This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
represents 1 whenever oc, fi EF. Hence every quaternion algebra is isomorphic to (1, —1) by Proposition 57:9. So there is essentially one quaternion algebra over a finite field F, namely
62:4. Remark. The factor groups n
,
of a regular n-ary quadratic space V (n 2) over a finite field are described by the equations (On+ :07,1 ) , 2, On = and + v} if dV = 1 with n even
nZn -
I.iv otherwise.
§ 63. Local fields Now consider a local field F. Let us recall some of the basic definitions and notation of §§ 13, 16 and 32. F is a field with a complete and discrete prime spot p and the residue class fieldF (p) is a finite field of Np elements. We let stand for the ring of integers (p), u for the group of units u (p), p for the maximal ideal m (p), n for a prime element of F at p, ord for the order function ordr , and I I for the normalized valuation We know from § 22E that the fractional ideals of F at p are the powers
pv= nvo
(v EZ) .
Remember that in this part of the book we are making the general assumption that the characteristic of F is not 2. However it is still possible for the residue class field to have characteristic 2. This is what happens for instance in the case of the 2-adic numbers. We shall call F a dyadic local field if its residue class field has characteristic 2; thus for a dyadic field we have z (1 ) = 0 and z(F (p)) 2. We call F non-dyadic if it is not dyadic; here we have (F (p)) >2 and (F) 0 or z (F) (p)) . This was originally proved by DICKSON and generalized by C. CHEVALLEY, Abh. Math. Sm. Hamburg (1935), pp. 73-75, to forms of any degree: every form of degree cl in cl 1- 1 variables over a finite field has Cl non-trivial zero.
Chapter VI. The Equivalence of Quadratic Forms
159
Note that F is dyadic if and only if
0 < 121 < 1
(or 0 < ord2 < 00) ;
it is non-dyadic if and only if
121 = 1
(or ord2 = 0) .
We saw in § 62 that exactly half the non-zero elements of a finite field of characteristic not 2 are squares; in particular this is true of the residue class field F (p) of a non-dyadic local field. On the other hand if F is a dyadic local field, then its residue class field is a finite field of characteristic 2; since all finite fields are perfect we can conclude that every element of F (p) is a square, i. e. that
(F (p)) 2 = F (p)
if p dyadic.
This has the following important consequence in F: if e, El are units in a dyadic local field, there is a unit 6 such that
s' 862 modn. In particular, in the dyadic case every unit is a square mod p. § 63A. Quadratic defect 63:1. Local Square Theorem. Let a be an integer in the local field F. Then there is an integer 13 such that
1 ± 4n a
(1 + 2n IV. Proof. The polynomial nx 2 + x — oc is reducible by the Reducibility Criterion of Proposition 13:9. Hence we have 13, IT EF such that
(x — 13) (x
j9').
Then the product of the roots is equal to —'oc, hence one of the roots, say j9 must be in e. But 1 ± 11n-2 + 4 .7-1 a) = __(_..,27---1 ,
2
by the quadratic formula. Hence
1 + 47t — (1 + 27c 13) 2 . q. e. d. 63: l a. Corollary. Suppose e, 6 are units in F such that e 6 mod 4.7. Then E E ô u2. 63: 1 b. Corollary. I":-2 is an open subset of F. Consider any element of the local field F. Then 4: has at least one oc in F. We write expression in the form = n 2+ oc with ?I ,
= n2 +
(n,
F)
160
Part Three. Arithmetic Theory of Quadratic Forms over Fields
in all possible ways and take the intersection
Then b (e) is either a fractional ideal or O. We call b (e) the quadratic defect of e. Clearly
b • If e is a square in F, then b (e) = O. Using the Local Square Theorem one can easily show that the converse is also true. Hence
E F2 44- b = O. In particular e always has an expression
= 772+ a with ao = b (e) . From this it follows that
b (a2 e) = a2 b (e)
V a, e E F.
We have
b (e) = eo if ordo e is odd. When ordo e is even we can write = n2re with s a unit, and then b (e) 7C2r b (e). So it is enough to study the quadratic defect on the group of units u. For a unit s we can write s 62 + a with 6 Eu and
ao = b(e) S 0. What is the intuitive meaning of the quadratic defect ? We have just seen that having defect 0 is equivalent to being a square. Consider a non-square unit s with defect b (s) = p4 Ç o. Then we can write s = 62 + a with 6 E u, a E pd, while such an expression is impossible with an a in p4 +'. So here the quadratic defect is that ideal pd with the property that 8 is congruent to a square modpd but not modpd +1.
63:2. Let e be a unit in the local field F. I/ F is non-dyadic, then b(s) is 0 or 0; if F is dyadic, then b(s) is one of the ideals 0C40C4p-1 C4p -3 (-••CP3 CP Proof. 1) If F is non-dyadic, then it follows from the Local Square Theorem that b (s) is 0 whenever it is not o, j. e. whenever 1)(0 p. 2) Now consider the case of a dyadic field F . If b (s) C 4o then b(s) ç 4p and so E is a square by the Local Square Theorem, hence b (s) = O. It remains for us to consider an s with b (s) = peg and 4 0 c pd o, and to prove that d must then be odd. Suppose if possible that we have such an s with an even d. Put d 2r. We can write s = 62 + a with 6 E u and ao — pd. Replacing s by 462 gives us a new s of the form
s
1 + sigr2r
Chapter VI. The Equivalence of Quadratic Forms
161
with si E u, b (s) = Or, and 2o C pr Ç o. By the perfectness of the residue class field there is a unit 61 such that O.= e modn. Then 1 H._ Eiger 1 + *or (1 + th ar)2 modulo
ger+ 1 •
Hence we have an expression 8 = (1 ± (51 7e)2+ al with al E p2r +1 .
This contradicts the fact that b (s) = p2r. Hence d must be odd. q. e. d. 63:3. Let s be a unit in the local field F. Then b(s) = 40 if and only if F is quadratic unramified. Proof. Here it is understood that F (VW) is provided with that unique spot which divides the given spot p on F. By Proposition 32:3, F(re-)
orivF
is also a local field. 1) First let us be given b (s) = 40. This certainly makes the given extension quadratic. Multiplying s by the square of a unit in F allows us to assume that s = 1 mod4, hence that —41 (s 1) E o. Now F can
(fr)
be obtained by adjoining a root a = ,c2 + x
(-1 + fr ) of the polynomial
—4- (1 — E)
to F. But Proposition 32:6 applies to this situation. Hence F (r)/F is unramified.
2) Now suppose F (11i)IF is quadratic unrarnified. We can assume that s is given in the form s = 1 A- oc with b (s) = oco. Since the given extension is quadratic we know that 4o S ao S o. This finishes the proof for the non-dyadic case. Now assume that F is dyadic and prove that oco = 40. Write A = —1+ I/E—. Then A (A -1- 2) = oc. Let 1 1 be the prolongation of the given valuation on F to F (fr). If we had 1A1 > 121 we would have loc1 = 1Al 2 > 141 ,
hence a would have even order in F (VW), hence it would have even order in F since the given extension is unrarnified; so b (s) would be equal to p2r with p2r 4o and this is impossible by Proposition 63:2. Hence 1A1 121. This implies that lai 141• Hence cco Ç 4o and so oco = 40 as required. q. e. d. 63:4. There is a unit s in the local field F with b (s) = 4o. If s' is any other such unit, then s' E s u2. Proof. By Proposition 32:9 there is an unramified quadratic extension EIF. Since x(F)--k 2 we can obtain E by adjoining a square root to F; O'Meara, lotroductigm to quadratic: fonts
11
162
Part Three. Arithmetic Theory of Quadratic Forms over Fields
since ELF is unramified this will have to be a square root of an element of even order. Hence we can write E = F (VW) with s a unit in F. Then b (s) = 4o by Proposition 63:3. Now consider e'. Then E' = F (V7) is quadratic unramified over F by Proposition 63:3. By Proposition 32:10 the two fields E and E' are splitting fields of the same polynomial over F, hence yir E F (ri), hence 1117E VF. So E' E 80. q. e. d. 63 : 5. Let E = 1 ± a be a unit in a dyadic field with 141 < ial <1 and orda odd. Then b(e) = ao. Proof. Put ow = pd. Clearly b (s) S pd. Suppose if possible that b (e) ç p4 + 1. Then there is a y Eo such that 1
oc (1 ± y) 2 modad+ 1 .
So 1 y (y 2)1 = loci by the Principle of Domination since Iciel = 1701. 141, contrary to the assumpIf we had I Y I 5- 121 we would have loci tions; if we had ly1 > 121 we would have loci = 1y1 2, contrary to the assumptions. Hence we cannot have b (s) S pd +1. So b (s) is indeed equal to ao. q. e. d. 63:6. Remark. Each of the ideals in Proposition 63:2 will actually appear as the quadratic defect of some unit e. To get defect 0 take E = 1, to get 4o apply Proposition 63:4, to get pd with d odd and 40 c p 4 Ç o take e - -- 1 + ad and apply Proposition 63:5. We conclude this subparagraph with local index computations that will be needed later in the global theory. For any fractional ideal pr with r> 0 the set 1 pr= {1 + curia E 0} is a neighborhood of the identity 1 under the p-adic topology on F. Clearly 1 ± pr is a subgroup of the group of units u and we have 1 pr+ic 1 + prs u
(if r > 0) .
63:7. Lemma. Let 0 be a homomorphism or a commutative group G into some other group, let 0 G be the image of G and Go the kernel of 0. Then for any subgroup H of G we have (G: H) = (0G: 0 H) (Go : H 0) , where the left hand side is finite if and only if the right hand side is.
Proof. By the isomorphism theorems of group theory we obtain (G : H) (G : G H) (G 0 H : H)
(0 G:0(G 0 11))(G 0 :G 0 nH) = (0G: OH) (G o : H0). q. e. d.
