, that is,
rad(V ) = 0.2

Definition 1.5 Let (V1 , B1 ) and (V2 , B2 ) be two quadratic spaces. The orthogonal sum
V1 ⊥ V2 is the quadratic space V1 ⊕ V2 with the symmetric bilinear form B defined by
B(x1 + x2 , y1 + y2 ) = B1 (x1 , y1 ) + B(x2 , y2 ) for all x1 , y1 ∈ V1 and x2 , y2 ∈ V2 .
So if V = V1 ⊥ V2 as above, then (Vi , Bi ) ∼ = (Vi , B|V ) for i = 1, 2, and V1 and V2 are
i
orthogonal in V .

Theorem 1.6 Let (V, B) be a quadratic space. Suppose that V = rad(V ) ⊕ W for some
subspace W . Then
(a) V = rad(V ) ⊥ W .

(b) W is nondegenerate.

(c) (W, B|W ) is determined up to isometry by V .

Proof. Part (a) is clear. For part (b), suppose that x ∈ W is orthogonal to all vectors in
W . Then x is orthogonal to all vectors in rad(V ) ⊥ W = V . Therefore, x ∈ rad(V ) ∩ W
and hence x = 0.
For part (c), suppose that V = rad(V ) ⊕ W1 . Then each x ∈ W can be written uniquely
as
x = y + α(x), y ∈ rad(V ), α(x) ∈ W1 .
One can check that the map α : W −→ W1 defined by x 7−→ α(x) is a vector space
homomorphism. It is clear that α is injective and hence surjective since dim(W ) = dim(W1 ).
If x = y + α(x) and x0 = y 0 + α(x0 ), then

B(x, x0 ) = B(y + α(x), y 0 + α(x0 )) = B(α(x), α(x0 )).

Hence α is an isometry.2
The isometry class of (W, B|W ) obtained in the above theorem is called the nondegen-
erate component of (V, B). The classification of general quadratic spaces reduces to the
classification of their nondegenerate components.

5
Lemma 1.7 Let V, W, V 0 , W 0 be quadratic spaces.

(a) V ⊥ V 0 ∼
=V0 ⊥V.

(b) If V ∼
= W and V 0 ∼
= W 0 , then V ⊥ V 0 ∼
= W ⊥ W 0.

(c) If A is a matrix for V and A0 is a matrix for V 0 , then


 
0 ∼ A 0
V ⊥V = .
0 A0

(d) V ⊥ V 0 is nondegenerate if and only if both V and V 0 are nondegenerate. In this case,
d(V ⊥ V 0 ) = d(V )d(V 0 ).

Proof. Everything is obvious.2

Proposition 1.8 If W is a nondegenerate subspace of a quadratic space (V, B) , then


V = W ⊥ W ⊥.
Proof. It suffices to show that V = W ⊕ W ⊥ . Since W is nondegenerate, 0 = ker(B
d W) =

W ∩ W . If x ∈ V and h = B̂(x)|W , then because W is nondegenerate there exists y ∈ W
with h = B W (y). Hence for all z ∈ W ,
d

B(x, z) = B̂(x)(z) = h(z) = B


d W (y)(z) = B(y, z)

and we can write x = y + (x − y) where y ∈ W and (x − y) ∈ W ⊥ by the above equation.


Thereby V = W + W ⊥ is proven.2
The above proposition implies that dim(W ) + dim(W ⊥ ) = dim(V ) whenever W is a
nondegenerate subspace of V . Below we show that the same additive property of the
dimension holds for every subspace provided V itself is nondegenerate.

Proposition 1.9 Let (V, B) be a nondegenerate quadratic space, and W be a subspace of


V . Then

(i) dim(W ) + dim(W ⊥ ) = dim(V ).

(ii) (W ⊥ )⊥ = W .

Proof. Part (ii) is a direct consequence of part (i) because W ⊆ (W ⊥ )⊥ . The map B̂ :
V −→ V ∗ is an isomorphism. Since W is a subspace of V , the canonical projection V ∗ −→
W ∗ is surjective. The kernel of the composition V −→ V ∗ −→ W ∗ is W ⊥ . Therefore,
dim(V ) = dim(W ⊥ ) + dim(W ∗ ) = dim(W ⊥ ) + dim(W ). 2

6
1.3 Witt’s Theorems
Let (V, B) be a quadratic space. A nonzero vector v ∈ V is called isotropic if Q(v) = 0.
Otherwise, v is called anisotropic.

Theorem 1.10 Every quadratic space has an orthogonal basis.


Proof. Let (V, B) be a quadratic space. Since rad(V ) is an orthogonal summand of V , we
may assume that V is nondegenerate. Suppose that Q(x) = 0 for all x ∈ V . Then for any
u, v ∈ V ,
1
B(u, v) = [Q(u + v) − Q(u) − Q(v)] = 0,
2
which implies that rad(V ) = V which is impossible.
Now, pick an anisotropic vector x ∈ V . Then F x is a nondegenerate subspace, and
hence we have a decomposition V = F x ⊥ V 0 for some subspace V 0 . The subspace V 0
is also nondegenerate. An induction argument on the dimension of V will complete the
proof.2

Corollary 1.11 Every invertible symmetric matrix in GLn (F ) is congruent to a diagonal


matrix.
We use the notation ha1 , . . . , an i to denote a diagonal matrix with a1 , . . . , an as the
diagonal entries. Then V ∼
= ha1 , . . . , an i if V has an orthogonal basis v1 , . . . , vn such that
Q(vi ) = ai for all i.
Let v be an anisotropic vector in a quadratic space (V, B). Define a map τv : V −→ V
by
2B(v, x)
τv (x) = x − v.
Q(v)
It is easy to verify that τv is linear. It is called the symmetry with respect to v (or a
reflection with respect to the hyperplane v ⊥ ).

Lemma 1.12 For each anisotropic vector v ∈ V , τv is an isometry of V and det(τv ) = −1.
Proof. The first assertion can be checked directly. Since Q(v) is nonzero, the subspace F v
is nondegenerate and hence V = F v ⊥ V 0 for some subspace V 0 . Let v2 , . . . , vn be a basis
for V 0 . Then v, v2 , . . . , vn is a basis for V . The matrix of τv with respect to this basis is the
diagonal matrix h−1, 1, . . . , i. Therefore, det(τv ) = −1.2

Lemma 1.13 Let (V, B) be a quadratic space. If x and y are anisotropic vectors of V with
Q(x) = Q(y), then there is an isometry σ of V such that σ(x) = y.
Proof. Let u = (x + y)/2 and v = (x − y)/2. Then B(u, v) = 0 and Q(x) = Q(u) + Q(v).
Either Q(u) or Q(v) is nonzero. In the first case, −τu (x) = y and in the second case
τv (x) = y.2

7
Lemma 1.14 Let (V, B) and (V 0 , B 0 ) be two quadratic spaces. If σ : V −→ V 0 is an
isometry and W is a subspace of V , then σ(W ⊥ ) = σ(W )⊥ .
Proof. Let x ∈ W ⊥ . For any w ∈ W ,

B 0 (σ(x), σ(w)) = B(x, w) = 0.

Therefore, σ(W ⊥ ) ⊆ σ(W )⊥ . The reverse inclusion can be proved similarly.2

Theorem 1.15 (Witt’s Cancellation Theorem) If V, V1 , V2 are nondegenerate quadratic


spaces such that
V1 ⊥ V ∼
= V2 ⊥ V,
then V1 ∼
= V2 .
Proof. Since V is the orthogonal sum of 1-dimensional subspaces, it suffices to consider the
case where dim(V ) = 1; thus V = F x. It follows from the hypothesis and Lemma 1.13 that
there exists an isometry
Σ : V1 ⊥ F x −→ V2 ⊥ F x
such that Σ(x) = x. By Lemma 1.14, Σ(V1 ) = V2 .2

Corollary 1.16 (Witt’s Extension Theorem) Let V and V 0 be isometric nondegenerate


quadratic spaces. Suppose that W and W 0 are nondegenerate subspaces of V and V 0 , respec-
tively, and that σ : W −→ W 0 is an isometry. Then there exists an isometry Σ : V −→ V 0
which extends σ, that is, Σ|W = σ.
Proof. Let τ : V 0 −→ V be an isometry. Then by Proposition 1.8,

W ⊥ W ⊥ = V = τ (W 0 ) ⊥ τ (W 0 )⊥ .

Since W ∼= τ (W 0 ), it follows from Theorem 1.15 that there exists an isometry σ 0 from W ⊥
to τ (W ) . Then Σ = σ ⊥ (τ −1 σ 0 ) is the required isometry.2
0 ⊥

1.4 Witt Index


Definition 1.17 A quadratic space V is said to be isotropic if it contains an isotropic
vector. Otherwise, it is anisotropic. The space V is said to be totally isotropic if every
nonzero vector in V is isotropic.

Theorem 1.18 Let (V, B) be a 2-dimensional nondegenerate quadratic space. The follow-
ing conditions are equivalent:

(a) V is isotropic.
 
∼ 0 1
(b) V = .
1 0

8
(c) V ∼
= h1, −1i ∼
= hα, −αi for any α ∈ F × .
 
× ∼ 0 α
(d) For any α ∈ F and any γ ∈ F , V = .
α γ

(e) −d(V ) = F ×2 .

Proof. Suppose that V is isotropic. Then V has an isotropic vector x, and V = F x ⊕ F y


for some y ∈ V . Let α = B(x, y) and γ = Q(y). Since V is nondegenerate, α ∈ F  × . If we

1 γ 0 1
let e = α x and f = − 2α x + y, then the symmetric matrix associated to {e, f } is .
1 0
This proves (a) =⇒ (b).
Suppose (b) holds and the basis which yields the matrix is {x, y}. For any α ∈ F × ,
Q( 2 x + αy) = α. So, the subspace spanned by u = 21 x + αy is nondegenerate and hence
1

V = F u ⊥ F v for some v ∈ V . Comparing the discriminants of both sides shows that v


can be chosen so that Q(v) = −α. This proves (b) =⇒ (c).
If (c) holds, then V has an isotropic vector, say x, which can be extended to a basis
{x, y} of V . Note that B(x, y) must be nonzero because V is nondegenerate. Now, let

α γ − Q(y)
e= x, f = x + y.
B(x, y) 2B(x, y)

Then {e, f } is a basis for M whose associated symmetric matrix is the one stated in (d).
The implication (d) =⇒ (e) is trivial. We now show that (e) =⇒ (a). If {x, y} is an
orthogonal basis for V and hα1 , α2 i is the associated symmetric matrix, then α1 α2 = −γ 2
for some γ ∈ F × . One easily checks that the vector x + γβ2−1 y is isotropic.2
The preceding theorem implies that there is only one isometry class of nondegenerate
isotropic 2-dimensional quadratic space. Any one of such space is called a hyperbolic plane
and is denoted by H.

Definition 1.19 A quadratic space V is said to be universal if V represents all elements


in F .

Corollary 1.20 Let (V, B) be a nondegenerate isotropic quadratic space. Then V ∼=H⊥
W for some subspace W . In particular, every nondegenerate isotropic quadratic space is
universal.
Proof. The second assertion is a consequence of Theorem 1.18. Let x be an isotropic vector
in V . Since V is nondegenerate, there exists y ∈ V such that B(x, y) = 1. The subspace
F x ⊕ F y is isometric to H.2
By the preceding corollary, every nondegenerate quadratic space V has an orthogonal
decomposition of the form

V ∼
= H ⊥ · · · ⊥ H ⊥ V0 = Hm ⊥ V0

9
where V0 is nondegenerate and anisotropic. The number m is well-defined, that is, it is
independent of the way we obtain the above orthogonal decomposition. For, suppose that
V ∼
= Hk ⊥ V1
where V1 is nondegenerate and anisotropic. If k > m, by Witt’s Cancellation Theorem it
follows that
Hk−m ⊥ V1 ∼ = V0 ,
which contradicts that V0 is anisotropic. Therefore, k = m, and V0 ∼
= V1 by Witt’s Cancel-
lation Theorem once again. In summary, we have
Corollary 1.21 (Witt’s Decomposition) Every nondegenerate quadratic space V has an
orthogonal decomposition
V = Hm ⊥ V0 ,
where V0 is nondegenerate and anisotropic. The integer m and the isometry class of V0 are
uniquely determined by the isometry class of V .
The number m, which is the number of copies of H in the above decomposition, is called
the Witt Index of V and is usually denoted by Ind(V ).
Definition 1.22 A quadratic space is called hyperbolic if it is an orthogonal sum of copies
of H.
Proposition 1.23 Let (V, B) be a nondegenerate quadratic space, and W be a totally
isotropic subspace of dimension k. Then W is contained in a hyperbolic subspace of V
of Witt index k.
Proof. Let {e1 , . . . , ek } be a basis for W . Since V is nondegenerate, there exists f1 ∈ V
such that B(e1 , f1 ) = 1 and B(f1 , ei ) = 0 for i = 2, . . . , k. The subspace H1 = F e1 ⊕ F f1
is isometric to H; hence V = H1 ⊥ V1 for some nondegenerate subspace V1 . Moreover,
e2 , . . . , ek are vectors in V1 which span a totally isotropic subspace. An induction on the
dimension of W will complete the proof of the proposition.2
Corollary 1.24 The dimension of a maximal totally isotropic subspace of a nondegenerate
quadratic space V is equal to the Witt index of V .
At last, the following lemma is useful when deciding which element in F is represented
by V .
Lemma 1.25 Let a be a nonzero element in F . If V is nondegenerate, then a is represented
by V if and only if V ⊥ h−ai is isotropic.
Proof. Let F e be a 1-dimensional quadratic space over F with Q(e) = −a. Suppose that
a is represented by V . Then there exists x 6= 0 in V such that Q(x) = a. Therefore,
Q(x + e) = 0 which implies that V ⊥ h−ai is isotropic.
Conversely, suppose that V ⊥ F e is isotropic. If V is isotropic, then we are done.
Otherwise, there exist x ∈ V \ {0} and t ∈ F × such that Q(x + te) = 0. Then Q(t−1 x) =
−Q(e) = a. 2

10
1.5 Quadratic spaces over C
The field of complex numbers C is algebraically closed. Therefore, every nonzero complex
number is a square. If a is a nonzero complex number, then hai ∼
= h1i ∼
= h−1i over C.
Let (V, B) be an n-dimensional nondegenerate quadratic space over C. Let r be the
greatest integer smaller than or equal to n/2. There are 2r pairwise orthogonal vectors
x1 , . . . , xr , y1 , . . . , yr of V such that

Q(xi ) = 1, Q(yj ) = −1, 1 ≤ i, j ≤ r.

By Theorem 1.18, the binary subspaces Cxi + Cyi are all isometric to H. This implies that
V has an orthogonal decomposition isometric to Hr ⊥ V0 , where V0 ∼ = h1i or 0. There are
several implications of this observation:

(i) All nondegenerate quadratic spaces over C of dimension ≥ 2 are isotropic.

(ii) Every nondegenerate quadratic space V has maximal Witt index, that is, Ind(V ) is
always bn/2c.

(iii) The dimension of V determines the isometry class of V .

1.6 Quadratic Spaces over R


The quotient group R× /R×2 has two elements which are represented by 1 and −1, respec-
tively. Therefore, if V is a nondegenerate quadratic space, V can be decomposed as

V ∼
= h1ip ⊥ h−1iq .

The integer p is called the positive index of V and is denoted by Ind+ (V ). Similarly,
q = Ind− (V ) is the negative index of V . The difference Ind+ (V ) − Ind− (V ) is called the
signature of V . It is a consequence of the Witt’s Cancellation Theorem that both Ind+ (V )
and Ind− (V ), and hence the signature, depend only on the isometry class of V .

Theorem 1.26 (Sylvester’s Law of Inertia) Let V and W be two nondegenerate quadratic
spaces over R. Then V ∼
= W if and only if Ind+ (W ) = Ind+ (V ) and Ind− (W ) = Ind− (V ).

A quadratic space (V, Q) over R is said to be positive definite if Q(x) > 0 for all
x 6= 0. A negative definite quadratic space over R is defined analogously. A nondegenerate
quadratic space over R is called indefinite if it is neither positive definite nor negative
definite. In particular, an nondegenerate indefinite quadratic space over R must be isotropic
and universal.

11
1.7 Quadratic Spaces over Finite Fields
Let F be a finite field of q elements. We always assume that q is odd. The number of square
classes in F× is 2. Let ∆ be a fixed nonsquare element in F× .

Proposition 1.27 For n ≥ 2, every nondegenerate n-dimensional quadratic space over F


is universal.
Proof. It suffices to prove the proposition for a binary quadratic space (V, B) over F. We
may assume that V ∼ = hδ, i where δ,  ∈ {1, ∆}. If V ∼ = h1, ∆i, then we are done. If
V ∼= h∆, ∆i, the set Q(V ) is equal to ∆·Q(h1, 1i). Therefore, we may assume that V ∼
= h1, 1i.
∼ ∼
Our goal is to show that V represents ∆. If −1 is a square, then h1, 1i = h1, −1i = H which
is universal. Therefore, We may further assume that −1 is a nonsquare.
The sets F×2 and 1 + F×2 have the same number of elements. They are not equal since
1 is not inside 1 + F×2 . Therefore, there exists α ∈ F× such that 1 + α2 is not in F×2 .
This element 1 + α2 cannot be zero because −1 is not a square. Hence V represents a
nonsquare.2

Corollary 1.28 Let V be a nondegenerate quadratic space over F. Then

V ∼
= h1i ⊥ · · · ⊥ h1i ⊥ hd(V )i.

Proof. We have a decomposition V ∼


= h1i ⊥ V0 whenever V is universal.2

Corollary 1.29 Every nondegenerate quadratic space of dimension ≥ 3 over F is isotropic.


Proof. Let (V, B) be a quadratic space of dimension ≥ 3 over F. We may assume that
dim(V ) = 3; hence V has a decomposition V ∼ = h1, 1, d(V )i. Since h1, d(V )i is universal, it
represents −1. As a result, V contains a binary subspace which is isometric to h1, −1i ∼ = H.
Thus V is isotropic.2

Theorem 1.30 Let V and W be nondegenerate quadratic spaces over F. Then V ∼ = W if


and only if d(V ) = d(W ) and dim(V ) = dim(W ).
Proof. If V ∼= W , then of course d(V ) is equal to d(W ) and dim(V ) = dim(W ). The
converse is a consequence of Corollary 1.28.2

2 The p-adic Numbers


2.1 Valuations
Let F be a field whose characteristic is different from 2. A valuation on F is a function | |
of F into R which satisfies:

V1 |a| > 0 if a 6= 0, |0| = 0;

12
V2 |ab| = |a||b|;

V3 |a + b| ≤ |a| + |b|,

for all a, b ∈ F . A function which satisfies V1, V2 and

V3’ |a + b| ≤ max{|a|, |b|}

will satisfy V3 and therefore is a valuation. Axiom V3 is called the triangle inequality and
V3’ is called the ultra triangle inequality. A valuation is called nonarchimedean if it satisfies
V3’; otherwise it is called archimedean.
There is always a valuation on F , namely the trivial valuation obtained by putting
|a| = 1 for all a ∈ F × . Since this valuation has no significant interest, therefore we assume
that every valuation in the subsequent discussion is nontrivial.
Given a valuation | | on F , the distance function d(a, b) = |a − b| makes F into a metric
space. By a completion of F with respect to d (or | |) we mean a field F̂ together with a
valuation which satisfies

(i) F̂ is complete with respect to the given valuation;

(ii) F is a subfield of F̂ and the given valuation on F̂ extends | |;

(iii) F is dense in F̂ .

Theorem 2.1 A completion of F with respect to a valuation exists


Proof. We know from topology that as a metric space F has a completion, that is, there is
a metric space F̂ which is complete and contains F as a dense subset. Moreover, the metric
on F̂ induces the metric d on F . Without causing any confusion, we denote the metric on
F̂ also by d. We have to define addition and multiplication on F̂ to make it become a field.
Let α and β be two elements in F̂ . They are the limits of two Cauchy sequences {an } and
{bn } in F , respectively. It is obvious to see that both {an + bn } and {an bn } are Cauchy and
hence they converge to some elements in F̂ . Define

α + β = lim(an + bn ), αβ = lim(an bn ).
n n

One can check that these definition are independent of the choices of {an } and {bn }. Take
the original 0 and 1 of F as the 0 and 1 of F̂ . These all together make F̂ into a field.
Finally define kαk = d(α, 0) for all α ∈ F̂ . It is clear that k k extends the original
valuation on F . Suppose that α is the limit of a Cauchy sequence {an } of F . By the
triangle inequality,
−d(an , α) ≤ d(an , 0) − d(α, 0) ≤ d(an , α).
Since |an | − kαk = d(an , 0) + d(α, 0), therefore

lim |an | = kαk.


n

13
Hence, if bn −→ β, then
kαβk = lim |an bn | = lim |an | lim |bn | = kαkkβk.
n n n

Similarly kα +βk ≤ kαk+kβk. Therefore, k k is a valuation on F̂ , and the metric associated


with k k is d because
kα − βk = lim |an − bn | = lim d(an , bn ) = d(α, β).
n n
2

Remark 2.2 Note that k k is nonarchimedean if | | is nonarchimedean.


From now on, by abuse of notation, the valuation on F̂ that extends | | on F is also
denoted by | |.

2.2 Nonarchimedean Valuations


Let | | be a nonarchimedean valuation on a field F . By Remark 2.2, the extension of | | to
F̂ is also nonarchimedean. Here are some consequences of the ultra triangle inequality V3’:
(1) (Principle of Domination) Let α1 , . . . , αn ∈ F̂ . If |αi | < |α1 | for all i, then
|α1 + · · · + αn | = |α1 |.
P
(2) Let {αn } be a sequence of elements of F̂ . If αn converges, then of course αn −→ 0.
Conversely, suppose that αn −→ 0. Then for all M ≥ N ,
|αN + · · · + αM | ≤ max{|αj | : N ≤ j ≤ M }.
P P
Therefore, the partial sums of αn form a Cauchy sequence. Thus αn converges
to some element in F̂ .
(3) Let α ∈ F̂ × . There exists a sequence {an } of elements in F which converges to α.
Therefore, for all sufficiently large n, |an − α| < |α|. This shows that |an | = |α| for
all sufficiently large n. This in particular shows that the two sets |F | and |F̂ | are the
same.
(4) Let o be the subset of F containing all elements of F with valuation ≤ 1. The principle
of domination implies that o is a subring of F . Let ô be the closure of o in F̂ . It
follows from (3) ô is a subring of F̂ . It is called the valuation ring of F̂ . Since F̂ is
complete and ô is closed, ô itself is a complete metric space.
(5) Let α ∈ ô such that |α| = 1. Then α 6= 0 and |α−1 | = 1 also. This means that α−1 is
in ô and hence α is a unit of ô. This shows that ô has only one maximal ideal
p = {x ∈ ô : |x| < 1}.
Note that the group of units of ô is precisely the set ô \ p.

14
The valuation | | is called a discrete valuation if the image the composition
log
F̂ × −→ R+ −→ R

is an infinite cyclic subgroup of R. Let π ∈ ô be a pull back of a generator of this cyclic


group. Such an element is called a prime element of F̂ . The ideal p is generated by π and
ô is a principal ideal domain. Every element in F̂ × is of the form π n u for some n ∈ Z and
a unit u of ô. Let C be a complete set of representatives of cosets of p in ô. We always pick
0 to represent p.

Proposition 2.3 Suppose that | | is a discrete valuation on F . Let C be a complete set of


representatives of p in ô as described above. Then every element α ∈ F̂ can be expressed
uniquely by a Laurent series X
cj π j
j≥n

where cj ∈ C for all j, |α| = |π|n and cn 6= 0.