163
Chapter VI. The Equivalence of Quadratic Forms
63:8. F is a kcal field at p, u is the group of units, and 1 + pr is a neighborhood of the identity with r> O. Then (1) (u : 1 -I- pr) = (N p — 1) (N p)r -1 ,
(2) (1 + pr) 2 = 1 + 2pr if prg, 2p . Proof. (1) Consider a residue class field F) of F at p. The restriction of the bar map is a multiplicative homomorphism of u onto the non-zero elements of F with kernel 1 + p. Hence (u: 1 + p) = Np — 1. Now the mapping yo(1 + = a. is easily seen to be a homomorphism of the multiplicative group 1 + pr onto the additive group F with kernel 1 + pr+ 1. Hence (1 + pr: 1 + pr+ 1) = Np. By taking the tower u 21+pDl+p 2 )•--D1+pr
we obtain (u : 1 + pr) (Np 1) (Np)* --- 1 . S 1 + 2pr when pr Ç 2p. We must reverse the (2) Clearly (1 + inclusion. So consider a typical element 1 + 2a rcr of 1 + 2pr with a E 0.
Then
+ 2 OE nr_, (1 + Goe )2 _ oc znar, + aye? (1 + /3 701 fl7eir E -t- p 2 whenever for some /3 E o. it is enough to prove that 1 fl Et' and prS 2 1'. By the Local Square Theorem there is a y E 21, such that (1 -I- y)2 .--- 1 + 9 7r2r. Then 17 + 21 = 121 and so
12 v1= 1Y(7 + 2)1= 113701 5- 127Er+ 1 hence y E pr + 1 Ç. Pr 63: 9. F is a local field at p and u is the group of units. Then
q. e. d.
(P: P2) = 2 (u : u2) = 4 (NWT".
Proof. 1) To prove the first equality apply Lemma 63:7 with G = H and Oa = lal. Then
(P :P2) = (1P1 1P1 2) (u : u2) = 2 (u : u2) . 2) To prove the second equality apply Lemma 63:7 and Proposition 63:8. This time take G = u, H = 1 + pi' for any r > 0 such that prç_ 2p, and 0 a = O. Then (u : 1 + pr) = 2(u 2 : (1 + p1 2) = 2 (u2 : 1 + 2 pr) = 2 (u2 : 1 p r +ord
Hence (u : u2) (u : 1 + pr)
2 (u : 1 + pr -Foed 2) .
Hence (u : u2) -,--- 2 (Np)°rd 2 q. e. d. 11*
164
Part Three. Arithmetic Theory of Quadratic Forms over Fields
§ 63B. The Hilbert symbol and the Hasse symbol In this subparagraph F can either be the local field under discussion or any complete archiMedean field. So F is either a local field at p, or p is real and complete, or p is complex and complete. In any one of these situations it is possible to replace the Hasse algebra by a simpler invariant called the Hasse symbol. In the definition of the Hasse symbol the quaternion algebras are replaced by Hilbert symbols which we now define. Given non-zero scalars a, ft in an arithmetic field of the above type the Hilbert symbol t a, 13\
k
P
or simply (a, 16), is defined to be + 1 if ce 4.2 + /3772. 1 has a solution E F; otherwise the symbol is defined to be — 1. 63:10. Example. Put E=F(110 So EIF is of degree 1 or 2. Then a E NE/FE if and only if (
P ) —1 .
Our first results refer to the local case only. 63:11. Let V be a binary quadratic space over a local field F and let the discriminant dV be a prime element of F. Let ZI denote a fixed unit of quadratic defect 4o. If y is any non-zero scalar, then V represents y or y.Z1 but not both. Proof. By scaling V we can assume that y = 1. 1) Our first task is to prove that V represents either 1 or 4. Since cl V is a prime element there is a splitting V <e>1 <671> in which e and 6 are units in F. If F is non-dyadic, e will either be a square or a square times 4 by Propositions 63:2 and 63:4. We may therefore restrict ourselves to the dyadic case. The above unit e can actually be any unit represented by V. In fact let it be a unit of smallest quadratic defect in Q (V). We can assume that this E has the form E 1 ± fi with b (6) = flo c o. Thus
V rz- <1 16> 1 <67c> One of three things can happen. (i) If b (e) = 0 we have 16' = 0 and so 6 = 1 and V represents 1 as desired. (ii) If b (e) = 4o then V represents a unit of quadratic defect 4o and so it represents 4 by Proposition 63:4. (iii) The one remaining possibility is 4o C b (e) C o. Let us prove that this is impossible. If 4o C fi'D C o we write 19 = el nk with el a unit and b pk Here k is odd. By the perfectness of the residue class field we can choose a unit A such that 2 2 6 = — e1 modn. Then V represents 1 ± egr k 2267.Ek 1
mode +1
Chapter VI. The Equivalence of Quadratic Forms
165
since k is odd. In other words V represents a unit whose quadratic defect is contained in plc +1 • This is contrary to the choice of e. 2) Finally we have to prove that V cannot represent both 1 and Ll. If it did we would have <1>1 <87c> 1
I.
and
Proof. The first equation is a direct consequence of the proposition. Let us do the second. In the non-dyadic case we can find n in o such
that
e 2 +4 77 2= 1 modn since a binary quadratic form over a finite field of characteristic not 2 is universal by Proposition 62:1. Here has to be a unit since A is a nonsquare, hence we have A E o such that ep(1 A n) ± A lp = 1 Then 1 ± An is a square by the Local Square Theorem. Hence <e> (e, ) — 1. In the dyadic case use the fact < > represents 1, hence that represents 1 modulo 4e together with the perfectness of the residue class field to show that <e> I represents 1 modulo 4p. q. e. d. Then apply the Local Square Theorem. 63:11b. The quaternion algebra (n, A) is a division algebra. All quaternion division algebras over F are isomorphic to it. Proof. The quadratic space <7i> I does not represent 1, by the proposition. Hence (n, A) is a division algebra by Proposition 57:9. It is easily seen that every quaternion algebra over F is similar to a tensor product of quaternion algebras of the form (e, 8n) where e and 8 are units. It therefore suffices to prove that this quaternion algebra is isomorphic to (n , Z1 0) where A 0 is one of the elements 1 and A. But
<8n> ± ce8nA 0 > . Computing Hasse algebras gives (e, 8n) (4 0, A 0 E 6)0 (A 0, n) . <e>
Part Three. Arithmetic Theory of Quadratic Forms over Fields
166
But (A 0 , A 0 e 6) 1 by Corollary 63:11a and Proposition 57:9. Hence q. e. d. (e, 6n) (40, a). 63:12. Example. Corollary 63: I I a shows the very simple nature of the Hilbert symbol over non-dyadic fields. For let e, 6 be units in a given ;16 ) — 1 if and — 1 always. And (-non-dyadic local field. Then ( only if 6 is a square. Now consider the local field or real complete field F. There are essentially two quaternion algebras over F. Hence œb Pi ) ", 1
tg) igign
if and only if the number of division algebras appearing in this tensor product is even. But ( /4 ) is a division algebra if and only if (") 1. Hence the given condition is equivalent to —
(ez b p i)
_ 1.
isin So we have 611
k 1Ztign
k F
if and only if (
°Lb
Pil
(
6
i)
iSigm‘
These results are of course trivial when F is a complex complete field. We can carry over to the Hilbert symbol the various formulas of Proposition 57:10. The first two of these formulas are trivial for the Hilbert symbol, the third now reads ice,a13\ . tœ,—P)
and the fourth,
k P/
k
P
fcc , P1(cc,31_(oc,PY) \PMP/ 63:13. Let F be either a local field or a complete archimedean field at
p.
Given any non-square /3 in P there is an a in F such that (a1P) --
1.
Proof. The complex case cannot arise. In the real case take a= — 1. Now the local case. We can suppose that /3 is either a prime element or a unit of the form /3 = 1 ± y with 10) = ye. Let LI 1 mod4 be a.unit of quadratic defect 4e. If /3 is a prime element take a . J. If ye — 4e
Chapter VI. The Equivalence of Quadratic Forms
take a =
7C. If
167
40C yo Co take a = — p; then
± <—cc 13> is isotropic at all spots except q where it is not. Clearly V ± <— > is isotropic at q where q is real; by Proposition 63:17 it is also isotropic for a discrete q since the discriminant X is then a non-square at q. Hence V .1 < X> is isotropic by Theorem 66:1. So V represents X. So V
NELFJE
E (d)' NEIFJE Ç
PFNEIFJE
Chapter VII. Hilbert's Reciprocity Law
203
and we have established our claim that
c p,NEI,JE On the other hand EIF is unramified of local degree 2 at q since is a unit of quadratic defect 404 at q. This is impossible by Proposition 71:17. q. e. d. 71: 19. Theorem. Let T be a set consisting of an even number of discrete or real spots on an algebraic number field F. Then there are a, p in 1 such that i ct,p\ J-1 if PE T kP) 1 1 if pED— T. -
Proof. fi can be any element of F which is a non-square at all spots in T. Such an element always exists: for instance the Weak Approximation Theorem provides a /3 which is a prime element at the discrete spots in T and a negative number at the real spots in T. Put E = F (11-g) and define a group homomorphism (± 1) by the formula
where i = (ip) pEs2 denotes a typical idèle in Ir. Then NELFJE is in the kernel of by § 65A; and Pip is in the kernel of (7) by the Hilbert Reciprocity Law; hence PFNE/FJE is in the kernel of Ti; now (p is surjective by Proposition 63:13, and (JF:PFNEIFJE). 2 by Proposition 65:21, hence PENE/FIE is precisely the kernel of 9). Take an idèle j E IF which is a local norm at all pED—T and is not a local norm at any p E T. Then 9? (j) = 1 since T contains an even number of elements, hence j E PpNE/Ej.E. But this relation can also be read as follows: there is an a in F which is a local norm at all p in D — T, and at no p in T. This a gives the desired values to the Hilbert symbol. q. e. d. 7 1:19a. Corollary. /3 can be any element of F which is a non-square at all spots in T. § 72. Existence of forms with prescribed local behavior 72:1. Theorem. A regular n-ary space Up is given over each completion F an algebraic number field F. In order that there exist an n-ary space V Up for all p, it is necessary and sufficient that over F such that V (1) there be a do in F with dUp . do for all p, (2) Sp Ur . 1 for almost all p, (3) // Sp Up . 1.