Proof. We may assume that α 6= 0. Suppose that α = π n u for some unit u of ô. Choose
cn ∈ C such that u ≡ cn mod p. Then cn 6= 0 and

α = cn π n + α1

with |α1 | ≤ |π|n+1 . Next apply this procedure to α1 to obtain an α2 , then to α2 , and so on.
From this procedure we obtain a sequence cn , cn+1 , . . . of elements of C such that for each
m ≥ n, there exists βm+1 ∈ pm+1 and

α = cn π n + cn+1 π n+1 + · · · + cn+m π n+m + βm+1 .

The partial sum


cn π n + · · · + cn+m π n+m
clearly converges to α.
Suppose that there are two Laurent series
X X
cj π j = dj π j
j≥n j≥n

with cj , dj ∈ C for all j but ci 6= di for some i ≥ j. Let i be the smallest index with this
property. Then ci − di 6∈ p which means that ci − di is a unit of ô. Then

X X
j
dj π = |π|i ,
j

0= cj π −
j≥n j≥n

which is impossible.2

15
2.3 Valuations on Q
Our primary objects of investigation here are the valuations of Q. The usual absolute value
is an archimedean valuation on Q. We denote it by | |∞ . The completion of Q with respect
to | |∞ is the field of real numbers R.
Beside | |∞ , Q has other valuations. Let p be a prime number. Any α ∈ Q× can be
written as
a
α = pi
b
where a and b are integers prime to p. Put
1
|α|p = .
pi
It is easy to show that this defines a discrete nonarchimedean valuation on Q. The comple-
tion of Q with respect to | |p is the field of p-adic numbers Qp . Its valuation ring is the ring
of p-adic integers Zp . The topology on Q or Qp induced by | |p is called the p-adic topology.
In the p-adic topology, every pn Zp is both open and closed.

Lemma 2.4 The maximal ideal of Zp is generated by p.


Let α be a p-adic integer. Since Zp is the closure of Z under the p-adic topology, therefore
there exists a ∈ Z such that α ≡ a mod pZp . Hence we can choose C = {0, 1, . . . , p − 1}
to be a complete set of representatives of Zp /pZp . In this way, every p-adic number can be
represented uniquely as a Laurent series in p

X
cj pj
j=n

with cj ∈ C.

Corollary 2.5 Zp /pZp is a finite field of p elements.


Proof. Let α ∈ Zp . Then α can be represented by a Taylor series in p

X
α= cj pj
j=0

with cj ∈ C. The function α 7→ c0 mod p is clearly a surjective ring homomorphism from


Zp onto Z/pZ with kernel pZp .2

Corollary 2.6 Zp is a compact topological space.


Proof. Let {Oλ : λ ∈ Λ} be an opening covering of Zp . Suppose that it has no finite
subcovering. Note that
p−1
[
Zp = (i + pZp ).
i=1

16
Therefore, there exists c0 ∈ C = {0, 1, . . . , p − 1} such that c0 + pZp is not covered by finitely
many of the Oλ . Similarly, there exists c1 ∈ C such that c0 + c1 p + p2 Zp is not finitely
covered. By continuing this process, we can construct a sequence c0 , c1 . . . of elements of C
such that for each j ≥ 0,
c0 + c1 p + · · · + cj pj + pj+1 Zp
is not finitely covered. Let

X
α= cj pj .
j=0

Then α ∈ Zp , and α must be in Oλ0 for some λ0 ∈ Λ. Since Oλ0 is open, there exists m ≥ 1
such that α + pm Zp ⊆ Oλ0 . But then

c0 + c1 p + · · · + cm−1 pm−1 + pm Zp ⊆ Oλ0

which is a contradiction.2

Definition 2.7 Two valuations | | and k k on a field F are said to be equivalent if there
exists a nonzero real number k such that | |k = k k.
Equivalent valuations on a field F define the same topology on F . It is not hard to see
that the absolute value | |∞ and the p-adic valuations | |p are inequivalent valuations on Q.

Theorem 2.8 (Ostrowski) Every nontrivial valuation of Q is equivalent to | |∞ or | |p for


some prime p.
Before giving the proof of Ostrowski’s theorem, we need a lemma to characterize the
nonarchimedean valuations on Q.

Lemma 2.9 A valuation k k on Q is nonarchimedean if and only if it is bounded on Z.


Proof. Suppose that k k is nonarchimedean. Then for any positive integer n,

knk = k1 + · · · + 1k ≤ k1k = 1.

Therefore, k k is bounded on Z.
Conversely, suppose that kmk < K for all m ∈ Z. Then for any positive integer k and
any rational numbers x, y,

kx + ykk ≤ K(|xkk + kxkk−1 kyk + · · · + kykk ) ≤ K(k + 1)max{kxk, kyk}k .

Taking the k-th root on both sides and letting k −→ ∞, we see that k k is nonarchimedean.2
Proof of Ostrowski’s theorem. Let k k be a nonarchimedean valuation on Q. Then knk ≤ 1
for all n ∈ Z. So, there must be a prime number p such that kpk < 1 because, if not, the
Fundamental Theorem of Arithmetic would imply kxk = 1 for all x ∈ Q× .

17
The set a = {a ∈ Z : kak < 1} is an ideal of Z satisfying pZ ⊆ a 6= Z. Thus pZ = a. If
a ∈ Z and a = pm b with gcd(p, b) = 1, then kbk = 1 and hence

kak = kpkm = |a|sp

where s = − log kpk/ log p. Consequently k k is equivalent to | |p .


Now, suppose that k k is archimedean. Let m > 1 and n > 1 be two natural numbers.
Then
m = a0 + a1 n + · · · + ar nr
where ai ∈ {0, 1, . . . , n − 1} and nr ≤ m. Observe that r ≤ log m/ log n and kai k ≤ ai k1k ≤
n. So, kmk < n(1 + knk + · · · + knkr ). If knk < 1, then kmk < n/(1 − knk) and this is true
for all m ∈ Z. This contradicts that k k is archimedean. Therefore, knk ≥ 1 and we obtain
the inequality
n  
X
r log m
kmk ≤ kai k · knk ≤ 1 + n · knklog m/ log n .
log n
i=1

Substituting here mk for m, taking k-th roots on both sides, and letting k tend to ∞, one
finally obtains
kmk ≤ knklog m/ log n , or kmk1/ log m ≤ knk1/ log n .
Interchanging the roles of m and n gives

kmk1/ log m = knk1/ log n .

Putting c = knk1/ log n and pick s so that c = es . Then for any positive rational number x,

kxk = es log x = |x|s .

Therefore k k is equivalent to | |∞ .2

Proposition 2.10 (Product Formula) For any a ∈ Q× , we have


Y
|a|∞ |a|p = 1.
p

Proof. The proof is obvious. Note that |a|p = 1 for almost all p.2

2.4 Square Classes of Qp


Let p be a prime number. It is clear that every square class of Q×
p must contain either a
unit or a prime element, but not both.

18
Lemma 2.11 (Local Square Theorem)Let α ∈ Q×
p and suppose that

|α − 2 |p ≤ |4p|p
for some unit  in Zp . Then α is a square in Z×
p.

Proof. The hypothesis implies that α is in Z× p . Write  as 0 and suppose that for any
non-negative integer k ≤ n, there is k ∈ Z×
p such that
α ≡ 2k mod 4pk+1 .
Let b ∈ Z such that
α − 2n
≡ bn mod p.
4pn+1
Putting n+1 = n + 2bpn+1 gives
α − 2n
 
α − 2n+1 = 4pn+1 − bn − 4b2 p2(n+1) .
4pn+1
The choice of b implies α ≡ 2n+1 mod 4pn+2 . Thus we have proved the existence of a
sequence {n } in Z× 2
p such that α ≡ n mod 4p
n+1 and 
n+1 ≡ n mod p
n+1 for every n ≥ 0.

Now, for any m > n,


|m − n |p = |(m − m−1 ) + · · · + (n+1 − n )|p
≤ p−(n+1)
Therefore {n } is a Cauchy sequence and it must converge to some x in Zp . As α ≡ 2n+1
mod 4pn+2 for all n, we see that α = x2 .2
Corollary 2.12 For any α ∈ Q× ×2 ×
p , the set αQp is an open subset of Qp .

Proof. Let αy 2 ∈ αQ×2 × 2 2


p . For any x ∈ Qp such that |x − αy |p ≤ |4pαy |p , we have

x
αy 2 − 1 ≤ |4p|p .

p

Therefore x(αy 2 )−1 is a square and hence x ∈ αQ×2


p .2

Corollary 2.13 Q× ×2
2 /Q2 is of order 8 and generated by 2, −1 and 5. For p 6= 2, let ∆ be
a non-square unit in Zp . Then Q× ×2
p /Qp is of order 4 and generated by p and ∆.

Proof. The assertion for p 6= 2 is clear since (Z/pZ)× has only


P∞2 square classes represented
2 × j
by 1 and ∆ respectively. Suppose α = x in Z2 . Let x = j=0 cj 2 with cj ∈ {0, 1} and
c0 = 1. Then
 2

X X∞
α = (c0 + 2c1 )2 + 2(c0 + 2c1 ) cj 2j +  cj 2j 
j=2 j=2
2
≡ (c0 + 2c1 ) mod 8
≡ 1 mod 8.

19
Therefore, α ∈ Z×2
2 if and only if α ≡ 1 mod 8. Thus the square classes represented by the
units are generated by −1 and 5.2
In the following proposition, δ ∈ {∆, p∆, p} if p > 2, and δ ∈ {−1, 3, 5, 2, −2, 6, 10} if
p = 2.

Proposition 2.14 (a) If p > 2, then x2 − δy 2 ∈ Zp if and only if x and y are in Zp .


(b) If p = 2 and δ 6= 5, then x2 − δy 2 ∈ Z2 if and only if x, y ∈ Z2 .
(c) If p = 2 and δ = 5, then x2 − 5y 2 ∈ Z2 if and only if x = s/2, y = t/2, s, t ∈ Z2 and
s − t ∈ 2Z2 .
Proof. (a) We assume that x2 −δy 2 ∈ Zp . The statement is clear if either x or y is zero. Thus
we further assume that xy 6= 0. Consider the case when δ = ∆ first. Suppose |x|p ≥ |y|p .
Since the space h1, −∆i is anisotropic over Zp /pZp ∼ = Z/pZ, therefore 1 − (y/x)2 ∆ is always
a unit in Zp . In other words, |x|p ≤ 1 and hence |y|p ≤ 1 also. Similar argument applies to
the case when |x|p ≤ |y|p . Now suppose δ = p with  = 1 or ∆. If |x|p ≥ |y|p , then
y
|x2 − y 2 p|p = |x|2p 1 − ( )2 p = |x|2p .

x p

Thus both x and y are in Zp . If, on the other hand, |y|p ≥ |x|p , then we consider
 2 !
x
x2 − y 2 p = y 2 − p
y

and deduce that |(x/y)2 − p|p is either 1 or |p|p . Therefore, y is in Zp and hence so is x.
For (b), we assume that x2 − y 2 δ ∈ Z2 and xy 6= 0. If 2 | δ, then the above argument
for p > 2 can apply here. Suppose δ = −1. If y/x ∈ Z2 , then (y/x)2 ≡ 0 mod 4 or 1 mod 8.
So, 1 + (y/x)2 is a unit or it is congruent to 2 mod 8. The former implies that |x|22 ≤ 1 and
thus x ∈ Z2 . This implies y ∈ Z2 as well. If 1 + (y/x)2 ≡ 2 mod 8, then 2x2 ∈ Z2 which
implies x ∈ Z2 . Thus y ∈ Z2 also. Same argument works when x/y ∈ Z2 . The case δ = 3
can be done in a similar manner.
For (c), it is clear that if x = s/2 and y = t/2 with s, t ∈ Z× 2 2
2 , then x − 5y ∈ Z2 and
xy 6= 0. Conversely, suppose x2 − 5y 2 ∈ Z2 . If y/x ∈ Z2 , then 1 − 5(y/x)2 is either a unit
or ≡ 4 mod 8. The first option implies that x ∈ Z2 and hence y ∈ Z2 as well. The second
happens only if y/x is a unit and x = s/2 with s ∈ Z2 . So, y = t/2 with t ∈ Z2 . Note that
s − t ∈ 2Z2 whenever |s|2 = |t|2 . The argument is similar when x/y ∈ Z2 .2

2.5 Quadratic Reciprocity


Let p be an odd prime and Fp be a finite field of p elements. For any x ∈ F×
p , define
  
x 1 if x is a square,
:=
p −1 otherwise.
It is called the Legendre symbol mod p. Its definition is often extended to integers that are
relatively prime to p. It is clear that the Legendre symbol is multiplicative in x.

20
Lemma 2.15 For any a ∈ F×
p,
 
p−1 a
a 2 = in Fp .
p

(p−1)n
Proof. Let ξ be a generator of F×
p . For any integer n, the order of ξ 2 is 1 if n is even
(p−1)n
and 2 otherwise. Therefore ξ 2is 1 if n is even and −1 otherwise. The lemma follows
from the fact that a can be written as ξ n for some n and a is a square if and only if n is
even.2

Lemma 2.16 For any odd prime p,


   
−1 p−1 2 p2 −1
= (−1) ,
2 = (−1) 8 .
p p

Proof. The first equality is clear. For the second one, let ξ be a primitive 8-th root of unity
in the algebraic closure of Fp . Then ξ 4 = −1 and ξ 2 + ξ −2 = 0. Therefore, if y = ξ + ξ −1 ,
then y 2 = 2. Hence 2 is a square in Fp if and only if y ∈ Fp , or equivalently, y p = y. The
last condition is satisfied if and only if ξ p+1 (ξ p−1 − 1) = ξ p−1 − 1, which is the same as
ξ p−1 = 1 or ξ p+1 = 1. Therefore, 2 is a square in Fp if and only if p ≡ ±1 mod 8. 2

Theorem 2.17 (Quadratic Reciprocity Law) Let q 6= p be two odd primes. Then
  
q p p−1 q−1
= (−1) 2 2 .
p q

 
Proof. For simplicity, we put χ(a) = for any a ∈ F×
a
p p and χ(0) = 0. Since half of the
×
P
elements in Fp are squares and the other half consists of non-squares, b∈Fp χ(b) = 0. Let
x be a primitive p-th root of unity in the algebraic closure of Fq . So it makes sense to talk
about xa for any a ∈ Fp ∼= Z/pZ. Put
X
g := χ(a)xa .
a∈Fp

Then X X X
g2 = χ(ab)xa+b = xc χ(a(c − a)).
a,b∈Fp c∈Fp a∈Fp

21
If c 6= 0, then we have
X X
χ(a(c − a)) = χ(−1) χ(a2 − ac)
a∈Fp a∈F×
p

X
= χ(−1) χ(1 − a−1 c)
a∈F×
p

X
= χ(−1) χ(1 − a)
a∈F×
p

X
= χ(−1) χ(b) − χ(−1)
b∈Fp

= −χ(−1).
Thus X X
g 2 = −χ(−1) xc + χ(−a2 ) = χ(−1) + (p − 1)χ(−1) = pχ(−1).
c6=0 a∈Fp

On the other hand, if we raise g to the q-th power, we get


X X
gq = χ(a)xqa = χ(q) χ(qa)xqa = χ(q)g
a∈Fp a∈Fp

and so g q−1 = χ(q). Therefore, in Fq ,


 
q−1 q−1 q
χ(−1) 2 p 2 = .
p
q−1 p−1
 
p
The left hand side is equal to (−1) 2 2
q in Fq . Therefore
  
p−1 q−1 q p
(−1) 2 2 = in Fq .
p q
This can be seen as an equality in Z because both sides are ±1.2

2.6 Hilbert Symbols and Hilbert Reciprocity


In this subsection, let F be either R or Qp for some prime p. Let a, b be two nonzero
elements in F . The Hilbert symbol (a, b)F is defined as

1 if ha, b, −1i is isotropic over F ,
(a, b)F =
−1 otherwise.
For simplicity, we will use (a, b) unless confusion arises. It is clear from the definition that
the Hilbert symbol is a symmetric function defined on F × /F ×2 × F × /F ×2 . Moreover,
(a, b) = 1 whenever a or b is a square.

22
Lemma 2.18 For any a, b, c ∈ F × ,

(a) (a, b) = (b, a), (a, b2 ) = 1.

(b) (a, −a) = (a, 1 − a) = 1 if a 6= 1.

(c) (a, a) = (a, −1).

(d) (a, bc) = (a, b)(a, c).

Proof. Part (a) and (b) are obvious. For part (c), since a 6= 0, ha, a, −1i is isotropic if and
only if h−1, −1, ai is isotropic.
For part (d), we may assume that a is not a square. The case F = R can be verified
directly: (−1, d) = 1 if and only if d > 0.
When F = Qp , let
Na = {x2 − ay 2 6= 0 : x, y ∈ F }.
Note that Na is in fact a subgroup of F × . Observe that (d, a) = 1 if and only if d ∈ Na .
Therefore, part (d) is valid if either b or c is in Na . It remains to show that bc ∈ Na if
b, c 6∈ Na . It suffices to show that [F × : Na ] ≤ 2.
Suppose p > 2. Then we may assume that a ∈ {∆, p, p∆}, where ∆ is a nonsquare
unit in Z× p . If a = p or p∆, then −a ∈ Na , which implies [Na : F
×2 ] ≥ 2 and thus
×
[F : Na ] ≤ 2. If a = ∆, the space h∆, −1, −1i is isotropic over Zp /pZp but h∆, −1i is
anisotropic. Therefore, there exist z ∈ Z× 2 2
p and x, y ∈ Zp such that (∆x − y ) − z ≡ 0
2
2 2 2 2
mod p. By Lemma 2.11, ∆x − y itself must be a square and hence ∆x − y = t for some 2

t ∈ Zp . Therefore, ∆t2 ∈ Na and [Na : F ×2 ] ≥ 2.


If p = 2, we have to show that [Na : F ×2 ] ≥ 4. It suffices to produce 3 different square
classes in Na . We may assume that a is coming from {3, 5, −1, 2, 6, 10, −2}. If a is divisible
by 2, the numbers 1, −a, 1−a fall into different square classes in Na . When a is not divisible
by 2, then 1, 1 − a, 1 − 4a will do the job.2.
Note that even when a is a square, the assertion that (d, a) = 1 if and only if d ∈ Na is
still true.
For a ∈ Q× m ×
p , we can write a = p u with m ∈ Z and u ∈ Zp . The integer m is called the
order of a, denoted ordp (a).

Lemma 2.19 Let ∆ be a nonsquare unit in Zp if p > 2 and ∆ = 5 if p = 2. Then,



1 if ordp (a) is even,
(∆, a)Qp =
−1 otherwise.

Proof. Obviously, we may assume that ordp (a) is 0 or 1. We first treat the case when
p > 2. Since the binary space h1, −∆i is anisotropic over Zp /pZp , therefore if x2 − ∆y 2 ≡ 0
mod p, both x and y are ≡ 0 mod p and hence ordp (x2 − ∆y 2 ) ≥ 2. In this case, a 6∈ N∆
and so (a, ∆)Qp = −1. On the other hand, suppose ordp (a) = 0. Since the space h1, −∆i

23
is universal over Zp /pZp , there exist x, y ∈ Z× 2 2
p such that x − ∆y ≡ a mod p. By the
×
Local Square Theorem, there exist z ∈ Zp such that x − ∆y = az 2 , which implies that
2 2

(a, ∆)Qp = 1.
Let us consider the case p = 2. Suppose ord2 (x2 − 5y 2 ) = 1. If x, y ∈ Z2 , then both of
them must be units in Z2 . So x2 − 5y 2 ≡ 4 mod 8 which is a contradiction. This implies
that x = s/2 and y = t/2 for some units s, t and s2 − 5t2 ≡ 0 mod 8. But this is impossible
because s2 − 5t2 ≡ 4 mod 8 for any units x, y. Therefore, 2 6∈ N5 and so (5, 2)Q2 = −1.
If a ∈ Z× 2 2
2 , we can always find x, y ∈ Z2 such that x − 5y ≡ a mod 8. By the Local
Square Theorem again, we see that a ∈ N5 and so (5, a)Q2 = 1. 2

Corollary 2.20 If p > 2 and , δ ∈ Z×


p , then (, δ)Qp = 1.

Corollary 2.21 If p > 2 and  ∈ Z×


p , then

1 if  is a square,
(, p)Qp =
−1 otherwise.
 

Consequently, (, p)Qp is the Legendre symbol p , where  is the class of  mod pZp .

Theorem 2.22 (Hilbert Reciprocity Law) For any a, b ∈ Q× , we have


Y
(a, b)R (a, b)Qp = 1.
p

Proof. Put Y
Fa (x) := (a, x)R (a, x)Qp .
p

Since (a, x)Qp = 1 for almost all p, therefore Fa (x) is well defined. Moreover, Fa (xy) =
Fa (x)Fa (y), Fa (x) = Fx (a) and Fa (a) = Fa (−1). So it suffices to show that F−1 (−1),
F−1 (2), F−1 (p), F2 (p), and Fp (q) are all equal to 1, where p, q are odd prime numbers.
Note that (a, b)R = −1 if and only if both a and b are negative. If a is a unit in Z2 , then
x2 + y 2 = a is solvable in Q2 if and only if a ≡ 1 mod 4. Therefore, (−1, −1)Q2 = −1 and
p−1
(−1, p)Q2 = (−1) 2 . The equation x2 − 2y 2 = p is solvable in Q2 if and only if p ≡ ±1 mod
p2 −1
8. Therefore (p, 2)Q2 = (−1) 8 and in particular (−1, 2)Q2 = 1. Together with Lemma
2.16 and Corollary 2.20, these show that

F−1 (−1) = F−1 (2) = F−1 (p) = F2 (p) = 1.

For Fp (q) = 1, we have to show that (p, q)Q2 = 1 if and only if p or q is congruent to 1 mod
4. We first assume that either p or q is congruent to 1 mod 4. Without loss of generality,
suppose p ≡ 1 mod 4. We may further assume that p is either 1 or 5. Clearly, (1, q)Q2 = 1
for all q and Lemma 2.19 shows that (5, q)Q2 = 1 as well.

24
Now suppose that (p, q)Q2 = 1. So there exist x, y ∈ Q2 such that x2 − py 2 = q. Assume
that p and q are ≡ 3 mod 4. Then x and y are in Z2 . Modulo 4, we have

3 ≡ q ≡ x2 − py 2 ≡ 1 mod 4

which is a contradiction. Thus either p or q is ≡ 1 mod 4.2


For our convenience, we extract the following from the proof above.

Corollary 2.23 If a and b are odd integers, then



(a, b) = (−1) (a−1) 2
(b−1)
2 ,
Q2
a2 −1
(a, 2)Q = (−1) 8 .
2

Our last remark is that if a ∈ Q× ×


p which is not a square, then ( , a) : Qp → {±1} is a
surjective group homomorphism with kernel Na .

3 Quadratic Spaces over Qp


3.1 Hasse Invariants
Within this subsection, F is either R or Qp for some prime p. Let (V, B) be a nondegenerate
quadratic space over F . Take an orthogonal basis for V and suppose

V ∼
= hα1 , · · · , αn i

in this basis. We define the Hasse invariant of V , written as SF (V ) or simply S(V ) if F is


clear from the context, to be Y
(αi , αj ),
1≤i≤j≤n

where all the Hilbert symbols are over F .

Theorem 3.1 S(V ) is independent of the choice of the orthogonal basis.


Proof. If n = 1 and V ∼
= hαi, then (α, α) = (α, −1) = (d(V ), −1) which depends only on V .
Suppose n = 2. A direct computation shows that
Y
(αi , αj ) = (d(V ), −1)(α1 , α2 ).
i≤j

Now, (α1 , α2 ) = 1 if and only if hα1 , α2 , −1i ∼


= V ⊥ h−1i is isotropic. This is certainly
dependent on V , not on the basis chosen.

25
When n ≥ 3, we need a couple of definitions. Let {vi } and {ui } be two orthogonal bases
for V . We write {vi } ∼ {ui } if F vi = F uj for some i, j. We say that {vi } and {ui } are
linked if there exist orthogonal bases {zij } of V such that

{vi } = {zi1 } ∼ {zi2 } ∼ · · · ∼ {zim } = {ui }.