PEI)
204
Part Three. Arithmetic Theory of Quadratic Forms over Fields
Proof. 1) Necessity. To obtain the first condition take do = dV. To obtain the second and third conditions consider a splitting Va4
Then Sp Up = Sp V p is a product of Hilbert symbols of the form
(oei E Pp)
at each p in T. Use the Weak Approximation Theorem to find an ai in F such that loci — ce:l p is small for all p in T. Since .F';; is an open subset of F1, we can obtain
YpET,
arEcF
provided our approximations are good enough. Do all this for 1 I Then take a quadratic space W over F with < oci >
•
n-1.
< oin—i>
Clearly Wp YpET.
Hence Sp Wp = Sp Up for all p in T. Let R denote the set of spots which Sp Wp and Sp Up are different. Of course R is a finite subset D — T. And Wp -- Up for all p in D — R by Theorem 63:20. If R T empty we are through. Otherwise R consists of those spots in D which S4, W1,= 1; now
./iS p W1, 11 sr wp
at
of is at
•H sp up = Hsp wp = 1 • 12-R
Hence R consists of an even number of spots in D — T. 3) We claim that there are quadratic planes P and P' with the same discriminant over F, with P PI; for all p in D R, and with Pp for all p in R. In fact, take any 0 in F which is a square at all archimedean spots and a non-square at all spots in R. By Theorem 71:19 and its corollary we can find an a in F with f —1 11 1
if PER — R. if p
Define
P <1> < 13> , P' ± <— cc 18> .
205
Chapter WI. Hilbert's Reciprocity Law
Then Pp and .13 , are isometric at all archimedean spots by choice of fl; applying Theorem 63: 20 at the discrete spots shows that .131„- P; for all p in D — R, and Pp P; for all p in R. So we have established our claim. 4) Consider Wp at any p in R. This Wp cannot be a hyperbolic plane since Wp and Up are non-isometric spaces with the same discriminant. We are also assuming that dim Wp 2. Hence
—>— (P _L
Vp ER
by Theorem 63:21. At any p in D R we also have such a representation since then P; P , by step 3). Hence there is a representation P'—>— (Pi_ W) by Theorem 66:3. Hence there is a quadratic space V over F with
This V has discriminant do = d W since ci P' = d P. For each p in D — R we have Pp s-:-. .13 , by choice of P, P', hence by Witt's theorem Up Vp ED — R . If p is in R we have Vp Wp sin.ce 134, Pi,; but Up * WI, by definition
of R; and dIgr = dUR = do ; hence by Theorem 63:20 we must have
Sp Vp — SW= SU R ; so Vp L-2_- Up for all p in R. Therefore Vp z'__ Up for all p and the space V has the desired property. q. e. d.
§ 73. The quadratic reciprocity law We conclude this chapter by finding an expression for the Hilbert symbol in terms of the Legendre symbol over the field of rational numbers Q. Recall the definition of the Legendre symbol: if p is an odd prime number, and if a is any rational integer that is not divisible by p, then
the Legendre symbol (f) is defined to be +1
or —1
according as a is or is not congruent to the square of a rational integer modulo p. In other words, (±i-) is 1 if the natural image of
z/p z is a square, otherwise
a in the finite field
(1,-) 6 is — 1. Now for any finite field K of
206
Part Three. Arithmetic Theory of Quadratic Forms over Fields
characteristic not 2 we have (K :K2) = 2 by § 62. Hence fa\ Ib\_lab\
UT)
lT)
Note that the Local Square Theorem tells us that
a (—
) --= 1 if and only
if a is a square in the field of p-adic numbers Q. b We shall use ( ) for the Hilbert symbol over Qp . We know from the formulas of § 63B that the Hilbert symbol is completely determined once its values are known, first for all rational integers a and b that are prime to p, and secondly for all rational integers a that are prime to p with b = p. We shall therefore restrict ourselves to these special cases. 73: 1. Let ft be an odd prime number, and let a and b be rational integers prime to p. Then a,b\ a\
‘25 ) —" k15 1 91 . Proof. The prime spot p is non-dyadic since the prime number p is odd. Apply Example 63:12, using the fact that the p-adic unit a is a square in Qp if and only if (7 a, = 1.
q. e. d.
73:2. Let a and b be rational integers prime to 2. Then
ta,b\
k
2)
a-1
(
—
b-1
(a2 -1)
2
1)
(—
Proof. 1) Every odd rational integer is clearly congruent to one of
the numbers 1, 3, 5, 7 modulo 8. Hence by the Local Square Theorem we can assume without loss of generality that a is one of these four numbers. The same with b. Now if u denotes the group of units of the ring of 2-adic integers Z2, then (u:u2) = 4 by Proposition 63:9. But every element of u is a square times 1, 3, 5 or 7 by the Local Square Theorem and the power series expansion of Example 31:5. Hence the numbers 1, 3, 5, 7 fall in the four distinct cosets of u modulo u 2. In particular 5 is a non-square, hence it is a unit of quadratic defect 4 Z2 . So by applying Corollary 63:11a we find that our proposition holds whenever a or b is 5. Of course the proposition holds whenever a or b is 1. We therefore restrict ourselves to a 3 or 7, b = 3 or 7. 2) We have 7 + 2(3)2= 25, hence - -
f7 2 2
k
1 8 -= 1
Chapter VII. Hilbert's Reciprocity Law
207
Then (3,2 2 ) ( 5,22 ) 2 ) = ( 7,2
8 •
Hence the second formula of the proposition holds for all a. 3) It is easily seen that (33\_ 2 )2= ( 3,7 ) _ (7,7 2 ) .
We will be through if we can prove that these three quantities are — 1. Now by Proposition 63 : 13 there is a 2-adic number c such that (--f 7 ) =— 1 2 since 7 is a non-square in Q2. But (1 ) = 1 by step 2), hence we can assume that c is a 2-adic unit, hence that c is 1, 3, 5 or 7. But c cannot be 1 or 5, hence c is 3 or 7. In either event we have our result. q. e. d. We cannot resist giving a proof of the famous Quadratic Reciprocity Law. This is obtained instantly from the Hilbert Reciprocity Law and the above formulas. Here then is the Quadratic Reciprocity Law with its first and second supplements. 73:3. Theorem. Let p and q be distinct odd prime numbers. Then p-1
P . ( I,q ) = (— 1) (-0 p-1 ---i-
-1
(T) — (
1)
q--1
2
2
f
/2
'
Proof. By the Hilbert Reciprocity Law we have
iii (12m _ 1 where 1 runs through all prime spots including 00. But ( Poe"' ) — 1 since p and q are positive real numbers. And if 1 is any prime number distinct from p, q, 2 we have — 1 by Proposition 73:1. Hence ( 111 / q
)
(P q\ p ) ( P yq )— — (—2 P q) Apply Propositions 73:1 and 73:2. This proves the reciprocity law. The first supplement is obtained in the same way from the equation ,
H ( ---1, ' P ) _
I
,
i
the second from the equation
H i2,p1_ 1. 1k 11
q. e. d.
208
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Part Four
Arithmetic Theory of Quadratic Forms over Rings Chapter VIII
Quadratic Forms over
Dedekind Domains
The rest of the book is devoted to a study of the equivalence of quadratic forms over the integers of local and global fields. Our first purpose in the present chapter is to state the nature of this problem in modern terminology and in the general setting of an arbitrary Dedekind domain. Our second purpose is to develop some technique in this general situation. The more interesting results must wait until we specialize to the fields of number theory. . We carry over the notation of Chapter II. F is a field, 0 = 0 (S) is a Dedekind domain defined by a Dedekind set of spots S on F, I = I(S) is the resulting group of fractional ideals, u — u (S) is the group of units of F at S. So here y is the ring of integers of our theory. As in § 22C we shall allow the same letter p to stand for a spot in S and also for the prime ideal of o that is determined by this spot. V will denote an n-dimensional vector space over F. In the second half of the chapter we will make V into a quadratic space by providing it with a symmetric bilinear form B: V x V -->- F. The general assumption that the characteristic of the underlying field F is not 2 will not be used in the first paragraph of this chapter.
§ 81. Abstract lattices § 81A. Definition of a lattice Consider a subset M of V which is an o-module under the laws induced by the vector space structure of V over F. We define
FM = lax! a EF, x EMI.. Since M is an y-module, and since F is the quotient field of o, we have
FM --, {ce-liel cc Eo,
oz+0,
x EM} .
From this it follows that FM is a subspace of V, in fact the subspace of V spanned by M. Given a C F and a E l we put
ocM = {cxxlx EM} ,
aM=afixiflEa,xEMI VIII
Chapter
VIII. Quadratic Forms over Dedekind Domains
209
These are again e-modules and the following laws are easily seen to hold:
a(M n N)=-- (OEM) n (OEN) (ao) M = OEM, (aa) M — (a + b) M = aM a (M
(a M)
(ab) M — a (6M)
bM ,
N) = QM aN
F(M N)— FM -PFN. We call the above o-module M a lattice in V (with respect to o, or with respect to the defining set of spots S) if there is a base x1 , . . x n for V such that M o x, • - • I- o x, ; we say that M is a lattice on V if, in addition to the above property, we have FM V. In particular, 0x 1 + • • • + ox a lattice on V. The single point 0 will always be regarded as a lattice. 8 1 : 1. Let L be a lattice on the vector space V over F. Then the o-module M in V is a lattice in V if and only if there is a non-zero a in o such that OEM L.
Proof. 1) First suppose that M is a lattice in V. So there is a base xn for V such that xi +
M
• • +
Oxn.
Since L is on V we can find n independent elements y 1 ,.. . , yn € L. Write
a1y j= 1
(ai F) •
These a" generate a certain fractional ideal, hence there is a non-zero a in such that aa15 (o for all 1,j. Hence
ocx5 (oy1 +»+ øyÇL , hence OEM C L. 2) Now the converse. We have a non-zero a in such that OEM c L. Since L is a lattice there is a base z1,...,z„ for V such that L Ç o z1 + • • •+ o Zn . Then
m
cc—iL
Hence M is a lattice in V.
o ( z.1 )
•••+
z:) . (-q. e.
d.
8 1 : I a. Let U be a subspace of V with Mc UC V. Then M is a lattice in V if and only i/it is a lattice in U. O'Meara, introduction to quadratic' forms 14
210
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Proof. Take a base z11 . .
x, for U and extend it to a base x1, .. . , for V. Put L'= x 1 + • • + ox, and L = oxi + • • • + ox.. If M is a lattice in U, then OEM Ç L'C L for some non-zero a in o, and so M is a lattice in V. If M is a lattice in V, then OEM CL for some non-zero a in o, hence OEM CLr\U=L', and so M is a lattice in U. q. e. d. It follows immediately from the definition that every submodule of a lattice is a lattice. In particular L n K is a lattice whenever L and K are lattices in V. And Proposition 81:1 shows that aL, aL, L K are lattices for any a E F, a E I. Clearly o x is a lattice for any x in V, and ax is also a lattice in V. Hence a1 z1 -1- - a, z,. is a lattice for any ai E I, zi EV. In particular, every finitely generated o-module in V is a lattice.
§ 81B. Bases Consider the lattice L in V. For any non-zero vector x in FL we define the coefficient of x in L to be the set a.= {a E F lax E L} . This is clearly an o-module in F, and it follows from the fact that L spans FL that it is not zero. Now
axx =LnFx, hence a. x is a lattice in Fx, hence a (a.x) Z ox for some non-zero a in o. Therefore aa.0 o, so that a. is actually a fractional ideal in F. Note that It is clear that
ccacco= (to aa; D
O
VocEiX
EL .•
We say that x is a maximal vector of .L if a= o. So x is a maximal vector of L if and only if L n Fx = ox
Every line in FL contains a maximal vector of L when the class number of F at S is equal to 1, i. e. when every fractional ideal is principal. For consider the line Fy in FL. Put ay . ao with a E F, then put x ---- ay; we have a.= o, hence x is a maximal vector of L that falls in the line Fy. 81:2. Given a lattice L on V, a hyperplane U in V, and a vector x o in V — U. Then among all vectors in x o U there is at least one whose coefficient with respect to L is absolutely largest. Let this coefficient be a. Then for any vector x o uo (uo E U) with coefficient a we have L = a (x0 /40) (L n U) .
Proof. 1) We claim that the set a . {a EFlax0EL
Chapter VIII. Quadratic Forms over Dedekind Domains
211
is a fractional ideal in F. It is clearly a non-zero o-module in F. And bY Proposition 81:1 there is a non-zero )3 in o such that /3L oxo + U. Hence (pa) xo _Ç /3L + U oxo + U. So fia cZ o. Hence a is indeed a fractionalideal as claimed. Now the coefficient of any vector in x0 + U is contained in a by definition of a. Hence the first part of the proposition will be proved if we can find a vector u in U such that a (x0 + u) Ç L. Since aa-'= o we can find an expression alfli+ • • • + GerPr= 1 Now each ai provides an expression mi x° = li + ui
E crl)
(ociE ftiE
by definition of a. Then x0=E
But ili a Ç o for 1
Pili+ E Pi ui•
r. Hence a (x0— E
piu)c L.
So we have found u E U such that a (x0 + u) Ç L and the first part of the proposition is proved. 2) We are given that a is the coefficient of x0 + uo, hence
a (x0 + 140) + (L r U) Ç L. . We must reverse this inclusion relation. So consider a typical vector in L which has the form ex (x0 + u) with a EF, uE U. Then axo EL + U and so a E a by definition of a. Hence ex (u
uo) = cc (xo + u) — ex (xo + uo)
Therefore
•
a (x0 + u) = oc (x0 + /40) +
EL•
(u — u0) E a (x0 + uo) + (L
81:3. Theorem. L is a lattice on the vector sfiace V and xl, a base for V. Then there is a base . . , y , with
y; EF xi + • • • + F
q. e. d. , x, is
(1 j n) ,
and there are fractional ideals a11 . . . , a., such that
L= + • ' + anYn • Proof. Let U be the hyperplane U = F xi + • • Proposition 81:2 L = (L U) + ayn
F x._1. Then by
14*
212
Part Four. Arithmetic Theory of Quadratic Forms over Rings
for some fractional ideal a n and some y„ E V — U. Proceed by induction
on n — dim V. 4. e. d. 81:4. Example. Let ,x„ be a base for V and letL = %xi + .. . + a„xn with ai E I. Then the coefficient with respect to L of any vector of the form CC1X1+ • • • + rX i
is equal to
(Cti
E F)
(a1 czT1) n • • n (a? ce71) In particular, the coefficient of x i is equal to ai. 81:5. Let L be a lattice on the vector space V. Then there is a tractional ideal a and a base z1, . . zn tor V such that L = azi + 02.2 + • • • + oz,,. Proof. Let us write L = aih+ • • • + any,, in the manner of Theorem 81:3. If n 1 we have L = alyi and we are through. The case of a
general n 3 follows by successive applications of the case n = 2. So let us assume that n = 2. By Proposition 22:5 we can find ch, tx2 E P such that + oc2 ail= o. Put x = z1 y 1 + oc 2y2. Then the coefficient of x in L is equal to
((Ilan (a2cV) = ( ohari. ce2a n-i. = by Examples 22:4 and 81:4. Hence L = ox + by by Theorem 81:3. q. e. d. 81:6. Example. An 0-module in V is a lattice if and only if it is
finitely generated. We say that the base z1, , z„ for V is adapted to the lattice L if there are fractional ideals al, .. . , a„ such that L = arzi d- • • + an zn .
Theorem 81:3 asserts that, there is a base for FL that is adapted to L, where L is any lattice in V. Consider a lattice L in V. It follows immediately from the fact that F is the quotient field of o that a set of vectors in L is independent over o if and only if it is independent over F. Hence a set of vectors of L is maximal independent over o if and only if it is a base for FL. In particular, any two such sets must contain the same number of elements. This number is called the rank of L and is written rank L. Thus rankL dimFL A set of vectors is called a base for L if it is a base in the sense of o-modules, I. e. if it is independent and spans L over o. So xl , . , x i. is a
Chapter VIII. Quadratic Forms over Dedekind Domains
213
base for L if and only if it is a base for FL with L ------ 0x 1 + - • • +oxr .
A lattice which has a base is called free. Any two bases of a free lattice L contain the same number of elements; this number is called the dimension of L and is written dim L; we have dimL = dimFL = rankL .
Every lattice L is almost free in the sense that it can be expressed in the form L = axi + ox2 -F • • • + ox7 with a a fractional ideal and xl, . . . , x, a base for FL. Clearly every lattice is free when the class number is 1, i. e. when every fractional ideal is principal. § 81C. Change of base Consider two lattices L and K on the same vector space V and let xl , . • . , xn and Yi'. . . , yin be bases in which L = ai x, ± • • • + anxn (ai E I) K= boi ± • • • -{- bnyn
Let
yi = E aii xi ,
(bf E I) .
xi = E biiyi
be the equations relating these two bases. So (ail) is the inverse of the matrix (k J). 81:7. K c L if and only if a ii bi Ç ai for all i,j. Proof. We have K S L if and only if bly, S L, i. e_ if and only if bi (aii xi + • • • + aii xi + • • •) S %xi + • • • + ai xi + • - • n. This is true if and only if ail bl S ai for 1 S i ‹ ti, 1<j‹ n. q. e. d. 81:8. Suppose K c L. Then K. L if and only if
for 1 ‹ j
a1 . • • an bi • • . b. • det (ail) . Proof. First consider K = L. Then by Proposition 81:7 we have c ai for all i., j, hence —
det (ail) = E ± atx . . . an .
E E (al bc71) • • • (an bT01) =
(al - - • an) Oh • • • bn) -1 •
Hence (b1 • . . b n ) det (ail)
(ai. • - an)
214
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Similarly (a1 . . an) det (k J) g (t.
.
Now det (a15) is the inverse of det (1)15). Hence the result follows. Now the converse. Since K L we have ai ,E ai br for all relevant 1,1. The cofactor A i; of a15 is equal to A ii — E
in which the first index avoids i and the second j. Hence A ij a i
(a1 .
an)
. . bn) -1 = 0 • det (ai i)
Therefore b a
A il —
det(a ii)
ab -
• 2
This is true for all i, j. Hence L S K by Proposition 81:7. Hence L
K.
q. e. d.