Claim: If dim(V ) ≥ 3, then every pair of orthogonal bases for V are linked.
Let us show how the theorem follows from the claim. Using the claim, we may assume
that
V ∼
= hα1 , . . . , αn i ∼
= hβ1 , . . . , βn i
and αh = βk for some h, k. It suffices to show that S(V ) is unchanged in the following two
situations:

(i) αk = β1 for some k 6= 1, and αi = βi for all i 6= 1, k. Note that we could then assume
that α1 = βk since α1 · · · αn and β1 · · · βn are in the same square class.

(ii) α1 = β1 .

For a fixed h,
Y Y Y Y
(αi , αj ) = (αh , αj ) (αi , αh ) (αi , αj )
i≤j h≤j i<h i≤j;i,j6=h
Y
= (αh , d(V )) (αi , αj ).
i≤j;i,j6=h

In case (i), we have


Y Y
(αi , αj ) = (α1 , d(V )) (αi , αj )
i≤j 1<i≤j
Y
= (α1 , d(V ))(αk , α1 d(V )) (αi , αj )
i≤j;i,j6=1,k
Y
= (α1 αk , d(V ))(αk , α1 ) (αi , αj )
i≤j;i,j6=1,k
Y
= (β1 βk , d(V ))(βk , β1 ) (βi , βj )
i≤j;i,j6=1,k
Y
= (βi , βj )
i≤j

In case (ii), if α1 = β1 , then


Y Y
(αi , αj ) = (α1 , d(V )) (αi , αj ) = (α1 , d(V ))S(hα2 , . . . , αn i).
i≤j 1<i≤j

26
Witt’s Cancellation Theorem implies that hα2 , . . . , αn i is isometric to hβ2 , . . . , βn i. There-
fore, case (ii) can be completed by an induction on dim(V ).
Proof of the claim Write v1 = a1 u1 + · · · + am um , where ai ∈ F and am 6= 0, and define
ν({ui }) = m. Observe that m = 1 means that {vi } and {ui } are linked. Therefore, all we
need to show is that if m > 1, then there exists another orthogonal basis {u0i } such that
{u0i } ∼ {ui } and ν({u0i }) < m.
Suppose aj = 0 for some j < m. Define {u0i } as follows:

ui
 if i < j,
0
ui = ui+1 if j ≤ i < n,

uj if i = n.

Then {u0i } and {ui } are linked and ν({u0i }) = m − 1. Thus we may assume that aj 6= 0 for
all j < m. If m ≥ 3, then we may suppose Q(a1 u1 + a2 u2 ) = a21 Q(u1 ) + a22 Q(u2 ) 6= 0 since
one of
Q(a1 u1 + a2 u2 ), Q(a2 u2 + a3 u3 ), Q(a1 u1 + a3 u3 )
is not equal to 0 (if they were all zero, then Q(u1 ) = Q(u2 ) = Q(u3 ) = 0 which is impossible).
Then we can extend {a1 u1 + a2 u2 , u3 , . . . , um } to an orthogonal basis for V that is linked
to {ui }, and the value of ν for this basis is m − 1.
Suppose m = 2. Then {a1 u1 + a2 u2 , a2 Q(u2 )u1 − a1 Q(u1 )u2 , u3 . . . , } is an orthogonal
basis whose value of ν is equal to 1, and hence it is linked to {vi }. But it is obvious that
this basis is also linked to {ui }. 2.
Suppose that
V ∼
= hα1 , . . . , αn i.
Lemma 2.18 implies that Y
S(V ) = (αi , α1 · · · αi ).
1≤i≤n

For a nontrivial decomposition V = U ⊥ W , we obtain

S(V ) = S(U )S(W )(d(U ), d(W )).

For any α ∈ F × , we let V α be the space (V, Qα ) where Qα (x) = αQ(x) for all x ∈ V . We
call V α the scaling of V by α. The discriminant of V α is αn d(V ). Also,

S(V α ) = (α, (−1)n(n+1)/2 d(V )n+1 )S(V ).

3.2 Classification
In this subsection, let p be a prime. The Hilbert symbol over Qp is denoted by ( , ). If V
is a nondegenerate quadratic space over Qp , then the Hasse invariant of V is denoted by
S(V ).

27
Proposition 3.2 A nondegenerate ternary space V over Qp is isotropic if and only if
S(V ) = (−1, −1).
Proof. After a suitable scaling, we may assume that V ∼
= ha, b, −1i. Therefore V is isotropic
if and only if (a, b) = 1. However, S(V ) is equal to (a, b)(−1, −1). As a result, V is isotropic
if and only if S(V ) = (−1, −1).2.

Corollary 3.3 Let a, b, c be units in Z×


p with p > 2. Then ha, b, ci is isotropic.

Proof. The Hasse invariant of ha, b, ci is equal to 1 = (−1, −1).2


The above corollary is false if p = 2. For example, the ternary space h1, 1, 1i is anisotropic
over Q2 .

Proposition 3.4 A nondegenerate quaternary space V over Qp is isotropic if and only if


one of the following holds:

(a) d(V ) = 1 and S(V ) = (−1, −1);

(b) d(V ) 6= 1.

Proof. If V is isotropic, then V ∼ = H ⊥ W . Suppose d(V ) = 1. Then d(W ) = −1 and thus


W is also a hyperbolic plane. A computation shows that S(V ) = (−1, −1) in this case.
For the converse, we have to show that for an anisotropic quaternary space V , both
d(V ) = 1 and S(V ) = −(−1, −1) hold. Suppose V ∼ = ha, b, c, di. Then both h1, ab, ac, adi
and h1, cd, ad, bdi are anisotropic. Suppose that (x, −ab) = 1 for some x ∈ Q× p . Then
hx, −ab, −1i is isotropic or x is represented by the space hab, 1i. Hence x is not represented
by h−ac, −adi and so hx, ac, adi is anisotropic. By Proposition 3.2, we see that

(x, xcd)(ac, cd)(ad, ad) = −(−1, −1).

The ternary space h1, ac, adi is also anisotropic. Using Proposition 3.2 again we have
(ac, cd)(ad, ad) = −(−1, −1) and hence (x, −cd) = 1. We thus show:

(x, −ab) = 1 =⇒ (x, −cd) = 1,
(x, −cd) = 1 =⇒ (x, −ab) = 1.

Hence (x, −ab) = (x, −cd) for all x ∈ Q×


p . This implies that d(V ) = abcd must be a square.
Another consequence is that S(V α ) = S(V ) for all α ∈ Q×p . Therefore

S(V ) = S(h1, ab, ac, adi)


= S(hab, ac, adi)
= −(−1, −1)

since hab, ac, adi is anisotropic.2

28
Theorem 3.5 If the dimension of a quadratic space V over Qp is at least 5, then V is
isotropic.
Proof. We may assume that the dimension of V is 5 and that V is nondegenerate. Suppose
V ∼= ha, b, c, d, ei, and yet V is anisotropic. Proposition 3.4 implies that the product of
any four different elements chosen from {a, b, c, d, e} is a square. This implies that V ∼ =
ha, a, a, a, ai which is a scaling of h1, 1, 1, 1, 1i. So we may assume that V ∼
= h1, 1, 1, 1, 1i. If
p > 2, then V contains an isotropic subspace isometric to h1, 1, 1i. If p = 2, then
√ 2
12 + 12 + 12 + 22 + −7 = 0.

Therefore, V is isotropic. 2

Corollary 3.6 Every nondegenerate quaternary space over Qp is universal.


If two nondegenerate quadratic spaces over Qp are isometric, then their dimensions,
discriminants, and Hasse invariants are the same. The main theorem in this subsection is
that these three invariants determine the isometry class of a quadratic space over Qp .

Theorem 3.7 Let V and W be two nondegenerate quadratic spaces over Qp . Then V ∼
=W
if and only if dim(V ) = dim(W ), d(V ) = d(W ), and S(V ) = S(W ).
Proof. Let the dimension of V be n. If n = 1, then the theorem is trivial. If n = 2, we may
assume that d(V ) = d(W ) 6= −1; otherwise both V and W are isometric to the hyperbolic
plane and hence V ∼= W . Let U be the quaternary space W ⊥ V −1 . Then d(U ) = 1 and
S(U ) = (−1, −1) by a routine calculation. Therefore, U is isotropic by Proposition 3.4.
Since both V and W are anisotropic, we can find nonzero vectors v ∈ V , w ∈ W such that
Q(v) = Q(w) = a. Thus
V ∼
= ha, ad(V )i ∼
= W.
Suppose now that n ≥ 3. Let U be the space W ⊥ V −1 whose dimension is at least
6. Therefore, it is isotropic which implies that there is an a ∈ Q×
ν which is represented by
both V and W . Hence
V = hai ⊥ V1 , W = hai ⊥ W1 .
Direct computations show that d(V1 ) = d(W1 ) and S(V1 ) = S(W1 ). So, an induction
argument (on the dimension) would show that V1 and W1 are isometric. Therefore, V and
W are isometric.2

Corollary 3.8 If V is an anisotropic quaternary space over Qp , then

V ∼
= h1, −∆, p, −∆pi.

Proof. From Proposition 3.4, we have d(V ) = 1 and S(V ) = −(−1, −1). These two con-
ditions are satisfied by h1, −∆, p, −∆pi. Theorem 3.7 implies that such a space must be
unique up to isometry.2

29
4 Quadratic Spaces over Q
4.1 Strong Hasse Principle
Let Ω be the set of all prime numbers and the formal symbol ∞. We call an element in Ω
a place of Q; a prime number is called a finite place and ∞ is the infinite place. Let V be
a quadratic space. For each ν ∈ Ω, let Vν be the vector space Qν ⊗Q V over Qν , where we
put Q∞ = R. We call Vν the local completion of V at ν. We identify V as a subset of Vν
via the embedding v 7→ 1 ⊗ v.
Let v1 , . . . , vn be a basis for V . Define a quadratic map Qν on Vν by
n
X n
X
Qν ( xi vi ) := xi xj B(vi , vj ), xi ∈ Qν for all i,
i=1 i,j=1

where B is the bilinear form on V . It is clear that (Vν , Qν ) becomes a quadratic space over
Qν . Since Q and Qν are essentially the same, we will drop the subscript and use the same
Q to denote the quadratic form on Vν . Note that if A is a symmetric matrix for V , then A
is also a symmetric matrix for Vν . We say that V is isotropic at ν if Vν is isotropic. The
Hasse invariant of Vν over Qν will be written as Sν (V ). In particular, if V ∼= ha1 , . . . , an i,
then Y
Sν (V ) = (ai , aj )Qν .
1≤i≤j≤n

Lemma 4.1 Let V be a nondegenerate quadratic space over Q. Then


Y
Sν (V ) = 1.
v∈Ω

Proof. This follows from the Hilbert Reciprocity Law.2

Theorem 4.2 (Strong Hasse Principle) A nondegenerate quadratic space V over Q is


isotropic if and only Vν is isotropic for every ν ∈ Ω.
Proof. The only if part is clear. We prove the converse by an induction on n, the dimension
of V . There is nothing to verify when n = 1 since all one dimensional nondegenerate space
are anisotropic.
Suppose that n = 2 and that V ∼ = ha, bi. Since Vν is isotropic, −ab is a square in Qν for
every ν ∈ Ω. Hence −ab is a square in Q and V is isotropic.
When n = 3, we scale V so that V represents 1. Hence we can suppose

V ∼
= h1, −a, −bi.

Moreover, by multiplying a and b by squares, we may assume that a and b are integers. We
handle the problem by induction on |a| + |b|. Here | | is the usual absolute value on Q. If

30
|a| + |b| = 2, then |a| = |b| = 1. Since V∞ is isotropic, therefore either a or b is positive. As
a result, V ∼= h1, 1, −1i or h1, −1, −1i and V is isotropic.
Suppose |a| + |b| > 2 and |a| ≤ |b|. Then we must have |b| ≥ 2. Moreover, by dividing
squares if necessary, we may assume that both a and b are squarefree integers. If a = 1,
then V is obviously isotropic. So we can assume that a 6= 1. We claim that the congruence
x2 ≡ a mod b is solvable with x ∈ Z. Up to ±, b is a product of distinct primes. Therefore
it suffices to prove the claim for every prime divisor p of b. We just need to look at those p
which do not divide a. Moreover, we may solve the congruence over Zp instead of Z. Since
Vp is isotropic, there exist x, y, z ∈ Zp , not all zero, such that x2 − ay 2 − bz 2 = 0. We may
assume that one of them is a unit in Zp . Since b 6≡ 0 mod p2 , y must be in Z× p . Then
a ≡ (xy −1 )2 mod p and xy −1 ∈ Zp .
Now, there exists an integer c such that c2 = a + bd for some integer d. We can further
assume that |c| ≤ |b|/2. Note that d 6= 0 because a is not a square. Then

|c2 − a| |b|
|d| = ≤ + 1 < |b|
|b| 4

by virtue of |b| ≥ 2. Note that bd = c2 − a 12 which is a norm of the extension Q( a)/Q.

Therefore, bd is also a norm of F ( a)/F whenever F is a field extension of Q. This means

that b is a norm of F ( a)/F if and only if d is.
By the assumption, h1, −a, −bi is isotropic over Qν and it is equivalent to saying that
√ √
b is a norm of Qν ( a)/Qν , and hence d is a norm of Qν ( a)/Qν . Apply the induction

hypothesis to h1, −a, −di, we find that d is a norm of Q( a)/Q and hence b is also a norm

of Q( a)/Q. This means that V is isotropic.
When n ≥ 4, put
V =U ⊥W
where U = ha1 , a2 i and W = ha3 , . . . , an i. All the ai can be assumed to be integers. For
every ν ∈ Ω, Vν is isotropic. Therefore, there exists bν ∈ Q× ν such that

(∗) bν ∈ Q(Uν ) ∩ −Q(Wν ).

Scaling V by −1 if necessary, we can assume that b∞ > 0. Write

bp = cp pe(p) , cp ∈ Z×
p

at every finite place p. We need to invoke an analytic result in number theory:

Theorem 4.3 (Dirichlet’s Theorem on Primes in an Arithmetic Progression) Let m > 0


and a be relatively prime integers. Then there exist infinite many primes p ≡ a mod m.
Let d be the product 2a1 · · · an and let
Y
c= pe(p) .
p|d

31
For every p | d, cp (cp−e(p) )−1 is in Z×
p . Therefore, there is a sufficiently large prime q such
that
q ≡ cp (cp−e(p) )−1 mod p3 ,
for all p | d. If we set b = qc > 0, then b/bp ≡ 1 mod p3 for p | d. This implies that
b/bp ∈ Z×2
p for all p | d. By (∗),

(∗∗) b ∈ Q(Uν ) ∩ −Q(Wν )

for all ν | d and for ν = ∞. If p is a prime which does not divide qd, then b and all the ai
are in Z× p . Therefore, both U ⊥ h−bi and W ⊥ hbi are isotropic at that p. So (∗∗) holds at
all p 6= q. We claim that (∗∗) also holds at q. The ternary space U ⊥ h−bi is isotropic at
all p 6= q. Hence Sp (U ⊥ h−bi) = (−1, −1)Qp at all p 6= q. It follows from Lemma 4.1 that
Y
Sq (U ⊥ h−bi) = (−1, −1)Qν = (−1, −1)Qq .
v6=q

Thus U ⊥ h−bi is isotropic at q also. When n = 4, W ⊥ hbi is isotropic at q by the same


reason. When n > 4, then W is already isotropic at q because the dimension of W is at
least 3 and each ai is in Z×
q (note that q is odd by choice). Hence W ⊥ hbi is isotropic at q.
Since the Strong Hasse Principle is proved for ternary spaces, therefore U ⊥ h−bi is
isotropic. By induction hypothesis, W ⊥ hbi is also isotropic. Therefore, b ∈ Q(U )∩−Q(W )
and thus V = U ⊥ W is isotropic.2

4.2 Weak Hasse Principle


Let a be any number in Q. Since V represents a if and only if V ⊥ h−ai is isotropic,
therefore we have

Proposition 4.4 A nondegenerate quadratic space V over Q represents a ∈ Q if and only


if Vν represents a for every ν ∈ Ω.
Suppose that U is another nondegenerate quadratic space over Q. A necessary condition
for U ∼
= V is that Uν ∼ = Vν for every ν ∈ Ω. The converse also holds and we are going to
prove it by induction on the dimension of U . The case where dim(U ) = 1 can be deduced
from the last proposition. Suppose now that dim(U ) > 1. Let a be an nonzero element in
Q(U ). Then a ∈ Q(Uν ) = Q(Vν ) for all ν ∈ Ω; hence a is represented by V . So we have a
decomposition
U = hai ⊥ U 0 , V = hai ⊥ V 0 .
It follows easily from Witt’s Theorem that Uν0 ∼
= Vν0 for all ν ∈ Ω, and V 0 ∼
= U 0 by the
induction assumption. Thus we have shown the following:

Theorem 4.5 (Weak Hasse Principle) Let U and V be nondegenerate quadratic spaces over
Q. Then U and V are isometric if and only if Uν and Vν are isometric for all places ν of
Q.

32
We have shown that the arithmetic of a quadratic space over Q (or its equivalent, a
quadratic form over Q) is completely determined by the local behavior of the space at every
place of Q. Using the earlier results proven for quadratic spaces over Qp and R, we obtain
the following complete set of isometry class invariants for a nondegenerate quadratic space
V over Q:

(i) the dimension dim(V ),

(ii) the discriminant d(V ),

(iii) the Hasse invariants Sp (V ) at all primes p,

(iv) the positive index Ind+ (V∞ ).

If V ∼
= ha1 , . . . , an i with a1 , . . . , an ∈ Q× , then there exists a finite set of places S, which
contains ∞ and 2, such that each ai is a unit in Zp for any p 6∈ S. Therefore, Sp (V ) = 1
for each p 6∈ S. So in practice one has to check the Hasse invariants for only finitely many
places. This also shows that Vν is isotropic for almost all ν if dim(V ) ≥ 3.
There is a version of Theorem 4.5, which is often referred as the Hasse Principle, which
concerns general representations of quadratic spaces. Let W and V be nondegenerate
quadratic spaces over a field F . We say that W is represented by V if there is an linear
map σ : W → V such that Q(σ(w)) = Q(w) for all w ∈ W . If dim(W ) = 1 and w is a basis
vector for W with Q(w) = a , then W is represented by V if and only if a is represented by
V.

Theorem 4.6 (Hasse Principle) Let W and V be nondegenerate quadratic spaces over Q.
Then W is represented by V if and only if Wν is represented by Vν for all ν ∈ Ω.
Proof. We proceed by an induction on dim(W ). If dim(W ) = 1, this is just Proposition
4.4. Suppose that dim(W ) ≥ 2. Let a be a nonzero rational number represented by W .
Then a is represented by Vν for all ν ∈ Ω. Therefore, a is represented by V . Hence

W = hai ⊥ W 0 , V = hai ⊥ V 0 .

By Witt’s Cancellation Theorem, Wν0 is represented by Vν0 for all ν ∈ Ω. By the induction
assumption, W 0 is represented by V 0 and we are done.2

5 Quadratic Forms over PID


In this section, R denotes a principal ideal domain (PID) and F is the field of fractions of
R. We keep the general assumption that the characteristic of F is not 2. In the subsequent
sections, R is usually taken to be Z or Zp for some prime p. Unless stated otherwise, (V, Q)
is a quadratic space over F .

33
5.1 Lattices on Quadratic Spaces
Definition 5.1 A subset L of V is called an R-lattice (or simply a lattice) in V if L is
a finitely generated R-module. We say that L is a lattice on V if it is a lattice in V and
FL = V .
Let L be a lattice on V . An element a ∈ F is said to be represented by L if there exists
a vector v ∈ L such that Q(v) = a. To see the connection with number theory, let us fix a
basis {v1 , . . . , vn } for L. It is also a basis for V over F . Let f (x1 , . . . , xn ) be the quadratic
form associated with this basis. Then a is represented by L if and only if the diophantine
equation
f (x1 , . . . , xn ) = a
has a solution (x1 , . . . , xn ) ∈ Rn , not only in F n . The fundamental question is still the
representation problem which asks for an effective determination of Q(L), the set of elements
of F represented by L.
Two lattices M and L are isometric, written M ∼ = L, if there exists an isometry σ :
F M −→ F L such that σ(M ) = L. It is clear that isometric lattices represent the same set
of elements of F . Therefore, it is also important to determine, say by means of computable
invariants, whether or not two given lattices are isometric. For any lattice L on V , the
isometry group of L is the set

O(L) = {σ ∈ O(V ) : σ(L) = L}.

Definition 5.2 Let L be a lattice on V . A sublattice N of L is called primitive if N is a


direct summand of L.
So a sublattice N of a lattice L is primitive if and only if L/N is free. A vector x ∈ L
is said to be primitive in L if the rank one lattice Rx is a primitive sublattice of L.

Lemma 5.3 Let L be a lattice on V and x1 , . . . , xn be a basis for L. Suppose that e is a


nonzero vector in L with e = a1 x1 + · · · + an xn . Then e is primitive in L if and only if
gcd(a1 , . . . , an ) = 1.
Proof. If d := gcd(a1 , . . . , an ) is not 1, then d−1 e is a vector in L and so L/Re is not free.
In other words, e is not primitive in L.
Conversely, suppose that e is not primitive in L. Then L/Re is not free. So there exists
x ∈ L, but not in Re, and n 6= 0 such that nx ∈ Re. Note that n is not a unit in R.
Therefore nx = me for some m ∈ R and we may assume that gcd(m, n) = 1. Then
m
x= (a1 x1 + · · · + an xn )
n
which implies that n divides gcd(a1 , . . . , an ). In particular, gcd(a1 , . . . , an ) 6= 1.2

Definition 5.4 A fractional ideal of F is a subset of F of the form (a) = {ar : r ∈ R} for
some a ∈ F × .

34
Lemma 5.5 A subset I of F is a fractional ideal of F if and only if I is a nonzero R-module
and λI ⊆ R for some nonzero λ ∈ F .
Proof. It is clear that every fractional ideal is a nonzero R-module. If I is the fractional
ideal (a), a ∈ F × , then a−1 I = R. Conversely, suppose that I ⊆ F is a nonzero R-module,
and that λI ⊆ R for some λ ∈ F × . Since λI is also a nonzero R-module and λI ⊆ R, it is
just an ideal of R. Therefore, λI = (α) from some nonzero α ∈ R. Then I = (λ−1 α).2
Let L be a lattice on V . For any x ∈ V , define the coefficient of x with respect to L to
be the set
ax = {t ∈ F : tx ∈ L}.
We claim that ax is a fractional ideal of F if x 6= 0. It is clear that ax is an R-module in F .
Let u1 , . . . , un be a basis for L and write x = a1 u1 + · · · + an un where ai ∈ F for all i. There
exists a nonzero r ∈ R such that rai ∈ R for all i. Therefore, r ∈ ax which shows that ax is
nonzero. For each i, write ai = αi βi−1 where αi and βi are in R such that gcd(αi , βi ) = 1.
Let g be the gcd of the αi and ` be the lcm of the βi . Suppose that t = αβ −1 ∈ ax with
α, β ∈ R. So, tαi βi−1 ∈ R for all i. Therefore, we must have

β | αi and βi | α for all i

and hence β | g and ` | α. If we set λ = g`−1 , then λt ∈ R, that is, λax ⊆ R.

Lemma 5.6 Let L be a lattice on V and x be a nonzero vector in V . Then there exists
a ∈ F × such that ax is primitive in L.
Proof. We know that ax = (a) for some a ∈ F × . In particular, ax ∈ L. Let y ∈ L and
suppose that there exists an nonzero t ∈ R such that ty ∈ Rax. Then ty = rax for some
r ∈ R, and hence t−1 rax ∈ L. This means that t−1 ra ∈ ax and therefore t−1 r ∈ R. As a
result, y = t−1 rax ∈ Rax and so L/Rax is torsion free.2
We define the radical of a lattice L in V to be the sublattice

rad(L) = {x ∈ L : B(x, L) = 0}.