Recall that the elements of 0 are the integers of our theory. Accordingly we say that an n X n matrix (a15) with entries in F is integral (with respect to o, or with respect to the defining set of spots S) if each of its entries is in 0. We shall call (a15) unimodular if it is integral with det (a15) a unit of F at S. By looking at cofactors one sees that the inverse of a unimodular matrix is integral, and hence unimodular. The defining equation of the inverse of a matrix shows that if an integral matrix has an inverse, and if this inverse is integral, then both the matrix and its inverse are unimodular. In other words, an integral matrix is unimodular if and only if it is invertible with an integral inverse. , x„ and consider 81:9. Suppose L is a free lattice with base vectors Yi' . . . , y ,, determined by yi = E aijx i (aiJEF) Then these vectors form a base for L if and only if the matrix (a15) is uni-
modular.
Proof. This is an easy application of Proposition 81:8. q. e. d. . 8 1:10. Example. Let v,. be vectors in V, let s be a unit in 0, and let oc2 , . . a,. be elements of 0. Then Ovi + 0v2 + • • • +
02. + 0v 2 + • - • 0v,.
where 2-11 = svl -F ot2 v2 + • • • + § 81D. Invariant factors 81:11. Theorem'. Given lattices L and K on the non-zero vector space V. 1 This theorem can be used to derive structure theorems for finitely generated modules over the Dedekind domain O. These structure theorems reduce to the Fundamental Theorem of Abelian Groups when 0 is the ring of rational integers Z.
Chapter VIII. Quadratic Forms over Dedekind Domains
215
Then there is a base x1,. . xn for V in which L = al xi + • • • + an xn K= ai r,xi + • • • + afi r„xn where ai and r i are fractional ideals with
r2 D • • • D tn
•
The ri determined in this way are unique. Proof. 1) We can suppose that K S L (if necessary we can replace K by ŒK where cc is a suitable non-zero element of o). For any x in V we let a. denote the coefficient of x in L and 6. the coefficient of x in K. Then put rx = b./a.. Since we are taking If S L we have b. S a2, and so
rx Ç 0 . 2) We can therefore take a yEÙ for which ry is maximal (though we are not yet sure it will be absolutely largest). By Theorem 81:3 there is a hyperplane U such that L = a„v + (L U) . We claim that b„+„ S b„ for any u E U. If not, then by Proposition 81:2 we can find auEU for which b i,+„ by. But ay+ .(v u) g L = a„,v + (L r U) , so that ay+ . S ay. Hence by+. b. r„ D tv+u — av+u and this contradicts the choice of v. So we do indeed have by+ .c b„ for all u in U. Now apply Proposition 81:2 again. We obtain K b„v + (K r\ U) .
3) An inductive argument gives us expressions L a„v + (a w w + • • + a„z) K a„r„v + (a w r w w + • • • + a z rz z)
with rw • • • D rx. We shall therefore be through with the first part of the theorem if we can prove that r„ r iy. By Example 22:9 we can pick cc, /3 E fr in such a way that + #(07,Ir.7,1) o I cca7, 1 + flawl = r.„ + rw . Put x ccv + 13w. By Examples 22:4 and 81:4 we obtain ox= (ovcc i) r (atufl -i) = (acql + '30;9' = (re + rw) -1 . And similarly b. = o. Hence rx r„ + r„). So ry„ S r„ by choice of v, and this is what we required.
216
Part Four. Arithmetic Theory of Quadratic Forms over Rings
4) Now the uniqueness. Let us call ri the i-th invariant factor of K
in L (a formal definition will be made once the theorem is proved). Our purpose is to prove that the invariant factors are indeed invariant, i.e. that they are independent of the base used in defining them. We make a start by remarking that the product of the invariant factors is invariant: the reader may easily verify this by using Proposition 81:8. , rn in the given base and let Consider the invariant factors r1, ,r„' be the invariant factors with respect to some other base. Suppose if possible that the second base gives rise to a different set of invariant factors. Take the first i (1 i n) for which ri r:. We can suppose that we actually have ri + r Dr;. We put =K
(rL) .
A i 1 the 2-th Consider the invariant factors of J in L. For 1 invariant factor is rl in either base; for i A n it is r2 + r in the first base and el; in the second. But this means that the product of all invariant factors in the first base is strictly larger than the product in the second base. We have already remarked that these products must be equal. So we have a contradiction. Hence r2 = ri for 1 A n. q. e. d. 81:12. Definition. The invariants r1 D • • • D rn of the last theorem are called the invariant factors of K in L. And ri is called the i-th invariant factor of K in L for 1 i n). Suppose we have lattices K and L on V with K S L. In this event the invariant factors of K in L are all integral ideals. Referring to Theorem 81:11 and § 22D we find —
(L:K)
11
(a i :airi) =
.11
Igign
(o:ri) .
In the important situations (e.g. local fields and global fields) the indices (o :ri) are all finite; hence (L:K) <0°; and the number of lattices between K and L is finite. § 81E. Localization
Let us give some attention to the completions Fr of F at the spots p in S. We let I Ir stand for any fixed valuation on Fr at p. Of course the Dedekind theory of ideals applies to Fp at p, and the theory of lattices applies to vector spaces over Fp . We let 1-r denote the group of fractional ideals of Fp at p. Recall that or, up , mr are used instead of o(p), u (p), m (p) on Fr at p. Here the same symbol p denotes two different spots, the one on F and the other on Fr. And p also denotes two different prime ideals, the prime ideal)) n mr of 0 and the prime ideal mr of or. A typical fractional ideal of F at S has the form pvp
E
Chapter VIII. Quadratic Forms over Dedekind Domains
217
with all vp in Z and almost all of them 0; and a typical fractional ideal of Fp at p has the form PAP Cup E Z) • We introduce the surjective homomorphism defined by the mapping 17 pvp
pvp
p ES
The image of a E / (S) under this mapping will be written ap. We call ap the p-ification or localization at p of the ideal a. The following laws follow easily from the results of §§ 22B and 22C: (Œa) p = cop , (a (a + b) p = ap bp , laplp lalp • This last equation shows that a is always contained in ap, and in fact
a= We have
p ES
(F n ap) .
a Sb
al,Sbp VpES,
a=
ap = bp VP ES.
and in particular It is easily seen, again using' the equation laplp op-ideal generated by a in Fp. 81:13. a p is the closure of a in Fp. Proof. We know that
la lp, that a;, is. the
ap = {cc E Fp lialp
ialp} ; but the map a –>– 14 of Fv, into R is continuous; hence ap is closed in Fp. Why is ap the closure of a? We must consider a typical a E ap and an s > 0, and we must find an a E a such that la — 4 s. We can assume that 0 < s < 14) By the Strong Approximation Theorem we can find an a in F such that la — alp < 6
Vci E S 5- lalq The Principle of Domination then insists that lai r = lair Hence lalq 5_ lak q. e. d. for all q in S. So a is in a. Now consider an n-dimensional vector space V over F. As in § 66 we use V,, for the localization Fp V of .V at p. Let L be a lattice in V (with respect to the set of spots S). By the p-ification or localization Lp laic'
Part Four. Arithmetic Theory of Quadratic Forms over Rings
218
of L in Vv at the spot p in S we mean the or-module generated by L in Vv. By the definition of a lattice there is a base xi., . . xn for V such that L
0 + - • • + 0 xt, ,
hence
Lr
+ • • + o v x„ ,
so Li., is a lattice in Vv (with respect to o r, or with respect to the spot p on F).
If L and K are lattices in V, then it is easily seen that VrES. (L K)4, = ±Kr Now (a x) = av x for any a E I (S) and any x in V, hence (not + • • • + ar ti.) v alv z, + • • • + arv zr for any ai E (S) and any zi E V. If we take a base x1 .. x n for V such that L= -F • • • + ar x, , then ,
Li)
p ± • • • +
ar
p .
Hence V n L = (F r a1p)x1 + • • • + (F r arp)xr ,
so
n
(V r\ Lr) L .
pES
We therefore have LSK
Lr Kr
V p ES,
L=K
Lr = Kr
V P ES.
and in particular, If FL = FK, then the Invariant Factor Theorem of § 81D shows that for almost all p in S. 81:14. An or-lattice hp) is given on VI, at each p in S. Suppose there is an o-lattice L on V with Lr = h o for almost all p. Then there is an 0-lattice K on V with Kr = he) for all p in S. Proof. 1) First we prove the following contention: given a spot p in S, there is a base . yn for V such that hp) = °ph ± • ' • + Op Ytt • To prove this we fix a base x1,. , xn for V; multiplying each of these basis vectors by a scalar which is sufficiently small at p allows us to assume that each xi is in hp) . Take a base for hp) , and write each
J4, ) = opih
• • + o plin
+ • + ani x„
(7/27 E 17,Y, (ccii E Fr)
.
Chapter VIII. Quadratic Forms over Dedekind Domains
219
Now -F,p, is the closure of F at p, hence we can find, for each i and j, an element aij CF such that 1
is as small as we wish. If the approximations are good enough we will obtain, in virtue of the continuity of multiplication and addition in the topological field Fp , the inequality det (ai — det(oc ii)1 1, < det (oc ii)i p .
Hence idet
= idet(oc ii)l p by the Principle of Domination. Put
y = a1 5 x 1 lor i
••
ani xn
n. Then 51; — m E op
by choice of the
(t ip
• • + opxn c= hp)
hence
°ph + • • • + opYn hp) • If we write y; f yij n i with all yi; Fp we have (7ii)
(Iii) -1 (a15)
hence det (y o) is a unit in vp , hence opYt ± • " opYrt
by Proposition 81: 8 . This proves our contention. 2) It is enough to prove the following: given a single spot p E S there is an D-lattice K on V such that L K —{ J (qr)
f f
CS —p = P•
The final result will follow by successive applications of this special case. In virtue of step 1) we can find an o-lattice J on V with J = hp) . Use the Invariant Factor Theorem of §81D to find a base z1, zn, for T7 such that Ç L = arzi + • • • + anz. + • • -1-- b.„z„ j=
Construct fractional ideals c i (1 c"
i
n) with
I Dig if q(-S — p bip if q = p
Then K
has the required property.