We call L nondegenerate if rad(L) = 0. It is easy to see that

F rad(L) = rad(F L)

and
rad(L) = L ∩ rad(F L).
In particular, L is nondegenerate if and only if F L is nondegenerate.

Proposition 5.7 Let L be a lattice in V . Then there exists a nondegenerate lattice K such
that L = K ⊥ rad(L).
Proof. It is clear that if K exists then it must be nondegenerate. It remains to demonstrate
the existence of K. It suffices to show that rad(L) is a primitive sublattice of L, equivalently,

35
L/rad(L) is a torsion free R-module. Suppose that there exists v ∈ L and r ∈ R such that
rv ∈ rad(L). Then B(rv, L) = 0 which implies that r = 0 or B(v, L) = 0, that is,
v ∈ rad(L).2

Remark 5.8 In the decomposition L = K ⊥ rad(L), the lattice K is not unique, but its
isometry class is uniquely determine by that of L. As a consequence, if L0 = K 0 ⊥ rad(L0 )
is another lattice, then L ∼
= L0 if and only if K ∼
= K 0 and rank(rad(L)) = rank(rad(L0 )).
Let L be a nondegenerate lattice and {v1 , . . . , vk } be a basis for L. If A is the symmetric
matrix (B(vi , vj )), we shall write
L∼ = A.
We call such an A a matrix for L. Since F L is a nondegenerate quadratic space, the
determinant of the matrix A is nonzero. The discriminant of L, denoted d(L), is the
canonical image of det(A) in F × /R×2 . Note that d(L) is well-defined because if A0 is the
symmetric matrix associated to another basis for L, then
A0 = T t AT
for some matrix T ∈ GLk (R); thus det(A0 ) = det(A) det(T )2 ∈ det(A)R×2 . Quite often we
also consider d(L) as an element of F × . In this case, d(L) is determined by L up to the
multiplication of an element in R×2 .
Let M be a sublattice of L of the same rank. Let [L/M ] be the product of the invariant
factors of the quotient R-module L/M . It is an ideal of R. By abusing the notation, we
also use [L/M ] to denote one of its generators.

Proposition 5.9 Let L be a nondegenerate lattice and M be a sublattice of L of the same


rank. Then d(M ) = [L/M ]2 d(L).
Proof. There exist a basis {e1 , . . . , ek } for L and nonzero elements a1 , . . . , ak of R such that
{a1 e1 , . . . , ak ek } is a basis for M . Then [L/M ] = a1 · · · ak , and
d(M ) = det(B(ai ei , aj ej ))
= (a1 · · · ak )2 det(B(ei , ej ))
= [L/M ]2 d(L)
2

Remark 5.10 If R = Z or Zp , the group index [L : M ] is a generator of [L/M ]. In either


case, we have d(M ) = [L : M ]2 d(L).
Consider a nondegenerate lattice L. The dual of L is the set
L# = {x ∈ F L : B(x, L) ⊆ R}.
Let {v1 , . . . , vn } be a basis for L. As F L is nondegenerate, there exists a basis {u1 , . . . , un }
for F L such that
(B(vi , uj )) = In .

36
It is easy to see that {u1 , . . . , un } is a basis for L# ; hence L# is a lattice on the space F L
and (L# )# = L. Moreover, L# is isomorphic to Hom(L, R) through the map x 7→ B(x, ).
Let T be the matrix defined by

(v1 , . . . , vn ) = (u1 , . . . , un )T.

Then In = (B(ui , uj ))T and (B(vi , vj )) = T . Therefore,

1 = d(L# ) det(T ), det(T ) = d(L).

Thus we have
d(L# ) = d(L)−1 .
If L has a decomposition L = J ⊥ K, then

L# = J # ⊥ K # .

Finally, for any nonzero a ∈ F ,


(aL)# = a−1 L# .

5.2 Modular Lattices


Consider a lattice L in V . The scale of L, denoted s(L), is the R-module generated by the
subset of B(L, L) of F . If {e1 , . . . , ek } is a basis for L, then
X
s(L) ⊆ B(ei , ej )R
i,j

So, either s(L) is either a fractional ideal or 0. We define the norm n(L) to be the R-module
generated by the subset Q(L) of F . Since Q(L) ⊆ B(L, L), it follows that n(L) is also a
fractional ideal or 0. Now, for all x, y ∈ L we have

2B(x, y) = Q(x + y) − Q(x) − Q(y) ∈ n(L).

Hence
2s(L) ⊆ n(L) ⊆ s(L).
If L = J ⊥ K, then it is easily verified that

s(L) = s(J) + s(K), n(L) = n(J) + n(K).

Let a be a nonzero element of F . Recall that V a denotes the vector space V provided
with a new quadratic map Qa such that Qa (x) = aQ(x) for all x ∈ V . We shall use La to
denote the lattice L when it is regarded as a lattice in V a . It is easy to see that

s(La ) = as(L), n(La ) = an(L), d(La ) = ak d(L)

where k = rank(L).

37
Definition 5.11 A lattice L is called unimodular if s(L) ⊆ R and d(L) is a unit. It is
−1
called (a)-modular if La is unimodular.

Lemma 5.12 Let L be a nondegenerate lattice in V . Then L is (a)-modular if and only if


aL# = L. In particular, L is unimodular if and only if L# = L.
Proof. First suppose that aL# = L. Then

B(L, L) = B(L, aL# ) ⊆ (a),


−1
and so s(L) ⊆ (a). Therefore, s(La ) ⊆ R. Furthermore, if rank(L) = k, then

d(L) = d(aL# ) = a2k d(L# ) = a2k d(L)−1 .


−1
Therefore, d(La ) = a−k d(L) is a unit of R.
−1 −1
Now, assume that L is (a)-modular. Then La is unimodular. Therefore, s(La ) ⊆ R
−1
and d(La ) = a−k d(L) is a unit. For any v ∈ L, we have

a−1 B(v, x) ⊆ R, for all x ∈ L.

So a−1 v ∈ L# or v ∈ aL# ; hence L ⊆ aL# . Now,

d(aL# ) = a2k d(L# ) = a2k d(L)−1 = d(L)2 d(L)−1 = d(L).

Thus L = aL# .2

Corollary 5.13 Suppose that L is an (a)-modular lattice in V . Then

L = {x ∈ F L : B(x, L) ⊆ (a)}.

Proof. This is clear because x ∈ aL# if and only if B(a−1 x, L) ⊆ R, which is equivalent to
B(x, L) ⊆ (a).2

Proposition 5.14 Let L be a nondegenerate lattice in V . If K is an (a)-modular sublattice


of L with s(L) = (a), then K is an orthogonal summand of L.
Proof. Since K is (a)-modular, it is nondegenerate. Therefore, there is an orthogonal
decomposition F L = F K ⊥ W for some nondegenerate subspace W of F L. We claim that

L = K ⊥ (L ∩ W ).

It suffices to show that L = K + (L ∩ W ). Let x ∈ L. Write x = y + z, where y ∈ F K and


z ∈ W . Then
B(y, K) = B(x, K) ⊆ B(L, K) ⊆ s(L) = (a).
But y is in F K; hence y ∈ K.2

38
6 Quadratic Forms over Zp
In this section, (V, Q) is always a quadratic space over Qp for some prime p. Every lattice
considered here is a Zp -lattice.

6.1 Classification in Terms of Jordan Decompositions


Lemma 6.1 (a) If p > 2, every nondegenerate Zp -lattice has an orthogonal basis.
(b) Every nondegenerate Z2 -lattice is an orthogonal sum of modular sublattices of rank 1 or
2.
Proof. Let L be a nondegenerate lattice on V . If s(L) = n(L), then there exists a vector
v ∈ L such that (Q(v)) = s(L) and by Proposition 5.14 it follows that Zp v is an orthogonal
summand of L. This proves (a).
If p = 2 and n(L) = 2s(L), then there exist u, v ∈ L such that Q(v), Q(u) ∈ 2s(L)
and (B(u, v)) = s(L). The binary submodule Z2 u + Z2 v is s(L)-modular and therefore is
an orthogonal summand of L. Thus repeating this argument, L is an orthogonal sum of
modular sublattices of rank 1 or 2.2
Let L be a nondegenerate lattice on V . We can write

L = L1 ⊥ · · · ⊥ Lt ,

where Li is s(Li )-modular and s(L1 ) ) · · · ) s(Lt ). Such an orthogonal decomposition is


called a Jordan decomposition of L, and each Li is a Jordan component. The first component
L1 is called the leading component of that Jordan decomposition. In general, a Jordan
decomposition is not unique.
For any fractional ideal a of Qp , let

La = {x ∈ L : B(x, L) ⊆ a}.

It is clear that La is a lattice in V .

Lemma 6.2 Suppose that M is a b-modular lattice and a 6= b. Then s(M a ) a.


Proof. It suffices to prove the case when a b. By scaling the bilinear form, we may assume
that M is unimodular. Consequently, a Zp . In particular, a = (pa ) with a ≥ 1. Let

M = M1 ⊥ · · · ⊥ M` ⊥ M`+1 ⊥ · · · ⊥ Mk

be an orthogonal decomposition of M such that rank(Mj ) is 1 for all j ≤ ` and 2 for all
j > `. Note that ` < n happens only when p = 2. Let x ∈ M a and write x = x1 + · · · + xk
where xj ∈ Mj for all j.
Suppose that j ≤ `. Let zj be a basis vector for Mj . Then Q(zj ) ∈ Z× p . From
a
B(xj , zj ) ⊆ B(x, Mj ) ⊆ B(x, M ) ⊆ a we see that xj ∈ p Mj .

39
If p = 2 and j > `, let {ej , fj } be a basis for Mj such that Q(ej ), Q(fj ) ∈ 2Z2 and
B(ej , fj ) = γ ∈ Z×
2 . Write xj = αej + βfj where α, β ∈ Z2 . Then

B(xj , ej ) = αQ(ej ) + βγ ∈ a (1)


B(xj , fj ) = βQ(fj ) + αγ ∈ a (2)

If α 6∈ a, then from (2) |α|2 = |βQ(fi )|2 . This implies that |αQ(ej )|2 = |βQ(fj )Q(ej )|2 <
|β|2 , and by (1) we have β ∈ a. But then (2) says that α must be in a which is a contradiction.
Therefore, α ∈ a and from (1) it follows that β is also in a.
Finally, the above shows that M a ⊆ pa M . Therefore, s(M a ) ⊆ p2a s(M ) a.2

Lemma 6.3 If L = L1 ⊥ · · · ⊥ Li ⊥ · · · ⊥ Lt is a Jordan decomposition and s(Li ) = a,


then Li is the leading component of a Jordan decomposition of La .
Proof. For any j, Laj is a subset of La . If x ∈ La and x = x1 + · · · + xt with xj ∈ Lj for all
j, then
B(xj , Lj ) = B(x, Lj ) ⊆ a.
Therefore, xj ∈ Laj and hence La = La1 ⊥ · · · ⊥ Lat . Lemma 6.2 implies that s(Laj ) a for
all j 6= i. Therefore, La = Li ⊥ M where s(M ) s(Li ). This proves the lemma.2

Theorem 6.4 Suppose p > 2. If L = L1 ⊥ · · · ⊥ Lt = K1 ⊥ · · · ⊥ Ks are two Jordan


decompositions of L, then t = s and Li ∼
= Ki for all i.
Proof. Suppose s(Li ) = a. Then La is not empty and hence there must be one and only
one j such that s(Kj ) = a. Similarly, for any Ki , there is a unique L` with s(L` ) = s(Ki ).
Therefore, t = s and s(Li ) = s(Ki ) for 1 ≤ i ≤ t.
Since Li and Ki are the leading components of two Jordan decompositions of the lattice
La , we may assume that i = 1 and s(L1 ) = s(K1 ) = Zp by scaling the quadratic form.
The quotient L/pL becomes a quadratic space over the finite field F = Zp /pZp . While L1
and K1 induce nondegenerate subspaces of L/pL, both L2 ⊥ · · · ⊥ Lt and K2 ⊥ · · · ⊥ Kt
become the radical of L/pL. Thus L1 /pL1 ∼ = K1 /pK1 as quadratic spaces over F, and
therefore rank(L1 ) = rank(K1 ) and d(L1 ) = d(K1 ) in F× /F×2 . Since p > 2, d(L1 ) = d(K1 )
in Z× ×2
p /Zp . Then theorem follows from Corollary 6.20.2

The above theorem does not hold when p = 2. For example, consider the Z2 -lattice L
with Jordan decomposition  
∼ 2 1
L= ⊥ h−2i.
1 2
Suppose that {x, y, z} is a basis which gives the symmetry matrix on the right. Then the
vectors x + z, y + z span a sublattices which is isometric to H. Therefore, L has another
Jordan decomposition  
0 1
L∼= ⊥ h6i.
1 0
Obviously, the leading components of the two Jordan decompositions are not isometric.

40
Theorem 6.5 Let L = L1 ⊥ · · · ⊥ Lt = K1 ⊥ · · · ⊥ Ks be two Jordan decompositions of a
Z2 -lattice L. Then
(a) t = s;
(b) s(Li ) = s(Ki ), rank(Li ) = rank(Ki ) for 1 ≤ i ≤ t;
(c) n(Li ) = n(Ki ) for 1 ≤ i ≤ t;
(d) n(L1 ) = n(L) and s(L1 ) = s(L).

Suppose that L = L1 ⊥ · · · ⊥ Lt is a Jordan decomposition of a Z2 -lattice L. For


1 ≤ i ≤ t, write si = s(Li ) and ni = n(Li ), and set ui = ord2 (n(Li )). The lattice L now has
the following set of invariants:
t, rank(Li ), si , ui , 1 ≤ i ≤ t.
They are called the Jordan invariants of L. For each i, let L(i) = L1 ⊥ · · · ⊥ Li .

Theorem 6.6 Let L and K be lattices on a nondegenerate quadratic space over Q2 . Sup-
pose that L and K have the same Jordan invariants. Let t be the number of Jordan compo-
nents in a Jordan decomposition of L. Then L ∼
= K if and only if the following conditions
hold for 1 ≤ i ≤ t − 1:
(i) d(L(i) )/d(K(i) ) ≡ 1 mod ni ni+1 /s2i ,
(ii) Q2 K(i) ⊥ h2ui i has a subspace which is isometric to Q2 L(i) when ni+1 ⊆ 4ni .

6.2 Maximal Lattices


Definition 6.7 Let a be a nonzero element in Qp . A lattice L on a nondegenerate space
V over Qp is called (a)-maximal if n(L) ⊆ (a) and for any lattice M containing L properly,
n(M ) * (a) holds.
Suppose that L is (a)-maximal but n(L) 6= (a). If n(L) ⊆ (p2 a), then p−1 L properly
contains L and n(p−1 L) ⊆ (a). This contradicts the maximality of L. Therefore, for any
(a)-maximal lattice L, n(L) is either (a) or (pa).

Lemma 6.8 Let L be a lattice on a nondegenerate quadratic space V of dimension n over


Qp . If n(L) ⊆ (a) and ((2a−1 )n d(L)) = Zp or pZp , then L is (a)-maximal.
Proof. Let M be a lattice on V which contains L and n(M ) ⊆ (a). Then
(d(M )) ⊆ s(M )n ⊆ (2−1 n(M ))n ⊆ (a/2)n .
However, d(L) = [M : L]2 d(M ) which implies that
((2a−1 )n d(L)) = ([M : L]2 (2a−1 )n d(M )) ⊆ [M : L]2 Zp .
Therefore, [M : L] = 1 and hence M = L.2

41
Corollary 6.9 If L is (a)-modular with n(L) = (2a), then L is (2a)-maximal. In particular,
if p > 2 then every (a)-modular lattice is (a)-maximal.
Over Z2 , there are unimodular lattices which are not Z2 -maximal. For example, the
Z2 -lattice N ∼
= h1, 3i, with respect to a basis {e, f }, is unimodular. However it is not
Z2 -maximal because it is a sublattice of M = Z2 21 (e + f ) + Z2 21 (e − f ) and n(M ) = Z2 .

Proposition 6.10 Let V be an anisotropic quadratic space over Qp . A lattice L on V is


(a)-maximal if and only if
L = {x ∈ V : Q(x) ∈ (a)}.
In particular, there is one and only one (a)-maximal lattice on V .
Proof. We first claim that the set M = {x ∈ V : Q(x) ∈ (a)} is a lattice. It is clear that M
is closed under multiplication by Zp . Let x, y ∈ M . Assume that 2B(x, y) 6∈ (a). Let t ≥ 1
be an integer such that (2pt B(x, y)) = (a). Then

Q(x)Q(y)B(x, y)−2 ∈ (4p2t )

and, since we also have Q(x + y) = Q(x) + Q(y) + 2B(x, y), 1 − Q(x)Q(y)B(x, y)−2 is a
square of some element c ∈ Z×
p . This implies that

Q(x)Q(y) − B(x, y)2 = −(cB(x, y))2 6= 0,

and, as a result, Qp x ⊕ Qp y is a hyperbolic plane which contradicts the assumption on V .


Next we show that M is finitely generated. Fix a lattice K on V with n(K) ⊆ (a). Then
K must be a subset of M . Let N be a finitely generated Zp -module such that K ⊆ N ⊆ M ,
and let {x1 , . . . , xn } be a basis for N . Since xi ∈ M for all i, the diagonal entries and the
off diagonal entries of the matrix (B(xi , xj )) are in (a) and ( a2 ) respectively. In particular,
det(B(xi , xj )) ∈ ( a2 )n and hence
 a −n
Zp ⊇ ([N : K]2 ) = (d(K) det(B(xi , xj ))−1 ) ⊇ (d(K)) .
2
This implies that [N : K] is bounded, and M must be finitely generated.
It is clear by now that M is an (a)-maximal lattice. If L is an (a)-maximal lattice on
V , then L ⊆ M . But n(M ) is clearly contained in (a). Therefore, L = M . 2

Lemma 6.11 Let L be a lattice on the hyperbolic plane. The following three conditions are
equivalent:
 
0 c
(i) L ∼
= ;
c 0

(ii) L is (c)-modular and n(L) ⊆ (2c);

(iii) L is (2c)-maximal.

42
Proof. (i) =⇒ (ii) is obvious, and (ii) =⇒ (iii) follows from Corollary 6.9 by putting a = 2c.
For (iii) =⇒ (i), let {x, y} be a basis for L such that Q(x) = 0. Write B(x, y) = a and
Q(y) = b for some a, b ∈ Qp . Then

a ∈ s(L) ⊆ 2−1 n(L) ⊆ (c), and b ∈ n(L).

We claim that (a) = (c). If not, then a ∈ (cp) and for any α, β ∈ Zp ,

Q(αp−1 x + βy) = 2αβp−1 a + β 2 b ∈ (2c).

Therefore, the lattice Zp p−1 x + Zp y contains L properly and its norm is contained in (2c).
This contradicts the maximality of L.
Now, write c = a for some  ∈ Z× −1
p . Then {x, −b(2a) x + y} is a basis for L whose
 
0 c
associated symmetric matrix is .2
c 0

Theorem 6.12 Let V be a nondegenerate isotropic quadratic space over Qp , and let L and
K be (a)-maximal and (b)-maximal, respectively, lattices on V . Then there exists a splitting
V = U ⊥ W in which U is a hyperbolic plane and

L = (L ∩ U ) ⊥ (L ∩ W ), K = (K ∩ U ) ⊥ (K ∩ W ).

Proof. For any isotropic vector x of V , let ax be the coefficient of x with respect to L, that
is,
ax = {t ∈ Qp : tx ∈ L}.
It has been shown that ax is a fractional ideal. We let bx be the coefficient of x with respect
to K. Now αK ⊆ L for some nonzero α in Qp . Therefore, rx := bx /ax is inside (α−1 ); hence
we can pick an isotropic vector x in V for which rx is maximal among all the coefficients of
the isotropic vectors in V . For our convenience, we let ` and k be generators of ax and bx
respectively. In particular, `x and kx are basis vectors of L and K respectively.
Note that
B(`x, L) ⊆ s(L) ⊆ 2−1 n(L) ⊆ (a/2).
We claim that B(`x, L) = (a/2). If not, then B(`x, L) ⊆ (pa/2). So, for any v ∈ L,

Q(v + p−1 `x) = Q(v) + 2p−1 B(v, `x) ⊆ (a).

This implies that the norm of the lattice L + Zp p−1 `x is contained in (a). But L $ L +
Zp p−1 `x and this contradicts the maximality of L.
Now, let w ∈ L such that B(`x, w) = a/2. Put y = w − a−1 Q(w)`x ∈ L. Then
 
∼ 0 a/2
Zp `x + Zp y = .
a/2 0

43
So, Zp `x + Zp y is an (a/2)-modular lattice on a hyperbolic plane. Since s(L) ⊆ (a/2),
Zp `x + Zp y must be an orthogonal summand of L.
Let k 0 be a generator of by . Then Zp kx + Zp k 0 y ⊆ K and so kk 0 /` ∈ (b/a), that is,

rx ry ⊆ (b/a).

As is done in the last paragraph, there exists isotropic vector z ∈ K such that Zp z + Zp k 0 y
splits K and B(z, k 0 y) = b/2. Let `0 be a generator of az . Then Zp `0 z + Zp y ⊆ L and so
`0 b/k 0 ∈ (a), that is,
(b/a) ⊆ rz ry .
But rx is chosen to be maximal. Therefore,

rx ry = (b/a)

which implies that the scale of Zp kx + Zp k 0 y is precisely (b/2). Hence Zp kx + Zp k 0 y splits


K. Thus U = Qp x + Qp y gives the desired splitting of V .2

Corollary 6.13 Let L be an (a)-maximal lattice on a nondegenerate quadratic space V


over Qp . Then  m
∼ 0 a/2
L= ⊥ L0
a/2 0
where m is the Witt index of V and L0 is anisotropic. Moreover, the isometry class of L0
is determined by that of V . As a result, all (a)-maximal lattices on V are isometric.
Proof. By Theorem 6.12, there is a splitting V = U1 ⊥ · · · ⊥ Um ⊥ V0 , in which U1 , . . . , Um
are hyperbolic planes and V0 is anisotropic, such that

L = (L ∩ U1 ) ⊥ · · · ⊥ (L ∩ Um ) ⊥ (L ∩ V0 ).
 
0 a/2
It is clear that each L ∩ Ui is isometric to a/2 0 . The isometry class of V0 is uniquely
determined by that of V and L∩V0 is the only (a)-maximal lattice on V0 . Thus the isometry
class of L ∩ V0 is uniquely determined by that of V .2

Corollary 6.14 Let L and K be (a)-maximal lattices on a nondegenerate quadratic space


over Qp . Then [L : L ∩ K] = [K : L ∩ K].
Proof. By Theorem 6.12, there are hyperbolic pairs xi , yi , 1 ≤ i ≤ m, and anisotropic lattice
N such that
m m
! !
X X
−1
L= Zp xi + Zp yi ⊥ N, K = Zp ai xi + Zp ai yi ⊥ N,
i=1 i=1
Pm
where a1 , . . . , am are nonzero elements in Zp . Then L ∩ K = ( i=1 Zp ai xi + Zp yi ) ⊥ N
and the corollary now follows immediately.2

44
Lemma 6.15 Let N be a lattice in a nondegenerate quadratic space V over Qp . If n(N ) ⊆
(a), then there exists an (a)-maximal lattice on V which contains N .
Proof. Let {v1 , . . . , vm } be a basis for N . Extend it to a basis {v1 , . . . , vn } for V (m ≤ n).
Let
X n
M =N⊕ Zp pt vi
i=m+1

for some integer t. If t is large enough, then n(M ) ⊆ (a). Let L be a lattice on V which
contains M and n(L) ⊆ (a). Then d(M ) = [L : M ]2 d(L) and d(L) ∈ s(L)n ⊆ (2−1 a)n .
Since [L : M ] is an integer, it must be bounded and hence there exists a lattice K which is
maximal with respect to inclusion such that K ⊇ M and n(K) ⊆ (a). Such K must be an
(a)-maximal lattice on V .2

Theorem 6.16 Let L be an (a)-maximal lattice on a nondegenerate quadratic space V over


Qp . For any b ∈ Q×
p , b is represented by L if b is represented by V and b ∈ (a).

Proof. Let v be a vector in V such that Q(v) = b. Then the norm of the rank 1 lattice
Zp v is (b) ⊆ (a). Using the previous lemma we find an (a)-maximal lattice K on V which
contains Zp v. By Theorem 6.12, K is isometric to L and therefore there exists w ∈ L with
Q(w) = b.2

Corollary 6.17 Let L be an (a)-maximal lattice on a nondegenerate quadratic space V of


dimension ≥ 4 over Qp . Then L represents every element in (a).
Proof. Since dim(V ) ≥ 4, V is universal and hence it represents all elements in (a). By
Theorem 6.16, L represents every element in (a).2

6.3 Modular Lattices over Zp , p > 2


In this subsection we shall present a solution to the classification problem of modular lattices
over Zp when p > 2. Together with Theorem 6.4 they provide a solution to the classification
of arbitrary Zp -lattices when p is odd. By Corollary 6.9, all (a)-modular lattices are (a)-
maximal. Therefore, if L is an (a)-modular lattice on a nondegenerate quadratic space over
Qp , then  m
∼ 0 a
L= ⊥ L0
a 0
where m is the Witt index of the ambient space and L0 is anisotropic.
 
0 a
Corollary 6.18 If p > 2 and L is an (a)-modular lattice of rank ≥ 3, then is an
a 0
orthogonal summand of L. In particular, L represents all elements in (a).
Proof. Under the assumption on the rank of L, the ambient quadratic space is isotropic.2

45
Lemma 6.19 If p > 2 and L is a unimodular lattice of rank 2, then L represents every
unit of Zp .
Proof. Let ∆ be a nonsquare unit of Zp . By Lemma 6.1, L has an orthogonal basis; hence
L is isometric to one of the following binary lattices:

h1, 1i, h∆, ∆i, h1, ∆i.

It is clear that the last lattice represents all units in Zp . The first two are isometric because
they are Zp -maximal lattices on isometric quadratic spaces over Qp ; see Theorem 6.12.
Therefore, it is enough to consider the case where L ∼ = h1, 1i. Then L is a Zp -maximal
lattice on the space h1, 1i, and this space represents all units of Zp . By Theorem 6.16, L
represents all units of Zp .2

Corollary 6.20 If p > 2, the isometry class of an (a)-modular lattice L is determined by


rank(L) and d(L). In particular, if L is unimodular, then

L∼
= h1i ⊥ · · · ⊥ h1i ⊥ hd(L)i.

Proof. We may assume that L is unimodular. By Corollary 6.18 and Lemma 6.19, L
represents 1 if rank(L) ≥ 2. Therefore, L ∼
= h1i ⊥ L0 for some unimodular lattice L0 . An
induction on the rank of L will complete the proof. 2

6.4 Modular Lattices over Z2


If L is an (a)-modular Z2 -lattice, then s(L) = (a) and n(L) is either (a) or (2a). We say
that L is improper if n(L) = (2a); proper otherwise. If L is an improper Z2 -modular lattice,
then n(L) = (2) and we call L even unimodular.
  On the hyperbolic plane H, there is a
0 1
unique unimodular lattice isometric to which we also denote by H.
1 0

Lemma 6.21 Let L be a binary even unimodular Z2 -lattice on a space V .


(i) If L is isotropic, then L ∼
= H, d(L) = −1, S(V ) = −1 and Q(L) = (2).
 
∼ 2 1
(ii) If L is anisotropic, then L = , d(L) = 3, S(V ) = 1 and Q(L) = {x ∈ Z2 :
1 2
ord2 (x) ≡ 1 mod 2} ∪ {0}.

Proof. The lattice L is (2)-maximal by virtue of Corollary 6.9. The case where L is isotropic
can be deduced from Lemma 6.11. Suppose that L is anisotropic. Let x be a basis vector
of L. Since L is unimodular, there exists y ∈ L = L# such that B(x, y) = 1. Then {x, y}
must be a basis for L and the corresponding symmetric matrix for L is
 
2a 1
.
1 2b

46
If a or b is divisible by 2, then d(L) = 4ab − 1 ≡ −1 mod 8. So L is isotropic which is a
contradiction. As a result, both a and b are units and d(L) = 3. Moreover, V ∼ = h2a, 6ai
and a direct computation shows that S(V ) = 1. By the classification of quadratic spaces
over Q2 , we see that  
∼ 2 1
V = ,
1 2
   
2 1 ∼ 2 1
which has a unique (2)-maximal lattice isometric to . Therefore, L = .
1 2 1 2
For any x ∈ Q× 2 , x ∈ Q(L) if and only if x ∈ Q(V ) and x ∈ (2) by Theorem 6.16.
However, x ∈ Q(V ) if and only if V ⊥ h−xi is isotropic, which is equivalent to requiring
that its Hasse invariant is equal to (−1, −1) = −1. A direct computation shows that the
Hasse invariant of V ⊥ h−xi ∼ = h2, 6, −xi is (5, −x) = (−1)ord2 (x) . This completes the proof
of the lemma.2
The last lemma says that there is only one isometry class of anisotropic binary even
unimodular Z2 -lattice. Any one of such lattices is denoted by A.

Lemma 6.22 Suppose that L ∼ = h1 , 2 , 3 i where i ∈ Z× ∼


2 for all i. Then L = P ⊥ hi
×
for some  ∈ Z2 and an even unimodular lattice P. Moreover P = H if and only if L is
isotropic.
Proof. Let {v1 , v2 , v3 } be an orthogonal basis for L in which Q(vi ) = i for all i. Let P be
the binary lattice spanned by {v1 + v2 , v2 + v3 }. Then
 
1 + 2 2
P∼= .
2 2 + 3

Since the sum of any two units of Z2 is divisible by 2, P is a binary even unimodular lattice
and hence P = H or A. Moreover, L ∼ = P ⊥ hi for some  ∈ Z× 2 . Lemma 6.21 implies that
L is isotropic if and only if P = H.2

Theorem 6.23 Let L be a unimodular Z2 -lattice. If L is even, then L ∼ = H or H ⊥ A,


where H is an orthogonal sum of some copies of H. If n(L) = Z2 , then L has an orthogonal
basis.
Proof. Suppose that L is even unimodular. Then L is (2)-maximal, and by Theorem 6.12
L is an orthogonal sum of binary even unimodular lattices. To complete the proof of this
case, it suffices to show that
H⊥H∼ = A ⊥ A.
Both lattices are (2)-maximal, and hence we only need to show that their ambient spaces
are isometric. This can be done by comparing the Hasse invariants and the discriminants.
Suppose n(L) = Z2 . Then L ∼ = hi ⊥ L0 for some  ∈ Z×
2 and a unimodular lattice L .
0

Therefore it suffices to show that for any binary even unimodular lattice P, hi ⊥ P has an
orthogonal basis. This is done by Lemma 6.22.2

47
Corollary 6.24 Let L be a unimodular Z2 -lattice with rank ≥ 5. Then H is an orthogonal
summand of L.
Proof. If L is even unimodular, then Theorem 6.23 implies the assertion. So we may suppose
that
L∼= h1 , . . . , n i
where i ∈ Z× 2 for all i. If h1 , 2 , 3 i is isotropic, then we are done by Lemma 6.22. Otherwise

h1 , 2 , 3 i = A ⊥ hαi for some α ∈ Z× ∼
2 . We may then assume that hα, 4 , 5 i = A ⊥ hβi for
× ∼
some β ∈ Z2 . But then A ⊥ A = H ⊥ H is an orthogonal summand of L.2

7 Quadratic Forms over Z


Unless stated otherwise, (V, Q) is always a quadratic space over Q and all lattices considered
in this section are Z-lattices. Since ±1 are the only units in Z, the discriminant of a
nondegenerate lattice in V is well-defined as a rational number.

7.1 Preliminaries
A lattice L in V is called integral if B(L, L) ⊆ Z. For an integral lattice L, L# contains L
and d(L) = [L# : L]2 d(L# ). Since d(L# ) = d(L)−1 , we have
[L# : L] = |d(L)|.

Lemma 7.1 Let L be an integral lattice on a nondegenerate quadratic space V over Q.


There are only finitely many integral lattices on V which contain L.
Proof. If M is an integral lattice on V which contains L, then L ⊆ M ⊆ L# . Since L# /L
is a finite abelian group, there are only finitely many such M .2

Lemma 7.2 Let L be an integral lattice on a nondegenerate quadratic space V over Q, and
N be a nondegenerate sublattice of L. Then d(N ⊥ ) divides d(L)d(N ).
Proof. Since L is integral, therefore for any x ∈ L the linear map z 7→ B(x, z) is a homo-
morphism from N to Z. But N is nondegenerate; hence there exists a unique y ∈ N # such
that B(x, z) = B(y, z) for all z ∈ N . Define a map φ : L −→ N # by φ(x) = y. Since
ker(φ) = N ⊥ and φ(x) = x for all x ∈ N , φ−1 (N ) = N ⊥ N ⊥ and this implies
[L : N ⊥ N ⊥ ] = [φ(L) : N ].
The right hand side of the above equation certainly divides [N # : N ] = |d(N )|. But
d(N ) d(N ⊥ ) = d(L)[L : N ⊥ N ⊥ ]2 . Therefore, d(N ⊥ ) divides d(L)d(N ).2

Theorem 7.3 (Hermite) Let L be a lattice on a nondegenerate quadratic space V over Q.


Let m(L) = min{|Q(x)| : x ∈ L \ {0}}. Then
  n−1
4 2 1
m(L) ≤ |d(L)| n .
3

48
Proof. We may assume that L is anisotropic. The proof will be by induction on n = dim(V ).
When n = 1, the theorem is obviously true. Assume that the theorem holds for lattices
on quadratic spaces of dimension ≤ n − 1. Let e1 ∈ L such that |Q(e1 )| = m(L). Then e1
must be primitive in L. If not, there would be an x ∈ L and nonzero integer a such that
ax = e1 . Then
|Q(x)| = a−2 |Q(e1 )| < |Q(e1 )|
which is impossible.
Now, extend {e1 } to a basis {e1 , e2 , . . . , en } for L. Let Φ : V −→ V be the orthogonal
projection of V onto (Qe1 )⊥ . Then

B(x, e1 )
Φ(x) = x − e1 , x ∈ V,
Q(e1 )
Pn
and L0 := Φ(L) = i=2 ZΦ(ei ). Let f1 = e1 and fi = Φ(ei ) for i ≥ 2. Then

|d(L)| = | det(B(ei , ej ))| = | det(B(fi , fj ))| = m(L)|d(L0 )|.

For any x0 ∈ L0 , there exist t ∈ [− 12 , 12 ] and x ∈ L such that x0 = x+te1 . Pick x0 ∈ L0 so that
|Q(x0 )| = m(L0 ). Then Q(x) = Q(x0 ) +t2 Q(e1 ) and hence m(L) ≤ |Q(x)| ≤ m(L0 ) + 41 m(L).
That is,
3
m(L) ≤ m(L0 )
4
  n−2
4 2 1
≤ |d(L0 )| n−1 ,
3

where the last inequality is from the induction assumption. Using |d(L)| = m(L)|d(L0 )| we
obtain the desired inequality for m(L).2

Theorem 7.4 For any given positive integer n and nonzero rational number d, there are
only finitely many non-isometric lattices of rank n and discriminant d.
Proof. The theorem is obvious if n = 1. Let us assume that the theorem holds for all lattices
of rank ≤ n − 1. Let L be a lattice of rank n and discriminant d. By scaling, we could
assume that L is integral. Pick e1 ∈ L so that |Q(e1 )| = m(L). Construct a sublattice N
of L as follows:
• If m(L) 6= 0, put N = Ze1 .

• Suppose that m(L) = 0. Since L is nondegenerate, the set B(e1 , L) is a nonzero ideal
of Z with a positive generator a. Then a2 | d; so there are only finitely many choices
for a. Let e2 ∈ L such that B(e1 , e2 ) = a. Note that e1 must be linearly independent
from e2 . Since
Q(te1 + e2 ) = Q(e2 ) + 2tB(e1 , e2 )

49
and
B(e1 , te1 + e2 ) = a,
we can choose e2 so that |Q(e2 )| ≤ a. We set N = Ze1 + Ze2 .

In both cases, there are only finitely many possibilities for the isometry class of N . By
Lemma 7.2, d(N ⊥ ) divides d(N ) · d. By the induction assumption, there are only finitely
many non-isometric lattices with rank and discriminant the same as those for N ⊥ . Since L
contains N ⊥ N ⊥ , the theorem is now a consequence of Lemma 7.1.2

7.2 Genus and Class


Let L be a lattice on V . If p is a prime, the p-adic completion of L at p is the Zp -lattice
Lp = Zp ⊗ L. Let K be another lattice on V . By the Invariant Factor Theorem, there exist
a basis x1 , . . . , xn for L and nonzero rational numbers r1 , . . . , rn such that r1 x1 , . . . , rn xn
is a basis for K. Consequently, Lp = Kp for almost all primes p. Since the only rational
numbers that are p-adic units for every prime p are ±1, therefore L = K if and only if
Lp = Kp for all primes p.
Let σ : V → V be a linear map. For every prime p, σ can be extended uniquely to a
linear map from Vp to Vp itself which we again denote by σ. It is not hard to check that
σ(Lp ) = σ(L)p .
Suppose that M is a lattice in another quadratic space over Q. A necessary condition
for M ∼= L is that Mp ∼ = Lp for all primes p and RM ∼ = V∞ . This necessary condition
implies that Qν M ∼ = νV for all places ν of Q. It follows from the Weak Hasse Principle that
there exists an isometry QM −→ V which sends M into V . Therefore, when studying the
problem of deciding when two lattices are isometric, it is enough to consider lattices on the
same quadratic space.
The class of L, written cls(L), is the set of all lattices M on V such that M ∼ = L. The
genus of L, denoted by gen(L), is the set of all lattices M on V such that Mp ∼ = Lp for all

primes p. Note that RM = V∞ is automatic since M is also on V . It is obvious that if
M ∈ gen(L), then d(M ) = d(L) and rank(M ) = rank(L).

Theorem 7.5 The genus of a lattice is partitioned into finitely many classes.
Proof. This follows from Theorem 7.4.2

Definition 7.6 The number of classes in gen(L) is called the class number of L.

Example 7.7 A lattice L is said to be positive definite if Q(x) ≥ 0 for all x ∈ L and
Q(x) = 0 if and only if x = 0. Now, suppose that L is a positive definite unimodular
lattice of rank ≤ 5. Then Hermite’s theorem implies that L must represent 1. Therefore,
L∼ = h1, . . . , 1i. In particular, the class number of L is 1. The same conclusion holds for
positive definite unimodular lattices of rank ≤ 7, but the proof requires a much involved
argument.

50
Theorem 7.8 Let L be a lattice on a quadratic space V over Q. For each prime p, let
M (p) be a Zp -lattice on Vp . Then there is a lattice M on V such that Mp = M (p) for all p
if and only if Lp = M (p) for almost all p.
Proof. We first assume that there exists M on V such that Mp = M (p) for all p. Since we
have seen that Mp = Lp for almost all p, therefore Lp = M (p) for almost all p.
Conversely, suppose that Lp = M (p) for almost all p. It suffices to show that given a
single prime q, there is a lattice K on V such that
(
Lp if p 6= q;
Kp =
M (q) if p = q.

The theorem will follow by successive applications of this special case.


We first prove the following contention: given a prime q, there is a basis {y1 , . . . , yn } for
V such that
M (q) = Zq y1 + · · · + Zq yn .
To prove this we fix a basis {x1 , . . . , xn } for V . By multiplying a suitable power of q to
each xi we may assume that all xi are in M (q). Suppose

M (q) = Zq z1 + · · · + Zq zn .

Write each zj as a linear combination of the xi , say

zj = α1j x1 + · · · + αnj xn , αij ∈ Qq .

Since Q is dense in Qq , we can find, for each i and j, a rational number aij ∈ Q such that
|aij − αij |q is as small as we wish. If the approximations are good enough we will obtain,
by virtue of continuity of multiplication and addition in Qq ,

| det(aij ) − det(αij )|q < | det(αij )|q .

Hence | det(aij )|q = | det(αij )|q . Put yj = a1j x1 + · · · + anj xn for 1 ≤ j ≤ n. Then

yj − zj ∈ Zq x1 + · · · + Zq xn ⊆ M (q)

by our choice the aij , hence

Zq y1 + · · · + Zq yn ⊆ M (q).
P
If we write yj = i γij zi with all γij ∈ Zq we have

(γij ) = (αij )−1 (aij ),

hence det(γij ) is a unit in Zq . As a result,

Zq y1 + · · · + Zq yn = M (q).

51
Now, we can find a lattice J on V such that Jq = M (q). By the Invariant Factor
Theorem of finitely generated modules over PID, there exists a basis e1 , . . . , en for L and
rational numbers r1 , . . . , rn such that r1 e1 , . . . , rn en is a basis for J. For any i, let ti be the
q-part of ri . Then K = Zt1 e1 + · · · + Ztn en has the desired property.2
A rational number a is said to be represented by the genus of L if V∞ represents a and
Lp represents a for every p. This is a necessary condition for a to be represented by L.

Theorem 7.9 If a is represented by the genus of L, then a is represented by some lattice


in gen(L).
Proof. We may assume that a 6= 0. From the Strong Hasse Principle it follows that a is
represented by V . Let v ∈ V such that Q(v) = a. Then for almost all p, v ∈ Lp . If v 6∈ Lp
for a prime p, then Witt’s extension theorem implies that there exists an isometry σp of Vp
such that σp (v) ∈ Lp . We define a lattice M on V by
(
Lp if v ∈ Lp ;
Mp = −1
σp (Lp ) otherwise.

This is well-defined by Theorem 7.8. Then M ∈ gen(L) and v ∈ Mp for all p; so v ∈ M .2

Corollary 7.10 If L has class number 1, then L represents all rational numbers that are
represented by its genus.

Example 7.11 Consider the Z-lattice L = h1, 11i. The equation


 2  2
8 1
3= + 11
5 5

shows that V∞ represents 3. Since 5 ∈ Z× p for all p 6= 5, Lp represents 3 for those p; but
L5 ∼
= H also represents 3. Therefore 3 is represented by gen(L). But it is obvious that 3 is
not represented by L. Incidentally this shows that there are at least two classes in gen(L).
One can show that  
3 1
M=
1 4
is a lattice in gen(L) which represents 3, and that M ∼
6 L.
=

7.3 Sum of Squares


For any integer n ≥ 1, let In be the lattice corresponding to the sum of n squares, that is,

In = h1, . . . , 1i.

When n ≤ 7, In has class number 1. Therefore, the set Q(In ) is precisely the set of
positive integers that are represented by the genus of In .

52
Theorem 7.12 (Euler) A positive integer m is a sum of two integer squares if and only if
ordp (m) is even for all primes p ≡ 3 mod 4.
Proof. Let p be a prime. We need to determine the set of p-adic integers that is represented
by the lattice h1, 1i over Zp .

(a) p ≡ 1 mod 4: In this case, −1 is a square in Zp . Therefore, h1, 1i ∼


= H over Zp and
hence h1, 1i represents every p-adic integer.

(b) p ≡ 3 mod 4: Since −1 is a nonsquare in Zp , h1, 1i is a Zp -maximal lattice on an


anisotropic quadratic space over Qp . It follows from Theorem 6.16 that a nonzero
m ∈ Zp is represented by h1, 1i if and only if the space h1, 1, −mi is isotropic, and the
latter is the same as saying that ordp (m) is even.

(c) p = 2: A direct computation shows that a nonzero 2-adic integer m is represented by


h1, 1i if and only if m = 2α b, where b ≡ 1 mod 4.

The theorem is now a consequence of (a), (b) and (c) together.2

Theorem 7.13 (Legendre) A positive integer m is a sum of three integer squares if and
only if m is not of the form m = 4a (8b + 7), a, b ∈ Z,
Proof. Since h1, 1, 1i is unimodular over Zp for all odd primes p, by Corollary 6.18 it repre-
sents all elements of Zp . Therefore, it suffices to show that over Z2 , h1, 1, 1i represents all
2-adic integers not of them 4a (8b + 7).
Over Q2 , it is direct to show that the space h1, 1, 1, −mi is anisotropic if and only
if m ∈ −Q×2 2 . Therefore, the lattice h1, 1, 1i does not represent any integer of the form
a
4 (8b + 7). But it is clear that h1, 1, 1i represents 1, 3 and 5. Furthermore, since

h1, 1, 1i ∼
= A ⊥ h3i

by Lemma 6.22, h1, 1, 1i represents all 2-adic integers m with ord2 (m) ≡ 1 mod 2; see
Lemma 6.21. Consequently, h1, 1, 1i represents all 2-adic integers of the form not of them
4a (8b + 7).2

Theorem 7.14 (Lagrange) Every positive integer is a sum of four integer squares.
Proof. By Corollary 6.18 I4 represents all elements of Zp when p > 2. Over Z2 , it can be
checked directly that I4 represents all 2-adic integers. Therefore, the genus of I4 represents
all positive integers.2

7.4 Spinor Norms


In this subsection, let (V, Q) be a nondegenerate quadratic space over an arbitrary field F
whose characteristic is not 2. If W is a nondegenerate subspace of V , then V = W ⊥ W ⊥ .
Thus every isometry σ of W can be extended to an isometry of V by defining, σ(x) = x for
all x ∈ W ⊥ . So we can identify O(W ) as a subgroup of O(V ).

53
Proposition 7.15 Every isometry in O(V ) is a product of symmetries.
Proof. The proof is by induction on n = dim(V ). The case n = 1 is trivial, so we assume
that n > 1. Take σ ∈ O(V ) and any v ∈ V such that Q(v) 6= 0. Then

Q(v − σ(v)) + Q(v + σ(v)) = 4Q(v) 6= 0.

Hence either Q(v − σ(v)) or Q(v + σ(v)) is nonzero. In the first case we have

τv−σ(v) (v) = σ(v),

while in the second case


τv+σ(v) τv (v) = σ(v).
So in either case there exists ρ ∈ O(V ) which is a product of one or two symmetries such
that ρ(v) = σ(v). Let W be the orthogonal complement of v. Then ρ−1 σ is in O(W )
and induction hypothesis implies that ρ−1 σ is a product of symmetries τv1 · · · τvk in O(W ).
Then
σ = ρτv1 · · · τvn
which is a product of symmetries in O(V ).2

Remark 7.16 The proof of the above proposition can be modified to show that if p > 2
and L is a modular Zp -lattice, then O(L) is also generated by symmetries. For, by scaling
the quadratic form suitably we may assume that L is unimodular. Then L contains a vector
v with Q(v) ∈ Z× ×
p . Hence either Q(v + σ(v)) or Q(v − σ(v)) is in Zp . The rest of the proof
will go over in nearly verbatim fashion.
If σ ∈ O(V ) and σ = τv1 · · · τvk for some anisotropic vectors v1 , . . . , vk ∈ V , we define
the spinor norm of σ to be

θ(σ) = Q(v1 ) · · · Q(vk )F ×2 ∈ F × /F ×2 .

Theorem 7.17 θ : O(V ) −→ F × /F ×2 is a well-defined group homomorphism.


We are not going to be too rigid in our use of the θ notation, although it will always be
clear from the context just what we have in mind. The equality θ(σ) = a with a ∈ F × will
often appear; this really means θ(σ) is the canonical image of a in F × /F ×2 . Occasionally
we regard θ(σ) as the full coset aF ×2 taken as a subset of F × . More generally, if X is a
subset of O(V ), θ(X) is the image of X in F × /F ×2 under θ; but we shall also regard θ(X)
as the union of the cosets [
θ(X) = θ(σ)F ×2 .
σ∈X

Lemma 7.18 If σ ∈ O(V ), then det(σ) = ±1.