+ • - • - I- enzn
q. e. d.
220
Part Four. Arithmetic Theory of Quadratic Forms over Rings
§ 82. Lattices in quadratic spaces
An additional structure is imposed on the situation under discussion: the vector space V is made into a quadratic space by giving it a symmetric bilinear form B with associated quadratic form Q. We shall call a lattice L in V binary, ternary, quaternary, quinary, , n-ary, according as its rank is 2, 3, 4, 5, . . . , n. § 82A. Statement of the problem Consider a lattice L in V. Let Ube some other quadratic space over F and consider a lattice K in U. We say that K is represented by L, and write K —›— L, if there is a representation o: FK —>— FL such that aK S L. We say that there is an isometry of K into L, and write L, if there is an isometry : F K >-- FL such that aK Ç L. We say that K and L are isometric, and write or K , if there is an isometry ci: FK >—). FL such that aK L: We pose the following fundamental question: can we determine, say by means of invariants, whether or not two given lattices L and K are isometric. This should be regarded as the integral analogue of the earlier work on quadratic spaces. And just as the question for spaces is a geometric interpretation of the classical problem on the fractional equivalence of quadratic forms, so the question for lattices can be regarded as a geometric interpretation of a classical question on the integral equivalence of quadratic forms. In fact it will be shown in § 82B that finding the lattices isometric to a given free lattice is the same as finding the integral equivalence class of a quadratic form. This poses the question. What about its solution ? It is certainly too much to expect an answer over an arbitrary Dedekind domain at the present time. However, a complete solution is known when o is the ring of integers of a local field, and considerable work has been done over the integers of a global field. These theories will be presented in the remaining chapters of this book. If a solution to the space problem is known over the field F (for instance if F is either a local field or a global field) then there is no loss of generality in assuming that FK and FL are identical. Under these circumstances we can ask our question in the following form: given lattices L and K on the quadratic space V, does there exist an element a of the orthogonal group 0(V) such that aK L? § 82B. The free case Let M be an m x m symmetric matrix and . /sI an n x n symmetric matrix, both over the field F. We write M —>— N (over c)
Chapter VIII. Quadratic Forms over Dedekind Domains
221
and say that M is integrally represented by N if there is an n X m matrix T with coefficients in o such that
M , tTNT , where tT stands for the transpose of T. If this can be done with a unimodular matrix T we say that M is integrally equivalent to N and write
M N
(over o) .
Integral equivalence of matrices is clearly an equivalence relation on the set of all n x n symmetric matrices over F. This equivalence relation depends, of course, on the set of spots S used in defining the ring o = o (S). We let clss N or clsN denote the set of n x n matrices which are integrally equivalent to the n x n symmetric matrix N over F, and we call this set the class of N at S. These classes partition the set of n x n symmetric matrices overF at S. If L is a free lattice on V there is a base x1 . . x. for V such that ,
L =oxi + • • • + oxn . By the matrix of L in the base x1 ,. N (B (xi , xi)), i.e. the matrix of V in
x„ we mean the matrix xn ; we write
LN If there is at least one base x. for which this holds, then we say that L has the matrix N and we write
LN. If
. • . , x is another base for L with x;
2: tl.ixA
(tau ( 0)
then T = (tu) is unimodular by Proposition 81:9, and the matrix N' of L in xi, . . x:, is equal to
N' ITNT In other words a change of base leads from N to a matrix N' in clss N, and every matrix of clss N can be obtained in this way. We can therefore associate one entire class of integrally equivalent matrices with a free lattice on a given quadratic space. And every class of integrally equivalent symmetric matrices can be obtained from a suitable free lattice on a suitable quadratic space. Given any symmetric matrix N we have agreed to let
222
Part Four. Arithmetic Theory of Quadratic Forms over Rings
82: 1. Let K and L be free lattices with matrices
M and N on the
quadratic spaces U and V, respectively. Then (1) K --->— L if and only if M —›— N (over t) if-and only if M N (over t) . (2) . K L
q. e. J. Proof. The proof parallels the proof of Proposition 41:2. , xn of the Consider the discriminant dB (x1, . . xn) of a base xi., x„' for L the equation free lattice L on V. If we take another base xi, 1 T/VT shows us.that N' dB (x;, . . x) = 62 dB (x 1, . xn) , xn) in for some unit e in o. Hence the canonical image of dB (x 1, 0 u ( 1 /u2) is independent of the base chosen for L, it is called the discriminant of L, and it is written dBL or dL . When L consists of the single point 0 we take dL — 1. We shall often write dL = a with cc in F; this will really mean that dL is the canonical image of cc in 0 u (F/112). It is equivalent to saying that L has a base L = °x1 + - - • ± ox, in which .
.
§ 82C. The class of a lattice
Consider a regular non-zero quadratic space V over F, and lattices K, L,... on V. We say that K and L are in the same class if K aL for some a E 0(V) . This is clearly an equivalence relation on the set of all lattices on V, and we accordingly obtain a partition of this set into equivalence classes. We use cisL to denote the class of L. The fundamental question of § 82A can now be regarded as a question of characterizing the class clsL. We define the proper class cisf to be the set of all lattices K on V such that K ciL
for some a E 0+ (V) .
The proper classes also put a partition on the set of all lattices on V, and this partition is finer than the partition into classes. In fact it is easily verified, using the fact that 0+(V) has index 2 in 0(V), that each class contains either one or two proper classes. Of course the class and
Chapter VIII. Quadratic Forms over Dedekind Domains
223
the proper class depend on several factors, such as the underlying set of spots S, the supporting vector space V, and the underlying bilinear form B. We define the group of units of L to be the subgroup 0 (L) = {a
0 (V) a L =
of 0(V). We put 0±(L) --= 0(L) r 0+(V)
.
The determinant map
det: 0(L)
(± 1) has kernel 0+ (L) , hence 0+ (L) is a normal subgroup of 0(L) with
1
(0 (L) : 0±(L))
2.
We shall see later that it is possible for this index to be either 1 or 2. It is 2 when there is at least one reflexion on V which is a unit of L; otherwise it is 1. We define the set 0- (L) 0 (L) n 0(V)
.
Then 0(L) = 0±(L) u 0- (L) , 0+(L) r 0- (L) 0.
82:2. Example. a 0 (L) a-1 = 0 (a L) for any a in 0(V). 82:3. Example. clsL = cls+L if and only if (0(L): 0+ (L)) , 2. 82:4. Example. clsL = cls4 L if dim V is odd, since —1 v is in 0(L) but not in 0+(L). 82:5. Example. Suppose L is free and let x1, with L = oxi + • • • + oxn
.
.
xn, be a base for V
Let M denote the matrix of V in this base. According to § 43A there is a group isomorphism of 0 (V) onto the group of automorphs of M, obtained by sending an isometry a onto its matrix T in the base x1,. This isomorphism carries rotations to proper automorphs. What does it do to 0 (L) ? It carries the units of L to the integral automorphs, Le. to the automorphs with integral entries. And to 0- '- (L)? These elements are carried to the proper integral automorphs of M. Thus (0 (L) : (L)) = 2 if and only if M has an integral automorph of determinant —1. So clsL = cls+L if and only if there is an improper integral automorph of M. 82:6. Example. Let K and L be free lattices on the same quadratic space V with matrices M and N respectively. Then clsK = cisL if and only if clsM = clsN. § 82D. Orthogonal splittings Consider the quadratic space V provided with its symmetric bilinear form B and its associated quadratic form Q. Let L be any lattice in V.
224
Part Four. Arithmetic Theory of Quadratic Forms over Rings
We say that L is a direct sum of sublattices L1 , .. L,. if it is their direct sum as o-modules, i.e. if every element x E L can be expressed in one and only one way in the form x=
+••-
(xi E L) .
x,.
We write L=LI ED••••L,.
for the direct sum. We know, for example, that L = L1 ED
L2
if and only if
L.4-1-L 2 with L1 nL2 =0. Suppose L is the direct sum L =L1 e • • • e L,. with = 0 for
B(Li,
1
<
r
We then say that L is the orthogonal sum of the sublattices L1, . . or that L has the orthogonal splitting L=4±•••±L,.. We call the L i the components of the splitting. We also use the notation
± L i and
ED Li 1
1
for orthogonal sums and direct sums respectively. We formally define j_ L i = 0 and ED L i = 0 . 0 0 We say that a sublattice K splits L, or that K is a component of L, if there is a sublattice J of L with L=KJ
If X1 . . X,. are sublattices of different quadratic spaces, not necessarily ,
contained in V, we write L
Xi 1 • • • 1 X,.
to signify that L has a splitting L=
1 • • •
± Lr
in which each sublattice L i is isometric to Xi. Suppose we are given quadratic spaces Vi (1 I r) over F and lattices L i in Vi. Then we know that there exists a quadratic space V over F such that V 24 Vi 1 • .j V,. . Hence there always exists a quadratic space V which contains a lattice L such that L ± • • • ± Lt .
.
Chapter VIII. Quadratic Forms over Dedekind Domains
225
There is a slight modification of the preceding construction which we shall call adjunction. Consider quadratic spaces U and V over F. Then there is a quadratic space W which contains the quadratic space V and is such that W = U' ± V with U' U .
We say that W is constructed from V by adjunction of U. Now let K and L be lattices in U and V respectively. Then there is a lattice J in W such that J = K' L with K' K .