54
Proof. Let B = {v1 , . . . , vn } be a basis for V and A be the symmetric matrix (B(vi , vj )).
For 1 ≤ j ≤ n, write
Xn
σ(vj ) = tij vi .
i=1

Then the matrix T = (tij ) is the matrix representation of σ with respect to the basis B. If
we identify V with the column space F n , then for any x, y ∈ V ,

xt Ay = B(x, y) = B(σ(x), σ(y)) = xt T t AT y,

hence A = T t AT . Therefore det(T ) = ±1.2


The special orthogonal group of V , denoted by O+ (V ), is the set

O+ (V ) = {σ ∈ O(V ) : det(σ) = 1}.

The kernel of the homomorphism θ : O+ (V ) −→ F × /F ×2 is written as O0 (V ). It is clear that


O0 (V ) is a normal subgroup of O(V ) (not only O+ (V )). Note that the groups O0 (V ), O+ (V ),
and O(V ) are unchanged upon scaling the quadratic map on V by any nonzero element of
F.

7.5 Spinor Genus


Suppose that V is a nondegenerate quadratic space over Q. Let L be a lattice on V . The
proper class of L is the set cls+ (L) of all lattices K on V such that K = σ(L) for some
σ ∈ O+ (V ). It is clear that
cls+ (L) ⊆ cls(L).
Each class in gen(L) breaks into at most two proper classes, and cls(L) = cls+ (L) if and
only if O(L) contains an isometry of determinant −1.
We can define the proper genus of L to be the set gen+ (L) of all lattices K on V such
that Kp = σp (Lp ) for some σp ∈ O+ (Vp ) at each p. However, this turns out to be an obsolete
definition.

Lemma 7.19 gen+ (L) = gen(L).


Proof. It suffices to show that at each p, O(Lp ) contains a symmetry. If Lp has an orthogonal
basis, say {x1 , . . . , xn }, then it is easy to verify that τxi is a symmetry in O(Lp ). Therefore
we are left with the case when p = 2 and L2 is an orthogonal sum of binary improper
modular lattices. Upon scaling the quadratic form on L suitably, we may assume that
L2 = J ⊥ K with J ∼ = H or A. Take any vector v ∈ J with Q(v) = 2. The symmetry τv is
readily an element in O(L2 ).2

Lemma 7.20 If dim(V ) ≥ 3, then for any prime p, θ(O+ (Vp )) = Q×


p.

Proof. If dim(V ) ≥ 4, then Vp is universal. Hence there is a symmetry with any prescribed
spinor norm and the proposition follows immediately.

55
Suppose that dim(V ) = 3. It is known that Vp represents a ∈ Q× p if the discriminant of
Vp ⊥ h−ai is not 1. So if d(Vp ) is a unit, Vp represents all prime elements of Qp and at least
one unit in Zp . Now every element of Q× p is a product of exactly two such elements times
× + ×
a square in Qp ; hence θ(O (Vp )) = Qp . A similar argument applies when d(Vp ) is a prime
element.2
A lattice K is in the proper spinor genus of L if there exist φ ∈ O+ (V ) and σp ∈ O0 (Vp )
at each prime p such that
φ(Kp ) = σp (Lp ) for all p.
The proper spinor genus of L is written as spn+ (L).

Theorem 7.21 If a is represented by gen(L) and rank(L) ≥ 4, then a is represented by


some lattice in spn+ (L).
Proof. We may assume that a 6= 0 and, by virtue of the Strong Hasse Principle, that V
represents a. Let v be a vector in V such that Q(v) = a. Let S be the set of primes for
which v 6∈ Lp . Note that S is a finite set. For p ∈ S, since Lp represents a, there exists
an isometry σp ∈ O(Vp ) such that σp (v) ∈ Lp . Since O(Lp ) contains a symmetry, we may
assume that σ is in O+ (Vp ). Let W be the orthogonal complement of v in V . By Lemma
7.20 there exists ρp ∈ O+ (Wp ) such that θ(ρp ) = θ(σp ) for each p ∈ S. Define a lattice K
on V by 
Lp if p 6∈ S;
Kp =
ρp σp−1 (Lp ) if p ∈ S.
Then K ∈ spn+ (L) and v ∈ K.2

8 Strong Approximation
8.1 Norms on orthogonal groups
In this subsection, let p be a prime and V be a nondegenerate quadratic space over Qp . We
shall define a norm k kp , or simply k k if no confusion arises, on O(V ). Let {x1 , . . . , xn }
be a basis for V . The norm k k which we are about to define is with respect to the basis
{x1 , . . . , xn } unless stated otherwise. Let M be the Zp -lattice spanned by this basis.
We first define the norm on V . For any x ∈ V , express it as

x = α1 x1 + · · · + αn xn , αi ∈ Qp ,

and define the norm of x by


kxk = max |αi |
i

where | | is the p-adic valuation on Qp . It is not hard to see that k k is a real-valued function
which satisfies the following three properties:

(i) kxk ≥ 0 for all x ∈ V and kxk = 0 ⇔ x = 0;

56
(ii) kαxk = |α|kxk for all α ∈ Qp and x ∈ V ;

(iii) kx + yk ≤ max{kxk, kyk}, with equality sign holds when kxk =


6 kyk.
Technically speaking, V becomes a normed vector space over Qp . We can make V into a
metric space by defining the distance between two vectors x and y to be kx − yk. Addition
and scalar multiplication will become continuous operations with respect to the topology
induced by k k.
Let L be the vector space of linear operators on V . Consider a typical element σ ∈ L.
For each j, let
n
X
σ(xj ) = αij xi .
i=1
Define the norm of σ to be

kσk = max |αij | = max kσ(xj )k.


i,j j

Then k k makes L into a normed vector space over Qp . Therefore, L is a metric space
and the distance between σ and τ is kσ − τ k. The addition and scalar multiplication are
continuous operations on L. However, we also have multiplication in L to consider. We
find that
kσ(x)k ≤ kσkkxk
for all σ ∈ L and x ∈ V . This shows that

kστ k ≤ kσkkτ k

for any σ and τ in L. Therefore, multiplication is continuous on L.


For any σ ∈ O(V ), we have det(σ) = ±1. Therefore kσk ≥ 1. Then kσk = 1 if and only
if kσk ≤ 1, and this is equivalent to σ(M ) ⊆ M . Hence σ(M ) = M since the discriminant
of σ(M ) is the same as that of M . Therefore,

O(M ) = {σ ∈ O(V ) : kσk = 1}.

Let {x01 , . . . , x0n } be another basis for V . Let k k0 denote the norm on V and L defined
by this new basis. Suppose that
n
X n
X
x0j = aij xi and xj = bij x0i .
i=1 i=1

Let A = maxij |aij | and B = maxij |bij |. Then for any x ∈ V and σ ∈ L,
kxk
≤ kxk0 ≤ Bkxk
A
and
kσk
≤ kσk0 ≤ ABkσk.
AB

57
Therefore, k k and k k0 induce equivalent topology on V and L. Now let K be the Zp -lattice
Zp x01 + · · · + Zp x0n . If σ ∈ O(V ) is sufficiently close to 1 under the metric induced by k k,
then kσ − 1k0 < 1. Each such σ satisfies kσk0 = 1, hence σ(K) = K.
Now consider a third Zp -lattice N on V , and suppose that N = λ(M ) for some λ ∈ O(V ).
We claim that σ(M ) = N for all σ ∈ O(V ) which are sufficiently close to λ. For, by choosing
σ sufficiently close to λ we can make

kλ−1 σ − 1k ≤ kλ−1 k kσ − λk

arbitrary small. But all λ−1 σ which are sufficiently close to 1 make λ−1 σ(M ) = M . Hence
all σ which are sufficiently close to λ make σ(M ) = λ(M ) = N .

8.2 Approximation for Isometries


Let V be a nondegenerate quadratic space over Q. Fix a basis for V and use it to define
the norms k kp on O(Vp ) for all primes p. The following Strong Approximation Theorem is
due to Eichler.

Theorem 8.1 (Strong Approximation for O0 (V )) Let V be a nondegenerate quadratic space


over Q with dim(V ) ≥ 3. Suppose that Vν is isotropic for some place ν (ν could be ∞ or
a prime). Let S be a finite set of primes with ν 6∈ S. For any  > 0 and any given family
{σp ∈ O0 (Vp ) : p ∈ S}, there is an isometry σ ∈ O0 (V ) such that

(i) kσ − σp kp <  for all p ∈ S;

(ii) kσkp = 1 for all p 6∈ S ∪ {ν}.

A quadratic space (V, Q) over Q is called indefinite if V∞ is an indefinite quadratic space


over R. This implies that Q takes on both positive and negative rational numbers. The
follow theorem is a consequence of the Strong Approximation Theorem.

Theorem 8.2 Let L be a lattice on a nondegenerate quadratic space V over Q. If V is


indefinite and dim(V ) ≥ 3, then cls+ (L) = spn+ (L).
Proof. Let K be a lattice in spn+ (L). There exist isometries φ ∈ O+ (V ) and σp ∈ O0 (Vp )
at each p such that
φ(Kp ) = σp (Lp ) for all primes p.
Let T be the set {p : φ(Kp ) 6= Lp }, which is a finite set. Let S be the set of primes p for
which p 6∈ T and Lp is not the Zp -lattice spanned by the basis for V that defines k kp . Note
that S is also a finite set. By the Strong Approximation Theorem, there exists σ ∈ O0 (V )
such that

(i) for all p 6∈ T ∪ S, kσkp = 1;

(ii) for all p ∈ T , σ and σp are sufficiently close so that σ(Lp ) = σp (Lp );

58
(iii) for all p ∈ S, σ and 1Vp are sufficient close so that σ(Lp ) = Lp .

For any p 6∈ T , σ(Lp ) = Lp = φ(Kp ). For any p ∈ T , σ(Lp ) = σp (Lp ) = φ(Kp ). As a


result, σ(Lp ) = φ(Kp ) for all p and hence K = φ−1 σ(L). In particular, K is in cls+ (L).2

Corollary 8.3 Let L be a lattice on an indefinite quadratic space of dimension ≥ 4 over


Q. If a is represented by gen(L), then a is represented by L.
Proof. This follows from Theorems 7.21 and 8.2.2
Corollary 8.3 does not hold for ternary lattices. For example, let L be the Z-lattice
corresponding to the quadratic form −9x2 + 2xy + 7y 2 + 2z 2 . In terms of symmetric matrix
 
−9 1 0
L∼ =  1 7 0 .
0 0 2

Since d(L) = −27 , Lp is unimodular for all primes p ≥ 3. In particular, Lp represents 1


for all p ≥ 3. Over Z2 , −9 is −1 times a square of a unit. Therefore, L2 also represents 1.
This means that gen(L) represents 1. We claim that L does not represent 1. The following
elementary proof is due to D. Zagier.
Assume on the contrary that L represents 1, i.e. there exist integers x, y, z such that

−9x2 + 2xy + 7y 2 + 2z 2 = 1.

We may rewrite this equation as

2z 2 − 1 = (x − y)2 + 8(x − y)(x + y).

The left hand side is odd; hence (x − y) and hence x + y as well are odd. This shows that
the right hand side is congruent to 1 mod 8, and thus z is odd and then 2z 2 − 1 ≡ 1 mod
16. But 8(x − y)(x + y) ≡ 8 mod 16. If follows that (x − y)2 ≡ 9 mod 16 and hence

(x − y) ≡ ±3 mod 8.

In particular, (x − y), and 2z 2 − 1 as well, has a prime factor p ≡ ±3 mod 8. On the other
hand, if such p divides 2z 2 − 1, then 2 is a square mod p and it follows that p ≡ ±1 mod 8
which is a contradiction.
Here is another consequence of the Strong Approximation Theorem. Its proof is similar
to that of Theorem 8.2

Theorem 8.4 Let L be a Z-lattice on a nondegenerate quadratic space V over Q, and q


be a prime for which Vq is isotropic. If dim(V ) ≥ 3, then Z[1/q]L ∼
= Z[1/q]M for any
+
M ∈ spn (L).
Suppose that V is an indefinite quadratic space over Q. Let L be a Z-lattice on V
and let M be a Z-lattice in gen(L). Let T be a finite set of primes such that Lp = Mp

59
for all p 6∈ T . For each p ∈ T , let σp ∈ O+ (Vp ) such that σp (Mp ) = Lp . If the Strong
Approximation Theorem held for O+ (V ), then there would be an isometry σ ∈ O+ (V ) such
that σ(Mp ) = σp (Mp ) for all p ∈ T and that σ(Mp ) = Mp for all p 6∈ T . Then σ(M ) = L
and this means that gen(L) has only one class. But from the class number formula obtained
in the next section we know that there exists indefinite Z-lattice whose class number is not
1. Therefore, the Strong Approximation Theorem does not hold for O+ (V ). However, there
is a weak approximation for O+ (V ).

Theorem 8.5 (Weak Approximation for O+ (V )) Let V be a nondegenerate quadratic space


over Q and let T be a finite set of primes on Q. Suppose that σp is given in O+ (Vp ) at each
p ∈ T . Then for every  > 0, there is a σ ∈ O+ (V ) such that

kσ − σp kp <  for all p ∈ T.

Proof. For each p ∈ T , we may write

σp = τx1 (p) · · · τxm (p)

where every xi(p) is in Vp . Since m must be even and all symmetries have order 2, we may
suppose that m is independent of p ∈ T . By Chinese Remainder Theorem, we can choose
xi ∈ V such that xi and xi(p) are sufficiently close in Vp for p ∈ T . Then τxi and τxi(p) are
sufficiently close so that σp and σ := τx1 · · · τxm are, for p ∈ T .2

Corollary 8.6 Let L be a Z-lattice on a nondegenerate quadratic space V over Q, and let
T be a finite set of primes. For every Z-lattice M in gen(L), there exists K ∈ cls+ (M ) such
that Kp = Lp for all p ∈ T .
Proof. For each p ∈ T , let σp ∈ O+ (Vp ) such that σp (Mp ) = Lp . It follows from the Weak
Approximation Theorem for O+ (V ) that there is a σ ∈ O+ (V ) satisfying σ(Mp ) = σp (Mp ).
Let K = σ(M ). Then Kp = Lp for all p ∈ T .2
Finally, we can improve Theorem 8.3 to representations with approximation property.

Theorem 8.7 Let L be a Z-lattice on an indefinite quadratic space V over Q with dim(V ) ≥
4 and T be a finite set of primes. Let a be a nonzero rational number which is represented
by gen(L), and suppose that for each p ∈ T there is a vp ∈ Lp with Q(vp ) = a. Then for
each  > 0, there exists v ∈ L with Q(v) = a and

kv − vp kp <  ∀ p ∈ T,

where all k kp are defined by L


Proof. Since a is represented by gen(L), there exists z ∈ V such that Q(z) = a. We enlarge
T so that

60
(a) T contains all the prime divisors of 2d(L), and

(b) z ∈ Lp for all p 6∈ T .

It suffices to prove the theorem for this enlarged T .


By Witt’s theorem, there is an isometry ρp ∈ O+ (Vp ) at each p ∈ T such that ρp (z) =
vp . Let W be the orthogonal complement of Qz in V . Since dim(W ) ≥ 3, there exists
τp ∈ O+ (Wp ) at each p ∈ T such that θ(τp ) = θ(ρp ). Let σp = ρp τp . Then σp ∈ O0 (Vp ) for
all p ∈ T . By the Strong Approximation for O0 (V ), there exists σ ∈ O0 (V ) such that

kσ − σp kp < ∀p ∈ T
kzkp

and
kσkp = 1 ∀ p 6∈ T.
Let v be σ(z). It is clear that kv − vp kp <  for each p ∈ T . Since vp ∈ Lp and  < 1,
v ∈ Lp as well. For p 6∈ T , σ ∈ O0 (Lp ) and z ∈ Lp . So v = σ(z) ∈ Lp . Consequently, v ∈ Lp
for all primes p; hence v ∈ L.2

9 Adelic Theory of Spinor Genus


In this section, let V be a nondegenerate quadratic space over Q and L be a Z-lattice on
V . We shall provide a formula for the number of proper spinor genera in gen(L). When V
is indefinite, this formula gives the number of proper classes in gen(L). We also describe a
procedure to decide if two given lattices in a given genus are inside the same proper spinor
genus.

9.1 Adeles
Let Ω be the set of all places of Q. Fix a basis B of V Q and use it to define all the norms
k kp on O(Vp ). If Σ is an element in the direct product ν∈Ω O+ (Vν ), then Σν will denote
its ν-component. The adelization of O+ (V ) is the set
Y
OA+ (V ) = {Σ ∈ O+ (Vν ) : kΣp kp = 1 for almost all primes p }.
ν∈Ω

Lemma 9.1 OA+ (V ) is well-defined, that is, it is independent of the choice of the basis B.
Proof. Let L be the lattice spanned by the vectors in B. Then for a prime p, kΣp kp = 1 if
and only if Σp (Lp ) = Lp . Now, suppose that B0 is another basis for V . Let L0 be the lattice
spanned by the vectors in B0 . Then for almost all p, Lp = L0p . So, if Σ is in OA+ (V ), then

Σp (L0p ) = Σp (Lp ) = Lp = L0p

for almost all primes p.2

61
An element in OA+ (V ) is called an adele of O+ (V ). It is clear that OA+ (V ) is a subgroup
of the direct product. The set of all elements Σ ∈ OA+ (V ) with the property

Σν ∈ O0 (Vν ) for all ν ∈ Ω

is clearly a subgroup of OA+ (V ). We shall denote this subgroup by OA0 (V ). It is evident that
OA0 (V ) contains the commutator subgroup of OA+ (V ).
Consider a typical element σ ∈ O+ (V ). Then σ can be regarded as an element in O+ (Vν )
for all ν ∈ Ω. Let T be the matrix representation of σ with respect to the basis B. Then
the entries of T and det(T ) are units in Zp for almost all p. Therefore, kσkp = 1 for almost
all primes p. Hence we can identify O+ (V ) as a subgroup of OA+ (V ) through the diagonal
embedding.
For any Σ ∈ OA+ (V ) and any lattice L on V , define the lattice Σ(L) by specifying its
local completions as
Σ(L)p = Σp (Lp ) for all p.
This definition is meaningful since Σp (Lp ) = Lp for almost all p. Therefore, we have an
action of the group OA+ (V ) on the set of lattices on V .

Proposition 9.2 We have cls+ (L) = O+ (V )L, spn+ (L) = O+ (V )OA0 (V )L, and gen(L) =
OA+ (V )L.
Proof. We only demonstrate the equality gen(L) = OA+ (V )L; the other two equalities can
be proved in a similar way. It is clear that Σ(L) ∈ gen(L) for any Σ ∈ OA+ (V ). Suppose
that M ∈ gen(L). Then for each prime p, there exists Σp ∈ O+ (Vp ) such that

Σp (Lp ) = Mp .

Since Lp = Mp for almost all p, therefore, kΣp kp = 1 for almost all p. Let Σ∞ be any
isometry in O+ (V∞ ) and put Σ = (Σν ). Then M = Σ(L).2
The stabilizer of L in OA+ (V ) is the set

OA+ (L) = {Σ ∈ OA+ (V ) : Σν (Lp ) = Lp for all primes p}.

9.2 Number of Spinor Genera


Lemma 9.3 For any Σ ∈ OA+ (V ), we have Σ(spn+ (L)) = spn+ (Σ(L)).
Proof. Let K be a lattice in Σ(spn+ (L)). Then there exist φ ∈ O+ (V ) and σp ∈ O0 (Vp ) at
each p such that
Kp = Σp φσp (Lp ) for all p.
Since O0 (Vp ) contains the commutator subgroup of O+ (Vp ), there exists ρp ∈ O0 (Vp ) such
that
Σp φσp = ρp φσp Σp .

62
However, as O0 (Vp ) is normal in O+ (Vp ), there exists τp ∈ O0 (Vp ) such that

ρp φσp Σp = φτp σp Σp .

Therefore, K ∈ spn+ (Σ(L)), and hence Σ(spn+ (L)) ⊆ spn+ (Σ(L)). The reverse inclusion
can be proved in the same fashion.2
If Σ, Λ are elements in OA+ (V ), then the normality of OA0 (V ) in OA+ (V ) implies that
ΣOA0 (V ) = OA0 (V )Σ, and the fact that OA0 (V ) contains the commutator subgroup of OA+ (V )
implies that ΣΛOA0 (V ) = ΛΣOA0 (V ), hence the set OA0 (V )ΣΛ is independent of the order
of OA0 (V ), Σ, Λ. From this it follows that the set O+ (V )OA0 (V )OA+ (L) is independent of the
order of O+ (V ), OA0 (V ), OA+ (L), and that this set is actually equal to the group generated
by O+ (V ), OA0 (V ), OA+ (L), which is a normal subgroup in OA+ (V ).

Proposition 9.4 The number of proper spinor genera in gen(L) is equal to the index

[OA+ (V ) : O+ (V )OA0 (V )OA+ (L)].

Proof. It is clear that the group OA+ (V ) acts transitively on the set of proper spinor genera
in gen(L) and O+ (V )OA0 (V )OA+ (L) is contained in the stabilizer of spn+ (L). Now, suppose
that Σ(spn+ (L)) = spn+ (Σ(L)) = spn+ (L) for some Σ ∈ OA+ (V ). Then Σ(L) ∈ spn+ (L)
and this means that there exist φ ∈ O+ (V ) and σp ∈ O0 (Vp ) at each p such that

Σp (Lp ) = φσp (Lp ) for all p.

Therefore, Σ ∈ O+ (V )OA0 (V )OA+ (L).2

Corollary 9.5 The number of proper spinor genera in a given genus is power of 2.
Proof. At each place ν, the square of any element in O+ (Vν ) has trivial spinor norm.
Therefore, the square of element in OA+ (V ) is in OA0 (V ), and hence the exponent of the
finite abelian group OA+ (V )/O+ (V )OA0 (V )OA+ (L) is 2.2
The set of ideles of Q is the set
Y
J = {x = (xν ) ∈ Q×
ν : |xp |p = 1 for almost all p }.
ν∈Ω

It is clear that J is a group.

Lemma 9.6 If p > 2 and Lp is a unimodular Zp -lattice of rank ≥ 2, then θ(O+ (Lp )) =
Z× ×2
p Qp .

Proof. By Lemma 6.19, Lp represents all units of Zp . Therefore, θ(O+ (Lp )) contains Z× ×2
p Qp .
+
For the reverse inclusion, we first note that O (Lp ) is generated by symmetries (see Remark
7.16). Therefore it is enough to show that the spinor norm of any symmetry in O+ (Lp ) is a

63
unit. Take a typical symmetry τy ∈ O+ (Lp ). We may assume that y is primitive in Lp and
thus p−1 y 6∈ Lp . For any x ∈ Lp ,

2B(x, y)
τy (x) = x − y.
Q(y)
Therefore B(y, Lp ) ⊆ Q(y)Zp . But B(y, Lp ) = Zp because Lp is unimodular. Hence θ(τy ) =
Q(y)Q×2 × ×2
p ⊆ Zp Qp .2

Corollary 9.7 Let L be a lattice on V . If dim(V ) ≥ 2, then for any Σ ∈ OA+ (V ), θ(Σp ) ⊆
Z× ×2
p Qp for almost all p.

Proof. This is because Lp is unimodular for almost all p.2


Let Σ be an element in OA+ (V ). If rank(L) ≥ 3, then by Corollary 9.7 there exists an
x ∈ J such that xν ∈ θ(Σν ) for all ν ∈ Ω. If y is another element in J which is associated
to Σ in this way, then y ∈ xJ 2 . Let JL be the subset of J defined by

JL = {x = (xν ) ∈ J : xν ∈ θ(O+ (Lν )) for all ν ∈ Ω }.

Then J 2 ⊆ JL , and hence the image of x and y in J/Q× JL are equal. We therefore have a
well-defined homomorphism
θA : OA+ (V ) −→ J/Q× JL .

Theorem 9.8 If rank(L) ≥ 3, then the number of proper spinor genera in gen(L) is equal
to the index [J : Q× JL ].
Proof. It suffices to show that θA is a surjective homomorphism with O+ (V )OA0 (V )OA+ (L)
as its kernel. Since the definition of OA+ (V ) is independent of the choice of the basis B of
V , we may very well assume that B is in fact a basis for L.
Take a typical element x ∈ J. By considering −x instead if necessary, we may assume
that x∞ ∈ θ(Σ∞ ) for some Σ∞ ∈ O+ (V∞ ). For almost all p, xp is a unit. Since Lp is
unimodular for almost all p and θ(O+ (Lp )) contains Z× +
p at those p, we can find Σp ∈ O (Lp )
such that xp ∈ θ(Σp ) for almost all p. For the remaining finite number of primes q, there
exist Σq ∈ O+ (Vp ) such that xq ∈ θ(Σq ) because dim(Vq ) ≥ 3; see Theorem 7.20. Then
Σ = (Σν ) is an element in OA+ (V ) and θA (Σ) = x.
It is clear that O+ (V )OA0 (V )OA+ (L) is a part of the kernel of θA . Now suppose that
θA (Σ) ∈ Q× JL . Let us assume the following statement at this moment:

{a ∈ Q× : a > 0} if V∞ is anisotropic,

θ(O+ (V )) =
Q× otherwise.

Then we can take σ ∈ O+ (V ) and Λ ∈ OA+ (L) so that

θA (ΣσΛ) ⊆ J 2 .

As a result, ΣσΛ ∈ OA0 (V ) and hence the kernel of θA is O+ (V )OA0 (V )OA+ (L).2

64
Proposition 9.9 If dim(V ) ≥ 3, then

{a ∈ Q× : a > 0}

+ if V∞ is anisotropic,
θ(O (V )) =
Q× otherwise.

Proof. Let a ∈ Q× and suppose that a > 0 if V∞ is anisotropic. If the latter situation
arises, then by multiplying −1 to the quadratic form if necessary, we may assume that V∞
is negative definite.
Let S be the set of primes for which Vp is anisotropic. Note that S is a finite set. If
p ∈ S, we take bp ∈ Q× p such that both bp d(V ) and abp d(V ) are not squares. Using the
Chinese Remainder Theorem we can find 0 < b ∈ Q× such that b and bp are sufficiently
close for all p ∈ S. If S is empty, then we set b = 1. Then both Vv ⊥ hbi and Vv ⊥ habi are
isotropic for every place v of Q. From the Strong Hasse principle it follows that both −b
and −ab are represented by V . Let x, y ∈ V such that Q(x) = −b and Q(y) = −ab. Then
θ(τx τy ) = Q(x)Q(y) ∈ aQ×2 . Thus

{a ∈ Q× : a > 0} if V∞ is anisotropic,

θ(O+ (V )) ⊇
Q× otherwise.

The reverse inclusion is obvious.2

Corollary 9.10 Let L be a nondegenerate Z-lattice. If rank(L) ≥ 3 and θ(O+ (Lp )) ⊇ Z×


p
for all primes p, then gen(L) = spn+ (L).
Proof. It suffices to show that the quotient group J/Q× JL is trivial. Take a typical element
x ∈ J. By multiplying −1 if necessary, we could assume that x∞ > 0 and hence x∞ ∈
θ(O+ (V∞ )). Let S be the set of primes p for which xp 6∈ Z× ×2
p Qp . Then there exists a
rational number a > 0 such that ordp (a) = 0 for all p 6∈ S and axp ∈ Z× p . As a result,
ax ∈ JL and hence x ∈ Q× JL .2

Example 9.11 Let L be a nondegenerate unimodular Z-lattice of rank ≥ 3. If p > 2, then


θ(O+ (Lp )) = Z× ×2
p Qp since Lp is unimodular. At p = 2, P is an orthogonal summand of L2 ,
where P = H or A, and θ(O+ (P)) ⊇ Z× ×2 +
2 Q2 . Therefore, gen(L) = spn (L). If, in addition,
L is indefinite, then gen(L) = cls+ (L) and the class number of L is 1.

9.3 Spinor Equivalence


Let L be a Z-lattice on a nondegenerate space V , and K is another Z-lattice in gen(L).
How do we know if K is inside spn+ (L)? By definition, K = ΣL for some Σ ∈ OA+ (V ).
Since the adelic spinor norm θA induces an isomorphism

OA+ (V )/O(V )OA0 (V )OA+ (L) −→ J/Q× JL ,

therefore K is in spn+ (L) if and only if θA (Σ) ∈ Q× JL . The remaining question is: how do
we construct Σ if L and K are given explicit enough?

65
It is clear that we can replace K by any lattice in cls+ (K). So by virtue of the Weak
approximation Theorem we can assume that L and K only differ at the primes not dividing
2d(L), that is Lp = Kp for all primes p | 2d(L). Now take a prime p so that Lp 6= Kp . Then
p > 2 and both Lp and Kp are unimodular Zp -lattices. Thus [Lp : Lp ∩ Kp ] = [Kp : Lp ∩ Kp ]
by Corollary 6.14. Consequently, [L : L ∩ K] = [K : L ∩ K].

Lemma 9.12 Suppose Lp 6= Kp . If σp is an isometry in O+ (Vp ) which sends Lp to Kp ,


then θ(σp ) ∈ [Lp : Lp ∩ Kp ]Z× ×2
p Qp .

Proof. For simplicity, let e = ordp ([Lp : Lp ∩ Kp ]). Since θ(O+ (Lp )) = Z× ×2
p Qp , it suffices
+ e
to exhibit an isometry in O (Vp ) of spinor norm p which sends Lp to Kp . By Theorem
6.12, we may assume that Lp and Kp are unimodular lattices on a hyperbolic plane with
a hyperbolic pair x, y such that Lp = Zp x + Zp y and Kp = Zp pe x + Zp p−e y. Then the
isometry τx−y τx−p−e y , which has spinor norm pe , sends Lp to Kp .2

Definition 9.13 The intersecting idele of L and K is the idele j(L, K) ∈ J whose p-
component is pordp ([L:L∩K]) .

Theorem 9.14 Let L and K be Z-lattices in the same genus. If [L : L ∩ K] is relatively


prime to 2d(L), then K ∈ spn+ (L) if and only if j(L, K) ∈ Q× JL .
Proof. If Σ(L) = K where Σ ∈ OA+ (V ), then θA (Σ) ∈ j(L, K)Q× JL by Lemma 9.12. Thus
K ∈ spn+ (L) if and only if j(L, K) ∈ Q× JL .2

10 Representations of Spinor Genus


Throughout this section, L is a Z-lattice on a nondegenerate space V of dimension ≥ 3.
Suppose that a is a nonzero rational number which is represented by gen(L). Then a is
represented by some lattice in gen(L). We say that a is represented by a proper spinor
genus S in gen(L) if a is represented by some lattice in S. The main question we would like
to answer is: which proper spinor genera in gen(L) represent a?
Let v ∈ V such that Q(v) = a and let W be the orthogonal complement of the space
spanned by v. For the questions with which we are concerned we may, without loss of
generality, assume that v ∈ L.

10.1 Generators
An adele Σ = (Σp ) ∈ OA+ (V ) is called a generator for L and v if v ∈ Σ(L). Let XA (L, v) be
the set of generators for L and v. It is clear from the definition that

XA (L, v)OA+ (L) = OA+ (W )XA (L, v) = XA (L, v).

The importance of XA (L, v) is illustrated in the following lemma.

66
Lemma 10.1 Let Σ ∈ OA+ (V ). Then a is represented by spn+ (Σ(L)) if and only if Σ ∈
O+ (V )OA0 (V )XA (L, v).
Proof. This is clear from the definitions.2
Note that in general XA (L, v) may not be a group. However, we have
Theorem 10.2 The set OA0 (V )XA (L, v) is a group.
Proof. It suffices to prove that θA (XA (L, v)) is a group. Let p be a prime, and let X(Lp , v)
be the set of local generators, that is,
X(Lp , v) = {σ ∈ O+ (Vp ) : v ∈ σ(Lp )}.
We need to show that θ(X(Lp , v)) is a group. Since X(Lp , v)O+ (Lp ) = X(Lp , v), it is only
necessary to consider the case when
[Q× +
p : θ(O (Lp ))] > 2.

This implies that Lp has only one dimensional Jordan components. When p > 2 this
follows from the fact that the spinor norm of the isometry group of any binary unimodular
Zp lattices contains all p-adic units (Lemma 9.6). For p = 2, this follows from the explicit
calculation of spinor norms of the isometry groups of modular Z2 -lattices by Hsia.
So, we may write Lp = Zp x1 ⊥ · · · ⊥ Zp xn with ordp (Q(x1 )) < · · · < ordp (Q(xn )). We
may assume that ordp (Q(v)) = ordp (Q(x1 )); otherwise one can replace Lp by L0p = Zp px1 ⊥
· · · ⊥ Zp xn because X(Lp , v) = X(L0p , v) and the argument can be repeated. Also, x1 = v
can be assumed.
Let σ ∈ X(Lp , v). Then σ −1 (v) ∈ Lp and hence σ −1 (v) = αv + w where α ∈ Z× p and
w ∈ W . One of 1 − α and 1 + α is not in 2pZp . Without loss of generality, we assume that
1 − α 6∈ 2pZp . A direct verification shows that τσ−1 v−v is in O(Lp ) and τσ−1 v−v (v) = σ −1 (v).
Therefore,
σ ◦ τσ−1 v−v ◦ τx2 (v) = v
which implies that σ ∈ O+ (Wp )O+ (Lp ). However, it is clear from the definitions that
O+ (Wp )O+ (Lp ) ⊆ X(Lp , v). Thus X(Lp , v) = O+ (Wp )O+ (Lp ) and θ(X(Lp , v)) is a sub-
group of J.2
We therefore have the following chain of containment of groups
O+ (V )OA0 (V )OA+ (L) ⊆ O+ (V )OA0 (V )OA+ (L)OA+ (W ) ⊆ O+ (V )OA0 (V )XA (L, v) ⊆ OA+ (V ).
Lemma 10.3 We have OA+ (V )/O+ (V )OA0 (V )OA+ (L)OA+ (W ) ∼
= J/Q× JL θA (OA+ (W )) and
OA+ (V )/O+ (V )OA0 (V )XA (L, v) ∼
= J/Q× θA (XA (L, v)).

Proof. This follows from the application of the adelic spinor norm map θA .2
Finally, we show that everything discussed thus far is independent of the choice of v and
L.

67
Lemma 10.4 Suppose that v 0 ∈ L0 with Q(v 0 ) = Q(v) and L0 ∈ gen(L). Then the
groups generated by XA (L, v) and XA (L0 , v 0 ) are conjugate in OA+ (V ) so that the subgroup
O+ (V )OA0 (V )XA (L, v) is independent of the choice of v and L.
Proof. Let G(L, v) be the group generated by XA (L, v). Let Σ ∈ OA+ (V ) such that L0 =
Σ(L). We first suppose that v = v 0 . Then Σ ∈ XA (L, v) by definition. If g ∈ XA (L0 , v), then
v ∈ gΣ(L); thus gΣ ∈ XA (L, v) and hence g ∈ G(L, v). This shows that G(L0 , v) ⊆ G(L, v).
By reversing the roles of L and L0 we see that G(L, v) = G(L0 , v).
In general, let φ ∈ O+ (V ) such that φ(v) = v 0 . Then

φG(L, v)φ−1 = G(φ(L), φ(v)) = G(φ(L), v 0 ) = G(L0 , v 0 ).

2
From now on we shall write XA (L, a) for XA (L, v).

10.2 Representations
We know from the previous subsection that the number of proper spinor genera in gen(L)
representing a is given by the following indices

[O+ (V )OA0 (V )XA (L, a) : O+ (V )OA0 (V )OA+ (L)] = [Q× θA (L, a) : Q× JL ].

For simplicity, we have used θA (L, a) to denote θA (XA (L, a)).


Suppose first that either dim(W ) ≥ 3 or W is a hyperbolic plane. Then θ(O+ (Wp )) =
+
Q× ×
p for all primes p. Therefore Q θA (OA (W )) = J, so that a is represented by every proper
spinor genus in gen(L). Note that this recaptures Theorem 7.21.
√Consider next the case dim(W ) = 2 and W is anisotropic. Let δ = d(W ) and E =
Q( −δ). Then E is a quadratic extension of Q. Let ΩE be the set of places of E (that
is, the set of equivalence classes of valuations on E). Each place ν of Q has one or two
extensions in ΩE . Let JE be the group of ideles of E defined by
Y
JE = {(xω ) ∈ Eω× : |xω |ω = 1 for almost all ω}.
ω∈ΩE

Define the norm map NE/Q : JE → J by


Y
NE/Q ((xω )) = ( Nω/ν (xω )),
ω|ν

where Nω/ν is the norm of the extension Eω /Qν .

Remark 10.5 The definitions of JE and the associated norm map NE/Q extend naturally
to all finite extensions of Q.

68
Lemma 10.6 We have θA (OA+ (W )) = NE/Q (JE ).
Proof. It suffices to show that θ(O+ (Wν )) = ω|ν Nω/ν (Eω× ) for every place ν of Q. This is
Q

obvious when ν splits since both groups are equal to Q× ν.


Suppose that ν does not split in E. Then there is only one place ω of E lying above
ν and Eω is a quadratic extension of Qv . Take an anisotropic vector x ∈ W and let
α = Q(x). Then W ∼ = hαi ⊥ hαδi. Thus, after scaling by α, the quadratic space Wν can be
identified with Eω and the quadratic map is simply the norm Nω/ν . Therefore, if t is given
in Nω/ν (Eν× ), then there must be a vector y ∈ W such that αQ(y) = t and so

t = α · α−1 t = θ(τx τy ) ∈ θ(O+ (Wν )).

Conversely, let σ ∈ O+ (Wν ). Then σ is a product of even number of symmetries of Wν .


It follows from the above discussion that θ(σ) ∈ Nω/ν (Eω× ). 2
By class field theory, [J : Q× NE/Q (JE )] = [E : Q] = 2. We offer another proof here.

Lemma 10.7 If E/Q is a quadratic extension, then [J : Q× NE/Q (JE )] = 2.


Proof. Define a function R : J → {±1} by
Y
R((xν )) = (xν , −δ)ν ,
ν

where ( , )ν is the Hilbert symbol over Qν . It is well-defined since for almost all odd primes
p, xp and −δ are p-adic units and hence (xp , −δ)p = 1. Since −δ 6∈ Q×2 , there is a place ν
at which −δ 6∈ Q×2 ×
ν . Hence there must be xν ∈ Qν such that (xν , −δ)ν = −1. This shows
that R is surjective. We shall show that ker R is precisely Q× NE/Q (JE ).

Recall that (xν , −δ)ν = 1 if and only if xν is a norm of the field extension Qp ( −δ)/Qp .
Thus NE/Q (JE ) ⊆ ker R. It follows from Hilbert Reciprocity Law that Q× ⊆ ker R. Hence
Q× NE/Q (JE ) ⊆ ker R. For the other inclusion, let (xν ) ∈ ker R and let ν1 , . . . , νm be the
places for which (xν , −δ)ν = −1. Note that m is even and nonzero. Let S be a finite set of
primes which contains the prime 2 and all the finite places among the νi , and such that for
any p 6∈ S, xp and δ are p-adic units. Put ep = ordp (xp ) and define
Y
h= pep .
p∈S

We take a prime q 6∈ S such that

q ≡ ηh−1 xp mod pt , for all p ∈ S,

where t is a sufficiently large integer and η = 1 if all the places νi are finite; otherwise
η = −1. We claim that qηh−1 (xν ) ∈ NE/Q (JE ), that is (qηh−1 xν , −δ)ν = 1 for all places ν.
This will complete the proof.
If a prime p is not in S and p 6= q, then p is odd and qηh−1 xp and −δ are in Z× p,
−1 −1 ×
and thus (qηh xp , −δ)p = 1. If p is in S, then qηh xp is a square in Qp (since t is

69
chosen large enough) and hence (qηh−1 xp , −δ)p = 1 also. If all the places νi are finite, then
qηh−1 > 0 and then (qηh−1 x∞ , −δ)∞ = (x∞ , −δ) = 1. If νi = ∞ for some i, then η = −1
and (x∞ , −δ)∞ = −1. This means that x∞ < 0. So, (qηh−1 x∞ , −δ)∞ = (−x∞ , −δ)∞ = 1.
Finally at the prime q, since qηh−1 (xν ) ∈ ker R, we have
Y
(qηh−1 xq , −δ)q = (qηh−1 xν , −δ)ν = 1.
ν6=q

2
Since there are inclusions of groups

Q× NE/Q (JE ) ⊆ Q× θA (L, a) ⊆ J,

therefore, Q× θA (L, a) 6= J if and only if Q× θA (L, a) = Q× NE/Q (JE ).

Lemma 10.8 If Q× θA (L, a) = Q× NE/Q (JE ), then θA (L, a) = NE/Q (JE ).


Proof. Assume the contrary that there exists (xν ) ∈ θA (L, a) which is not in NE/Q (JE ).
Then there must be a place ` for which (x` , −δ)` = −1. Construct an idele j = (jν ) such
that jν = 1 if ν 6= ` and jν = xν . Then j ∈ θA (L, a) and so we can write j = αn for some
α ∈ Q× and n ∈ NE/Q (JE ). For all ν 6= `, since j` = 1, we have (α, −δ)ν = 1. By Hilbert
Reciprocity Law, (α, −δ)` = 1 as well. Thus (xν , −δ)ν = 1 which is a contradiction.2
Therefore, Q× θA (L, a) 6= J if and only θA (L, a) = NE/Q (JE ). When this is the situation
we shall call a a spinor exception. In this case, we have

[J : Q× JL ] [J : Q× JL ] 1
[Q× θA (L, a) : Q× JL ] = ×
= ×
= [J : Q× JL ]
[J : Q θA (L, a)] [J : Q NE/Q (JE )] 2

so that a is represented by exactly half of the proper spinor genera in gen(L). To summarize:

Theorem 10.9 Suppose that a is represented by the genus of a ternary Z-lattice L and
−ad(L) 6∈ Q×2 . Then a is a spinor exception if and only if θA (L, a) = NE/Q (JE ), where
E = Q( −ad(L)). Moreover, spn+ (ΣL) represents a if and only if θA (Σ) ∈ Q× θA (L, a).
p

The groups θA (L, a) are explicitly determined by Schulze-Pillot.

10.3 A Class Field Interpretation


Let F be a number field and O be the ring of integers in F . Let ΩF be the set of places of
F . The group of ideles of F is defined by
Y
JF = {(xν ) ∈ Fν× : xν ∈ Oν× for almost all ν}.
ν∈ΩF

70
Let ΩfF and Ω∞F be the set of finite and infinite, respectively, places of F . We impose a
topology on JF as follows. Let
Y Y
E= Fν× Oν× .
ν∈Ω∞
F ν∈ΩfF

It is easy to see that E is a subgroup of JF . Each factor of E has its metric topology induced
by the associated valuation. Each Fν× is locally compact while each Oν× is compact. We
give E the product topology so that it becomes a locally compact topological group. Next
we consider the following collection of subsets of JF
{jA : j ∈ JF and A is open in E}.
We declare each subset in this collection open in JF . It turns out that these open subsets
form a basis of a topology on JF which we call the restricted product topology.
An open subgroup H of JF is called principal if F × ⊆ H and [JF : H] < ∞. For a finite
abelian extension E/F , F × NE/F (JE ) is a principal open subgroup of JF . A main theorem
in Class Field Theory says that
E 7→ F × NE/F (JE )
is a bijective inclusion-reversing correspondence between finite abelian extensions of F and
principal open subgroups of JF . Moreover, Artin Reciprocity Law, which is a generalization
of Hilbert Reciprocity Law, gives an isomorphism
JF /F × NE/F (JE ) −→ Gal(E/F ).
Let us go back to our discussion of spinor genera. In our case, the base field is Q (even
though all the results discussed thus far about spinor genera can be generalized to any
number field setting). The group Q× JL is clearly a principal open subgroup of J(= JF ).
Associated to it by Class Field Theory is an abelian extension Σ/Q. Since JL ⊇ J 2 , Σ/F
is a multiquadratic extension. We call Σ the spinor class field of gen(L).
It is clear that we can rephrase all the results, especially Theorem 10.9, in terms of the
spinor class fields. Here we offer a result that can be deduced easily from this class field
setting but is not obvious from the previous discussion. The group Q× θA (L, a) is also a
principal open subgroup of J and hence one can associate to it an abelian extension ΣL/a
called a relative spinor class field. Since we have a chain of containments of groups
Q× JL ⊆ Q× θA (L, a) ⊆ J,
we have a chain of fields
Q ⊆ ΣL/a ⊆ Σ.
p
If a is a spinor exception of gen(L), then θA (L, a) = NE/Q (JE ) where E = Q( −ad(L)).
p
In particular, ΣL/a is in fact the quadratic extension Q( −ad(L)). Since Σ can have only
finitely many quadratic subextensions of Q, therefore we have
Proposition 10.10 Let L be a ternary nondegenerate Z-lattice. Then the spinor exceptions
of gen(L) fall into finitely many square classes.

71
11 Representations of Definite Quadratic Forms
In this section, we will prove the following theorem.

Theorem 11.1 (Tartakowsky, Kloosterman, Pall and Ross) If L is a positive definite Z-


lattice of rank ≥ 5. Then there exists a constant c(L) such that a is represented by L if a
is represented by the genus of L and a > c(L).
The above theorem is false if the rank of L is 4. Consider the Z-lattice L = h1, 1, 25, 25i.
If p 6= 5, Lp is just the sum of 4 squares. Hence Q(Lp ) = Zp for p 6= 5. Since −1 is a
square in Z5 , L5 is split by h1, 1i ∼ = H; hence Q(L5 ) = Z5 . Consequently, the genus of L
represents all positive integers. We claim that L does not represent all integers of the form
3 · 2r . Let {v1 , v2 , v3 , v4 } be an orthogonal basis for L whose associated symmetric
P matrix
r
is h1, 1, 25, 25i. Suppose that there exists v ∈ L such that Q(v) = 3 · 2 . If v = ai vi , then

3 · 2r = a21 + a22 + 25a23 + 25a24 .

If r ≤ 2, then a3 = a4 = 0. But then the equation

3 · 2r = a21 + a22

is not soluble over Z. If r ≥ 3, then all the ai must be even, and hence 3 · 2r−2 is also
represented by L. Eventually we arrive at the conclusion that L represents an integer of
the form 3 · 2m with m ≤ 2, which is impossible as shown earlier.

11.1 Some Local Results


In this subsection, L is a Zp -lattice on a nondegenerate quadratic space V over Qp . Recall
that L is a maximal lattice if L is an (a)-maximal lattice for some a ∈ Q×p.

Proposition 11.2 If L is a maximal lattice on V and dim(V ) ≥ 3, then θ(O+ (L)) ⊇ Z×


p.

Proof. By scaling the quadratic form on V , we may assume that L is (2)-maximal. If V


is anisotropic, then L is the unique (2)-maximal lattice on V . This means that O+ (L) =
O+ (V ). By Lemma 7.20, θ(O+ (L)) = θ(O+ (V )) = Q× p.
Suppose that V is isotropic. Then L is split by the hyperbolic plane H; see Theorem
6.12. So, θ(O+ (L)) ⊇ θ(O+ (H)) ⊇ Z× p .2

Lemma 11.3 Suppose that p > 2 and that L is a unimodular lattice. If x and y are
primitive vectors of L with Q(x) = Q(y), then there exists σ ∈ O(L) such that σ(x) = y.
Proof. Let w = y − x. Then Q(w) = −2B(x, w). If Q(w) ∈ Z× p , then τw ∈ O(L) and
τw (x) = y. Therefore, we may assume that Q(w) ∈ (p). Suppose that we can find an u ∈ L
such that
Q(u), B(u, x), B(u, y) ∈ Z×
p. (∗)

72
Let v = y − τu (x). Then v ∈ L because τu ∈ O(L). Moreover,

Q(v) = Q(w) + 4B(u, x)B(u, y)Q(u)−1 ∈ Z×


p.