We say that J has been constructed from L by an adjunction of K. As an example of the notation consider the equation L
±
with M and N symmetric matrices over F. This means that L is a free lattice in a quadratic space with
L=—( Similarly L
(cci >
11/1 0 1— O/sT) . • • 1.
with all cci in F means that L has a base x11 . . x„ in which Q (xi) — cci for 1 I n and B(xi, xj) = 0 for 1 I <j Sn. A base with this property is called an orthogonal base for L. It is easily seen that the direct sum of lattices L= Ll e-••eL,. implies the direct sum of spaces FL = FL, e-••e FL, . Similarly with orthogonal sums. These equations show that rankL = rankLi + • •
rankL,..
We define the radical of a lattice L in a quadratic space V to be the sublattice radL =
EL IB (x, L) = 0} .
We call L regular if radL = O. It is easily seen that radFL = F radL
and radL
L
radFL
In particular, L is regular if and only if FL is regular. We have rad(L i K) = (radL) j (radK) . O'Meara, Introduction to quadratic forms
15
226
Part Four. Arithmetic Theory of Quadratic Forms over Rings
In particular, L K is regular if and only if L and K are regular. Suppose J—LIK with L regular; then K is uniquely determined by and L, and in fact K {x J B (x, L) 0} . Let us show that L has a radical splitting, i.e. that there is a lattice K such that L = K j radL. If L is regular just take K L. Now assume that L is not regular and take a base x11 . . xn for FL in which x1,. . x,. span the radical of FL. By Theorem 81:3 there is a base yi, . y,z for FL in which ' • • + arYr + • • • + anYn and radFL == Fyi • • + F y,. . Then radL L radFL = a1 y1 + • • a,.y,.. So taking K = ar_Fo r+i + • • • + anYn gives us the desired radical splitting L = K radL. Incidentally, the lattice K in any radical splitting L = K j radL is always regular since the equations radL = radK j rad (radL) radK j radL
imply that radK is O. If L = K j radL and L1 = K1 I radLI are two radical splittings, then it can be shown (using .a proof somewhat similar to the proof of Proposition 42:8) that L is isometric to L1 if and only if K K 1 with
radL212- radLi •
82 : 7. Let L be a lattice in the quadratic space V and let K be a regular sublattice of L. Suppose K has a splitting K=
_1_ • • • ± K,. .
Then K is a component of L if and only if each K i is. Proof. If K splits L, then it is clear that each K i does too. Conversely suppose L is split by each K. Since FK is regular we have a splitting FL = FK in which U is the orthogonal complement of FK in FL, by Proposition 42:4. We claim that L K (U n . In order to prove this it is enough to show that L = K (U L). So consider a typical x in L. For r we can write x = xA 3/2 with xA E KA and B (y A, K A) = 0, since KA splits L. Then B (x—E xA , K i) = B(x— xi, K i) B (y i , K i) = 0 2
Chapter VIII. Quadratic Forms over Dedekind Domains
for 1
227
i < r. Hence B (x — E xA,F K) . 0. a ‘
So x— E xA E U n L . a
Hence -47 XA) E K + (U n L) . X = (I ' XA)± (X q. e. d.
§ 82E. Scale, norm and volume Consider a lattice L in the quadratic space V. By the scale sL
of L we mean the o-module generated by the subset B (L, L) of F. Clearly sL = 1 E B (x, y)jx,y E
9.
I fin
Since L is finitely generated we can write L = o z1 + • • • + o zr, hence sL C E B(z i, zj)o , ci
so sL is either a fractional ideal or 0. We define the norm nL
to be the o-module generated by the subset Q (L) of F. Since Q (L) s B (L , L) it follows that nL is also either a fractional ideal or 0. Now for any x, y in L we have 2B (x, y) = Nx + A - 12 (x) - Q(y) E nL . Hence 25L ç la C sL . We also have Q (F L) = F2 Q (L) .
So all the sets oL, nL, B(FL,FL), Q(FL)
are 0 if one of them is. If L has the splitting L = J ± K, then it is easily verified that sL = s, f + 6K,
nL—ni-l-nK.
82:8. Let the lattice L in the quadratic space V have the form L = arzi -I- • • • + ar zt with the a i in I and the zi in V. Then (1) sL = E a i ai B(z i,zi), iI
(2) nL -- E 4Q(z) +26L. i
15*
Part Four. Arithmetic Theory of Quadratic Forms over Rings
228
Proof. 1) First we do the case r — 1. Clearly n(a01) c (a01) Ç a7+2' (z1) .
Let us prove that aTQ (xi) C n(a01). Write al — ao + fib with a, /3 in al. Then cc2 Q (xi) = 02 (azi) E 02 (al zi) ç n (al zi) , and similarly 132 Q (xi) E n(aiz,L). Hence by Example 22:3
alQ (zi) —
(x20 + 1320)
Q(z) ç n (alz1) .
So n (a1 z1) . s (aizi) . alo2 (z1) .
This proves the proposition in the case r — 1. 2) Next we find the scale for any r> 1. Clearly B (x, y)
E E ai aj B (xi, z1) , ‘i
for any x, y in L, hence
E ai ai B (xi, z).
5 L ._Ç
i,
Conversely, for any oc iE di and any pi E al we have oci fli B(zi, xi) — B(aizi ,
thzi) E sL ,
hence ai aj B(zi, zi) S sL , hence f d i al B(zi, xi) = sL .
i,i 3) Finally the norm equation. For a typical vector
x = oh; + - - - ± ar zr in L we have
(aci
E ai)
Q (x) = f 04 Q ( z1) + 2 f oc; B (z 1 , 2 .1) ,
hence tIL S
E alQ (xi) + 2sL . t
On the other hand we have 2sL S nL, and by step 1) aN (zi) . n (aizi) nL , hence nL — E aN(zi) + 2sL . i
q. e. d. 82:8a. For any fractional ideal a we have s (a L) . a 2 (5 L) and n(aL) ----- a2(nL).
Chapter VIII. Quadratic Forms over Dedekind Domains
229
82:9. Let K be a regular non-zero lattice in the quadratic space V. Then there is a lattice L on V which is split by K and has the same scale
and norm as K. Proof. Since K is regular we have a splitting V = (FK)1 U. Take any lattice J on U. Now sK and nK are non-zero, 'fence they are fractional ideals, hence there is a non-zero a in o such that s (c J) = oJ) Ç sK ,
and n (ccj) = a2 (nj) SnK
Then L = K I (aJ) is a lattice on V with the desired properties. q. e. d. If L is not 0 there is a base xi, x,. for FL such that L a1 x1 + • • • + arx,.
with the ai in I. We define the volume to be L = a? . . . 4. • d
, x,.) .
This quantity is 0 when L is not regular, it is a fractional ideal in F when L is regular. Is it independent of the choice of base for FL? Consider another adaptation L = b 1 y1 + • • • + by
with yi
E aii xi .
Then by §41B we have d•, y,.) = (det a15) 2 • d (x1, . . „
,
hence by Proposition 81:8 we obtain 1)? . . .
• d(y 1 , • • • ,
= a? . . . 4. • d
, x,.) .
So bi- is indeed well-defined. If L is the lattice 0 we define 1.11, = 0. We note that »L = • »K when L has a splitting L =J I K. 82:10. Let L be a non-zero lattice in the quadratic space V. Then
uL C (en? where r denotes the rank of L. Proof. Take an adaptation L a ixi + • • • ar x,.
in the usual way. Put gii = B(xi, xj).Soa i aj gij Ç sL by Proposition 82:8.
230
Part Four. Arithmetic Theory of Quadratic Forms over Rings
By definition of the determinant we obtain »L= ... 4.) • E
g„)
(al aAŒ) . . . (ar aco g,..)
ç_ (51,) ? . q. e. d. 82:11. Let K and L be non-zero lattices on the regular quadratic space V with K L. Let a be the product of the invariant factors of K in L. Then a is an integral ideal and a2 (»L) , aL
Proof. Let is a base
K
rn be the
invariant factors of K in L. Then there xn for V and there are fractional ideals ., an suchta
I
L . alx,_ + • - • + anxn K= mi x,. ± • • • + an rn x,z .
Here we have all ri ç o since K C L, hence a C 0. If we compute volumes with respect to the above base we find UK . a2 (bL). Finally, we have q. e. d. a S ri for 1 < i < n, hence aL S K. 82:11a. »KC»/. if K C L.
Example. Consider the lattice L on the regular quadratic space V and let a be an element of 0(V) such that aL C L. Then a, being an isometry, will preserve volume. Hence aL . L, i.e. a is actually in 82:12.
0(L).
dual of a lattice Consider a lattice L in the quadratic space V. Suppose that L is regular. We define the dual of L to be § 82F. The
L4t . {x EFL [ B (x, L) Ço}.
If L is the trivial lattice 0, then 1.4* . O. Suppose L is not O. Then we have a base xl , . . ., x,. for FL and fractional ideals al., .. ., at such that L= al x„. + • • • + af xr .
We claim that L* =aj y1 +•••+ ç l y
where y„ . . y? is the dual base of 0 if i == j, and B (ai xi, B (ai xi, a-ph)
x,. on FL.
Now
C o otherwise. Hence
I-4f • On the other hand, if we take a typical vector z in L 4* we must have ar l Yi + • • • + (171 Yr
Thai
B
13iYi+ • • • + 13,.Yr
ai xi) = B (z, a i xi) c B (z, L) s; o,
Chapter VIII. Quadratic Forms over Dedekind Domains
231
and so /3, E aTl. Hence we have established our Claim. An incidentally this also shows that L 4 is a lattice. As immediate consequences we have FL 4 = FL,
L4 # L ,
and
(a L)* = a- L# for any fractional ideal a. In virtue of the fact that
d(xl, .
xr) d (yl, . .
yr)
1
(see Example 42:5) we must have
(DL) -1 .
DE*
If L has a splitting L J I K, then L4 = j4 K4 . Finally, it is easily seen that if L, J,K are all on the same space FK, then
L# K 4
L DK and (J
K ) -It =
n K4 .