So, τv ∈ O(L) and a direct computation shows that τv (y) = τu (x). This implies τv τu (x) = y.
Now it remains to show the existence of v satisfying (∗). Let L̄ be the quadratic space
L/pL over the finite field F = Zp /pZp . Since L is unimodular, L̄ is a nondegenerate space
over F. For any vector t ∈ L, let t̄ be its canonical image in L̄. By abusing the notation,
we also use Q and B to denote the quadratic form and bilinear form, respectively, induced
on L̄. Condition (∗) is equivalent to

Q(L̄ \ (x̄⊥ ∪ ȳ ⊥ )) 6= 0.

Since x is a primitive vector of L, there exists z ∈ L such that B(z, x) = 1. Therefore,


dim(x̄⊥ ) = dim(L̄) − 1. Similarly, dim(ȳ ⊥ ) = dim(L̄) − 1.
Suppose on the contrary that

Q(L̄ \ (x̄⊥ ∪ ȳ ⊥ )) = 0.

Let t̄ ∈ x̄⊥ ∩ ȳ ⊥ and s̄ ∈ L̄ \ (x̄⊥ ∪ ȳ ⊥ ). Then at̄ + s̄ 6∈ x̄⊥ ∪ ȳ ⊥ for all a ∈ F. Therefore,

Q(at̄ + s̄) = a2 Q(t̄) + 2aB(t̄, s̄) + Q(s̄) = 0, for all a ∈ F.

Since p > 2, this implies that

Q(t̄) = B(t̄, s̄) = Q(s̄) = 0 ∀t̄ ∈ x̄⊥ ∩ ȳ ⊥ and ∀s̄ ∈ L̄ \ (x̄⊥ ∪ ȳ ⊥ ). (∗∗)

Note that w̄ ∈ x̄⊥ ∩ ȳ ⊥ and the set L̄ \ (x̄⊥ ∪ ȳ ⊥ ) spans the whole space L̄. Therefore,
B(w̄, L̄) = 0 and then x̄⊥ = ȳ ⊥ . Together with (∗∗) we deduce that Q(x̄⊥ ) = Q(L̄\ x̄⊥ ) = 0,
and hence Q(L̄) = 0 which is impossible.2

Theorem 11.4 Suppose that p > 2 and that L is a unimodular lattice. If M and N are
isometric primitive sublattices of L and τ : M → N is an isometry, then there exists
σ ∈ O(L) such that σ|M = τ . In particular, σ(M ) = N .
Proof. We proceed by an induction on the rank of M . If the rank of M is 1, then the
theorem is a consequence of Lemma 11.3. Suppose now that the rank of M is at least 2. If
there exists x ∈ M with Q(x) ∈ Z×
p , then

M = Zp x ⊥ M 0 , L = Zp x ⊥ L0 ,

where L0 is unimodular. By Lemma 11.3, there exists an isometry of L which sends x to


τ (x). So, we may assume that x = τ (x), and hence

N = Zp x ⊥ N 0 .

73
Both M 0 and N 0 are primitive sublattices of L0 . Moreover, by Theorem 6.4, M 0 and N 0
are isometric. By the induction assumption, there exists an isometry φ ∈ O(L0 ) such that
φ|M 0 = τ |M 0 . Then σ = 1 ⊥ φ is an isometry of L and φ|M = τ .
Now, suppose that Q(M ) ⊆ (p). Let {x1 , . . . , xn } be an orthogonal basis for M . Since
M is primitive in L, there exists z ∈ L such that B(z, x1 ) = 1 and B(z, xi ) = 0 for all i ≥ 2.
Then the binary lattice H = Zp x1 + Zp z is a hyperbolic plane. Therefore,

L = H ⊥ L1 ,

where L1 is unimodular. Note that the lattice M1 = Zp x2 + · · · + Zp xn is a primitive


sublattice of L1 . As is in the previous paragraph, we may assume that x1 = τ (x1 ). Then
N1 := Zp τ (x2 ) + · · · + Zp τ (xn ) is also a primitive sublattice of L1 , and N1 is isometric to
M1 . By Lemma 11.3, there exists an isometric ρ ∈ O(L1 ) such that ρ|M1 = τ |M1 . Then
σ = 1 ⊥ ρ, which is an isometry of L, is the desired isometry.2

Remark 11.5 When p = 2, Theorem 11.4 holds if L is an even unimodular Z2 -lattice.

11.2 Main Result


From now on, let L be a Z-lattice on a positive definite quadratic space V over Q. Let
Q(gen(L)) be the set of rational numbers represented by L. For any positive integer r, let

Qr (gen(L)) = {n ∈ Q(gen(L)) : ordp (n) ≤ r for all p for which Vp is anisotropic }

and
Qr (L) = Q(L) ∩ Qr (gen(L)).
If dim(V ) ≥ 5, then Q(gen(L)) = Qr (gen(L)) for every r, since Vp is isotropic for every
prime p. If dim(V ) = 4, the above definitions make sense because there are only finitely
many p for which Vp is anisotropic. We will prove the following theorem which has Theorem
11.1 as a consequence.

Theorem 11.6 Let L be a positive definite Z-lattice of rank ≥ 4. Then Qr (gen(L)) \ Qr (L)
is finite.

Lemma 11.7 Let a be a positive rational number and M be a positive definite Z-lattice of
rank ≥ 3. Suppose that the genus of M contains only one spinor genus. Let q be a prime
for which Mq is isotropic. Then there exists s ∈ Z such that

(a) Q(gen(q s M )) ⊆ Q(M );

(b) Qr (gen(q s M ⊥ hai)) \ Q(M ⊥ hai) is finite.

74
Proof. (a) By Theorem 8.4, Z[1/q]M ∼
= Z[1/q]K for every K ∈ spn+ (M ). Therefore, there
s
exists s ∈ Z such that q K ⊆ M and hence
1
Q(K) ⊆ Q(M ).
q 2s
This s can be chosen to be independent of K because there are only finitely many classes
in spn+ (L). Let n ∈ Q(gen(q s M )). Then q −2s n is represented by some K ∈ gen(M ) =
spn+ (M ). So n ∈ Q(M ).
(b) For our convenience, let L0 be the lattice q s M ⊥ hai. Let S be the set containing all the
prime divisors of 2d(L0 ). Note that L0` is unimodular for all ` 6∈ S. For each prime p, let

Qr (L0p ) = {t ∈ Q(L0p ) : ordp (t) ≤ r if L0p is anisotropic}.

Let p ∈ S. Suppose that t ∈ Qr (L0p ) and t = Q(v) for some v ∈ q s Mp . If t 6= 0, then for
a sufficiently large integer k, t − ap2k = t2 for some  ∈ Z× 2k
p . So, t = Q(v) + ap . If t = 0
and v is an isotropic vector in q Mp , then there exist integer m ≥ 0 and w ∈ q s Mp such
s

that Q(w) = −ap2m . So, t = Q(w) + ap2m . If t = 0 but q s Mp is anisotropic, then L0p must
be isotropic. So, there exists v 0 ∈ q s Mp and nonzero u ∈ Zp such that t = Q(v 0 ) + au2 .
In summary, for each t ∈ Qr (L0p ), there exists u ∈ Zp , u 6= 0, such that t ∈ Q(q s Mp ) +
au2 . Since the set Qr (L0p ) is compact and Q(q s Mp ) is open, there is a finite set Ip of nonzero
p-adic integers such that
[
Qr (L0p ) ⊆ (Q(q s Mp ) + au2p ).
up ∈Ip

Therefore, the product Y


Q(L0p )
p∈S

is contained in the union of the sets


Y
(Q(q s Mp ) + au2p )
p∈S

where up ∈ Ip for each p ∈ S. By the Chinese Remainder Theorem, we can find a finite set
of nonzero integers J such that for each collection {up : p ∈ S} there exists u ∈ J with
Y Y
(Q(q s Mp ) + au2p ) = (Q(q s Mp ) + au2 ).
p∈S p∈S

Suppose that t ∈ Q(gen(L0 )) and that t ≥ au2 for all u ∈ J. Then for each p ∈ S, there
exists u ∈ J such that
t − au2 ∈ Q(q s Mp ).

75
For each ` 6∈ S, t − au2 ∈ Z` since L0` is unimodular and u ∈ Z` . But q s M` is a unimodular
lattice of rank ≥ 3 and ` is odd. Therefore, Q(q s M` ) = Z` and hence t − au2 ∈ Q(q s M` )
for all ` 6∈ S. Consequently, it follows from part (a) that

t − au2 ∈ Q(gen(q s M )) ⊆ Q(M ).

As a result, t ∈ Q(M ⊥ hai).2


Proof of Theorem 11.6. Let L be a Z-lattice on a positive definite quadratic space V of
dimension at least 4 over Q. Choose a prime q - 2d(L) such that the Witt index of Vq is
at least 2. When dim(V ) ≥ 5, q can be chosen to be any prime not dividing 2d(L). When
dim(V ) = 4, Dirichlet’s theorem implies that there are infinitely many primes ` which are
congruent to 1 mod p3 for all p | 2d(L). By the Quadratic Reciprocity Law, d(L) is a square
mod `. We then choose q to be one of these `.
Let p ∈ S. For any α ∈ Q(Lp ), α ∈ Q(Zp u) for some u ∈ Lp with |Q(u)|p maximal.
Since Q(Lp ) is compact and each Q(Zp u) is open, there exists a finite subset Ip of Lp such
that [
Q(Lp ) = Q(Zp up ).
up ∈Ip

Let {up ∈ Ip : p ∈ S} be given. Choose t ∈ Q with the following properties:

(1) t ∈ Z×2
p Q(up ) for all p ∈ S;

(2) the prime factors of the denominator of t are in S;

(3) if p 6∈ S, then t is either a prime or a unit in Zp .

We can construct such a t as follows. Let Q(up ) = pap p , where p ∈ Z×


p . By the Chinese
Remainder Theorem, there exists an integer k such that
Y
k `a` ≡ p mod p3 , for all p ∈ S.
`∈S\p

By Dirichlet’s theorem there exists a prime p̃ 6∈ S such that


Y
p̃ ≡ k mod `3
`∈S

Then Y
p̃ `a` ∈ pap p Z×2 ×2
p = Q(up )Zp .
`∈S

We then set Y
t = p̃ `a`
`∈S

and we are done.

76
From (1), it is clear that t ∈ Q(Lp ) for all p ∈ S. For any p 6∈ S, p is odd and Lp
is unimodular; hence t ∈ Q(Lp ). Consequently, t ∈ Q(gen(L)). It follows form Theorem
7.21 that t ∈ Q(spn+ (L)). In other words, there exists K ∈ spn+ (L) such that t ∈ Q(K).
Let ` be a prime not in S such that ` - t and the Witt index of V` is at least 2. Then
Z[1/`]K ∼ = Z[1/`]L by Theorem 8.4, and hence there exists v ∈ V such that Q(v) = t and
v ∈ Lp for all p 6= `.
Choose an integer e ≥ 0 such that `e v is a primitive vector in L. We shall construct a
sublattice M of L of rank rank(L) − 1. For the sake of convenience, let J be the rank one
lattice spanned by `e v. We construct M by specifying its localization Mp for each prime p.
Let W be the orthogonal complement of QJ in V .
If p 6∈ S, p 6= ` and p - t: since Jp is unimodular, it splits Lp . We let M (p) be the orthogonal
complement of Jp in Lp . Then θ(O+ (M (p))) ⊇ Z× s
p and Q(q M (p) ⊥ Lp ) = Zp because
q s M (p) is unimodular of rank ≥ 3.
If p ∈ S: let L0 (p) be a maximal lattice which contains Lp ∩Wp . Then pa L0 (p) ⊆ Lp ∩Wp for
some a ≥ 0. Let M (p) be the Zp -lattice pa L0 (p). Since M (p) is maximal, θ(O+ (M (p))) ⊇ Z×
p
by Proposition 11.2. Also, Q(q s M (p) ⊥ Jp ) ⊇ Q(Jp ) = Q(Zp up ).
If p = `: since the Witt index of V` is at least 2 and L` is unimodular, L` = H1 ⊥ H2 ⊥ K,
where H1 and H2 are hyperbolic planes and K is a unimodular Z` -lattice. Also, since `e v
is primitive in L` , there exists z ∈ L` such that B(z, `e v) = 1. Let G be the binary Z`
lattice spanned by z and `e v. If Q(`e v) is not a unit, then G is a hyperbolic plane and
hence L` = G ⊥ H with H ∼ = H2 ⊥ K. In this case, we let M (`) = I ⊥ H where I is the
orthogonal complement of J` in G.
If Q(`e v) is a unit, then J` splits L` . Therefore, L` = J` ⊥ E = H1 ⊥ H2 ⊥ K, where
E is a unimodular Z` -lattice. Let w be a primitive vector of H1 such that Q(w) = Q(`e v).
By Lemma 11.3, there exists an isometry φ ∈ O(L` ) such that φ(w) = `e v. Hence

E∼
= H2 ⊥ K ⊥ ( orthogonal complement of w in H1 ).

In this case, let M (`) be E.


In any case, θ(O+ (M (`))) ⊇ Zp since M (`) has a rank 2 unimodular component. Also,
Q(q s M (`)) ⊇ Q(q s M (`)) = Z` .
If p 6∈ S and p | t: by the construction of t, we see that p2 - t. Therefore, v is primitive
in Lp . Let z ∈ Lp such that B(z, v) = 1. Then the binary sublattice spanned by v and z,
denoted H, is a hyperbolic plane. So, Lp = H ⊥ G for some unimodular Zp -lattice G. Also,
H has a vector w satisfying B(w, v) = 0 and Q(w) = −Q(v) = −t. Let M (p) be Zp w ⊥ G.
Then θ(O+ (M (p))) ⊇ Z× s
p and Q(q M (p) ⊥ Jp ) = Zp .

Note that or almost all primes p, M (p) is the orthogonal complement of Jp in Lp . By


Theorem 7.8, there exists a Z-lattice M on W such that Mp = M (p) for all p. Since
θ(O+ (M (p))) ⊇ Z×
p for all p, the genus of M contains only one spinor genus. Note that all
together we have constructed only finitely many t and M .

77
Let α ∈ Qr (gen(L)). Then at each p ∈ S, α ∈ Q(Zp up ) for some up ∈ Jp . For p 6∈ S,
Lp is unimodular; hence α ∈ Zp . Let t, J and M be as constructed as above. Then
α ∈ Qr (gen(q s M ⊥ J)). By Lemma 11.6, all but finitely many α are in Q(L).2

12 Classes and Genera of Representations


Throughout this section, R is either Z or Zp for some prime p, and F is the field of fractions
of R. An R-lattice L is said to be integral if s(L) ⊆ R.

12.1 Representations of Lattices


Definition 12.1 Let L and M be R-lattices. A representation of L by M is an injective
R-homomorphism σ : L → M such that Q(φ(v)) = Q(v) for all v ∈ L.
This generalizes the representation of an element in R by a lattice. For, let a ∈ R, L be
the rank 1 R-lattice hai, and M be an R-lattice. Then having a representation of L by M
is equivalent to the existence of a vector x ∈ M with Q(x) = a.
Definition 12.2 Two representations φi : L → Mi (i = 1, 2) of R-lattices are in the same
class if there exists an isometry σ : M1 → M2 such that φ2 = σφ1 .
Proposition 12.3 Suppose that L and M are nondegenerate R-lattices. If φ : L → M is
a representation, then there exists an R-lattice M 0 such that L ⊆ M 0 and φ is in the same
class of the inclusion L ,→ M 0 .
Proof. Let U be a quadratic space which is isometric to φ(F L)⊥ , and let τ : U → φ(F L)⊥
be an isometry. Then φ ⊥ τ is an isometry from F L ⊥ U to F M . Let M 0 be (φ ⊥ τ )−1 (M ).
Then L ⊆ M 0 and φ is in the same class of the inclusion L ,→ M 0 . 2
Lemma 12.4 Suppose that L and M are unimodular Zp -lattices, p > 2. Then there is only
at most one class of representations of L by M .
Proof. Let φi : L → M , i = 1, 2, be two representations of L by M . Then φ1 (L) and φ2 (L)
are two primitive sublattices of M . Let τ : φ1 (L) → φ2 (L) be an isometry. By Theorem
11.4, there exists an isometry σ ∈ O(M ) such that σ|φ1 (L) = τ . Then σφ1 = φ2 , hence φ1
and φ2 are in the same class.2
Lemma 12.5 Suppose that L and M are nondegenerate R-lattices with L ⊆ M . Let N
be the orthogonal complement of L in M . Then d(N ) | d(M )d(L). In fact, N # /N is a
homomorphic image of a subgroup of M # /M ⊕ L# /L.
Proof. This is just Lemma 7.2. Since M/N is torsion free, N is a primitive sublattice of
M . So, the restriction map induces a surjective homomorphism
M # /(L ⊥ N ) → N # /N.
On the other hand, the map M # /(L ⊥ N ) −→ L# /L ⊕ M # /M which sends x to (P (x), x)
where P is the orthogonal projection from F M onto F L, is injective.2

78
Theorem 12.6 Suppose that L is a nondegenerate R-lattice. Given any positive integer n
and element d ∈ R \ {0}, there are only finitely many classes of representations φ : L → M
with rank(M ) = n and d(M ) = d.
Proof. Let φ : L → M be one of such representations. If N is the orthogonal complement
of φ(L) in M , then
φ(L) ⊥ N ⊆ M ⊆ M # ⊆ (φ(L) ⊥ N )# .
By Lemma 12.5, d(N ) divides d(M )d(L). Therefore, there are only finitely many isometry
classes of N . Let {Ni } be a complete set of representatives of these classes. For each i,
there are only finitely many integral R-lattices Mij with rank(Mij ) = n, d(Mij ) = d, and

L ⊥ Ni ⊆ Mij ⊆ (L ⊥ Ni )# .

Since there are only finitely many Ni , the set {Mij } is finite. Let φij : L → Mij be the
representation induced by the inclusion L ⊥ Ni ,→ Mij . We claim that every representation
of L by some lattice M with rank(M ) = n and d(M ) = d is in the class of some φij .
Let ρ : L → M be such a representation. Let N be the orthogonal complement of ρ(L)
in M . Then N must be isometric to Ni for some i. Let τi : N → Ni be an isometry, and σ
be the isometry φ−1 ⊥ τi from φ(L) ⊥ N to L ⊥ Ni . Then σ((φ(L) ⊥ N )# ) = (L ⊥ Ni )# ,
and hence σ(M ) = Mij for some j. It is immediate that σρ = φij .2
The above theorem is false if d(L) = 0. Consider a Z-lattice L = Zx ∼ = h0i, and a
hyperbolic plane M with a basis {e, f } whose associated symmetric matrix is ( 01 10 ). Then
x 7→ nf , n ∈ Z \ {0}, are all different representations. The orthogonal group of M is clearly
finite because {±e, ±f } are the only primitive isotropic vectors in M and any isometry of
L must send e (or f ) to a primitive isotropic vector.

Lemma 12.7 Let M1 and M2 be two R-lattices on a quadratic space over F . Then [O(Mi ) :
O(M1 ) ∩ O(M2 )] is finite for i = 1, 2.
Proof. There exist two nonzero elements a and b in F such that bM1 ⊆ M2 ⊆ aM1 . We
have a natural homomorphism

O(M1 ) → AutZ (aM1 /bM1 ).

Let G be the kernel of this homomorphism. Then [O(M1 ) : G] is finite since aM1 /bM1 is a
finite group. The elements of G acts trivially on M2 /bM1 , hence G ⊆ O(M2 ). as a result,
G ⊆ O(M1 ) ⊆ O(M2 ) and thus [O(M1 ) : O(M1 ) ∩ O(M2 )], which is small than [O(M1 ) : G],
is finite. Similarly, [O(M2 ) : O(M1 ) ∩ O(M2 )] is also finite.2

Theorem 12.8 Let L be a nondegenerate Z-lattice. If M is a positive definite Z-lattice or


QM is the hyperbolic plane, then there are only finitely many representations of L by M .
Proof. Since there are only finitely many classes of representations of L by M , it suffices to
show that O(M ) is finite when M is positive definite or QM is the hyperbolic plane.

79
Suppose that M is positive definite. Let {e1 , . . . , en } be an orthogonal basis for QM
with Q(ei ) ≤ Q(ei+1 ) for all i, and let M 0 = Ze1 ⊥ · · · ⊥ Zen . Then |O(M 0 )| is finite. By
Lemma 12.7, |O(M )| is finite.
If QM is the hyperbolic plane, let M 0 be a lattice on QM with symmetric matrix ( 01 10 ).
Then |O(M 0 )|, and hence |O(M )| as well, is finite.2

Remark 12.9 It can be shown that if M is an indefinite Z-lattice and QM is not the
hyperbolic plane, then O(M ) is an infinite group.

12.2 Genus of Representations


Definition 12.10 Two representations φ : L → M and φ0 : L → M 0 of Z-lattices are in
the same genus if there are isometries σp : Mp → Mp0 for all primes p such that φ0 = σp φ for
all p.
For a prime p and a pair of Z-lattices L and M , let c(Lp , Mp ) be the number of classes
of representations of Lp by Mp . By Theorem 11.4, c(Lp , Mp ) is equal to 1 for almost all p.

Theorem 12.11 Let L and M be nondegenerate Z-lattices.


Q The number of genera of rep-
resentations of L by lattices in gen(M ) is equal to p c(Lp , Mp ).
Proof. Let G be the set of all genera of representations of L by lattices in gen(M ). For each p,
let Cp be the set of classes of representations of Lp by Mp . Let φ : L → M 0 be a representation
with M 0 ∈ gen(M ). For each prime p, φ induces representations φ : Lp → Mp0 . Let
σp : Mp0 → Mp be an isometry. Then σp φ : Lp → Mp is a representation. Different choices
of the isometry σp result in different representations of Lp by Mp that are in the same class.
So we have a well-defined map Y
Φ : G −→ Cp .
p

Let φ : L → K and τ : L → N be two representations of L by lattices in gen(M ).


Suppose that for each p, the classes of representations of Lp by Mp induced by φ and τ are
the same. Then for each p, there are isometries σp , ψp and ρp which make the following
diagram commute
φ σp
Lp −−−−→ Kp −−−−→ Mp

ψ
y p
τ ρp
Lp −−−−→ Np −−−−→ Mp
Let ηp = ρ−1
p ψp σp . Then τ = ηp φ. Therefore, φ and τ are in the same genus. This shows
that Φ is injective.
Suppose that a representation φp : Lp → Mp is given for each p. Let T be a finite set of
primes satisfying

(a) c(Lp , Mp ) = 1 for all p 6∈ T ;

80
(b) Lp ⊆ Mp for all p 6∈ T .

Let W and V be the Q-spaces spanned by L and M , respectively. By the Hasse Principle,
there is an isometry sending W into V . So we may assume that L is a Z-lattice in V . For
each p, it follows from Witt’s extension theorem that there exists an isometry σp ∈ O(Vp )
such that σp |Lp = φp . Define a Z-lattice M 0 on V by specifying its localization as follows:

Mp for all p 6∈ T ;
Mp0 =
σp−1 (Mp ) for all p ∈ T .

Then at each p ∈ T ,
Lp = σp−1 (φp (Lp )) ⊆ Mp0 .
Note that for each p 6∈ T , Lp ⊆ Mp = Mp0 . Therefore, L ⊆ M 0 . Let j : L ,→ M 0 be the
inclusion. Then for each p 6∈ T , φp and j are in the same class because c(Lp , Mp ) = 1. For
each p ∈ T , σp (φp (x)) = x for all x ∈ Lp ; thus j = σp−1 φp for each p ∈ T . Consequently, j
and φp are in the same class for all p. This proves that Φ is surjective.2

81

You might also like