§ 82G. Modular lattices Consider a lattice L of rank Y in the quadratic space V. Suppose that the scale of L is the fractional ideal a. Then we know from Proposition 82:10 that 5L — a and DL Ç ar. Suppose that L actually satisfies 6L,= a
and UL =a'.
In this event we call L a-modular, or just modular. For any oc in I we call L oc-modular if it is a-modular with a oco. We call L unimodular if it is a-modular with a = o. It is clear from Proposition 82:10 that L is a-modular if 5L a and 1)L = ar An a-modular lattice is regular, non-zero, of scale a, and of volume DL = If L is a-modular, then ocL is oc2 a-modular for any oc in F, and bL is b2 a-modular for any b in I. If we take a non-trivial splitting L J I K, then it is easily seen that L is a-modular if and only if both J and K are. The lattice ox with x an anisotropic vector of V is Q(x)-modular. 82: 13. The free lattice L in the quadratic space V has the matrix M. Then L is a unimodular lattice if and only if M is a unimodular matrix. Proof. Take a base x1 ,. . x,. for L in which L has the matrix M. So L = oxi + • • - +ox,. and DL = (det M)o. If L is unimodular, then all B(x I , x3) are in o since 5L — o, hence M is an integral matrix. Further-
232
Part Four. Arithmetic Theory of Quadratic Forms over Rings
more (detM)o .----- o and so detM is a unit. Hence M is unimodular. Now let M be unimodular. Then M is integral and so sL = f B(x i,xi)o S o. ci q. e. d. And detM is a unit so that 13./. = o. Hence L is unimodular. 82: 14. L is a non-zero regular lattice in a quadratic space V. Then L is a-modular if and only if al.* = L. Proof. First suppose that aL4* , L. Then B(L, L) = B(L, aL 4t) C a ,
and so sL C a. Furthermore b.L = » (a./,) = a2 ? (»L4) _, a2r 03 4-1 , hence u.L = ar. Therefore L is a-modular. Now assume that L is a-modular. Then B(L, a--1 L) C o so that a-1 /- C 1.4 t. On the other hand we have I) (a-1 L) -, a-2 r (» L) = (a)-1 = »L*. q. e. d. Hence a-'/, — L 4t by Corollary 82:11 a , 82:14a. Suppose L is a-modular. Then L = {x EFL f B(x, L) S a} . Proof. Since L is a-modular we have B(L, L) S sL = a and so L S {x EFL 1B (x, L) _Ç a} . Conversely consider an x in FL with B(x, L) .0 a. Then B (x, 1-#) = B (x, a --3-1.) C o , hence x E L# 4t = L. 82:14b. L is unimodular if and only if .1.4t = L.
q. e. d.
82 : 15. L is a lattice in a quadratic space V and I is an a-modular sublattice of L. Then J splits L if and only if B(J, L) C a. Proof. If j splits L we have L— JIK and then
B (j, L) = B (J, J) _Ç a.
Conversely suppose j satisfies the condition B(J, L) C a. Since J is modular it is regular, hence by Proposition 42:4 we have a splitting FL = (F J) _i_ U. We claim that L = J _I_ (L r\ U) . It is enough to show that a typical x in L is also in J + (L n U). Write x=y+z (y EFL z E U) Then B(y, J) = B(x, J) _ç_ B (L, J) s; a .
Chapter VIII. Quadratic Forms over Dedekind Domains
233
But y is in F J. Hence y is in / by Corollary 82:14a. So z is in L and we are through. q. e. d. 82:15a. If J. is an a-modular sublattice of L with sL a, then j splits L. 82:16. Let L be an a-modular lattice and let x be an isotropic vector in L. Then there is a binary lattice J which splits L and contains x. Proof. By Theorem 81:3 there is a base for FL which is adapted to L and includes x among its members, say L=bx+••••
Then by § 82F we have 1.4* = b-ly + • • •
where y is a vector in FL with B(x, y) = 1. Now a/At = L since L is a-mQdular. Hence = bx + b-iay
is a sublattice of L which contains x. But is easily computed and found to be a2. And 15/ S sL S a. Hence J is a-modular and therefore splits L. q. e. d. 82: 17. Suppose is a principal ideal domain. Then a non-zero lattice L in a quadratic space V is a-modular if and only if B(x, L) — a for every maximal vector x in L. Proof. First suppose L is a-modular. Then by Theorem 81:3 we can find a base x, . . . for L in which
L = ox + • • • .
So L 4t oy + • • •
where y is a vector with B (x, y) = 1. Now aL4* = L since L is a-modular, hence a B(x, L) B(x, ay) 2 a , so B(x, L) = a. Conversely suppose B(x, L) = a holds for every maximal vector in L. Every vector y in L falls in ox where x is a maximal vector of L in the line Fy, hence L is regular and s L -= a. So B (a-1 L, L) S 0. This proves that a-1 /, S LA and hence that L S aL4*. On the other hand B(aL 4t, L)S a. If y is a vector of FL—L, then y = ocx with x maximal in L and a not in o, hence B (y , L) = B(x, L) = oca
and this is not contained in a; so no vector y of FL — L can fall in a/ * . Hence aLA .0 L. So aL 4t = L and L is a-modular. q. e. d.
234
Part Four. Arithmetic Theory of Quadratic Forms over Rings
§ 82H. Maximal lattices
Let L be a non-zero lattice on the regular quadratic space V, and let a be a fractional ideal. We say that L is a-maximal on V if nL S a and if for every lattice K on V which contains L we have nKÇ a => K--=--L.
We say that L is maximal on V if it is a-maximal for some a. If L is an a-maximal lattice on V, then it is easily seen that ccL is ea-maximal for j_ K is a every cc in F , and bL is 62a-maximal for every b in /; if L non-trivial splitting of the a-maximal lattice L, then J. and K are a-maximal on F/ and F K respectively. 82 : 18. Let L be a lattice on a non-zero regular quadratic space V and suppose nL C a where a is a fractional ideal. Then there is an a-maximal lattice K on V with K D L. Proof. Ler r denote the dimension of V. First observe that for any L on V with nL C. a we have 1 )r L (sL)f The Unique Factorization Theorem shows that the number of fractional
ideals between bL and (a)i. is some finite number v. If L is not a-maximal 21 we have a lattice Li on V with L (LI and nL1 a. If LI is a-maximal we are through. Otherwise repeat the preceding step. Continue in this way to obtain a chain LCLI C• • •CLt
with each nLi c a. This gives rise to the chain of fractional ideals )r
So t v. Thus the process must terminate before 3, steps. In other words we obtain an a-maximal lattice Lt before y steps. This L t is our K.
4. e. d. 82: 18a. Every non-zero regular quadratic space contains an a-maximal lattice for every fractional ideal a. 82: 19. Let L be a lattice on the r-dimensional regular quadratic space V
and let a be a tractional ideal such that nL S a. Then the ideal 2r 03 ÷ (a)*
is integral. If this ideal has no integral square factors, then L is a-maximal. 1 . Proof. The ideal 2' (»L) Mr is integral since (D L) C (8 L) ( -1 Suppose this ideal has no integral square factors. If L is not a maximal we can find a lattice K on V with L (K and nK c a. Then there is a proper integral ideal b with DL = b2 (DK) by Proposition 82:11. And -
Chapter VIII. Quadratic Forms over Dedekind Domains
235
2f (bK)/ar is integral since nK C a. This implies that 2? Oa L)/ar= 62 2? (» K)/a'
has an integral square factor, and this is contrary to hypothesis. q. e. d. 82:20. Let L be an a-maximal lattice and let x be an isotropic vector in L. Then there is a binary lattice J . which splits L and contains x. Proof. Let V denote the regular quadratic space on which L is a-maximal. Let b denote the coefficient of x in L. We have 1 1 sL ç— --2 (nL) S -2a. Hence B (2 a-1 bx , L) S B(2a -1 L, L) So , hence (2 a-lb) x C 1,4, so 2a-1 6 S c where c denotes the coefficient of x in 1,4. We claim that 2a-1 6 = c. Suppose not. Then 2a-1 6 C c, so that 1 a cx + L is a lattice in V which properly contains L. Now an easy computation gives 1 Q ( 2 a cx ± L) S a . it So — 2 OCX + L is a lattice which properly contains L and has its norm contained in a. This denies the maxirnality of L. So we do indeed have 2a-1 b = c. By Theorem 81:3 we have a base for V which includes the vector x and such that L4 = 2a-1 bx +... . Then by § 82F there is a vector y with B (x, y) = 1 such that L . 1, 4*4*. -21 a 6-13, + • • - -
So we have a sublattice 1 ab-1 y I —bx-F-T
i a-modular. of L with s j S sL S -I1a and » J. = -4-1a2 ; therefore J. is T q. e. d. By Proposition 82:15 j will split L. 82:21. Let L be a lattice on the hyperbolic plane V. Then L is 2 amaximal if and only if L is a-modular with nL S 2a. Proof. Take x, y in V with Q (x) = Q(y) = 0 and B (x, y) . 1. First suppose that L is 2a-maximal on V. By Theorem 81:3 we can write L. bx+ c (ccx + y) for some a in F and some 6, c in I. Now b c S sL S a by Proposition 82:8. If we had I) c C a, then ac--1 x + c(ccx + y)
236
Part Four. Arithmetic Theory of Quadratic Forms over Rings
would be a lattice of norm 2a which strictly contained L and this would deny the maximality of L. Hence b c = a and L bx ab -1 (ccx y) .
So 6L Ç a and »L = 02. Hence L is a-modular. And nL S 2a since L is 2 a-maximal. Conversely, suppose that L is a-modular with nL S 2a. Then 22 (» L) -4 (20) 2 =0, hence L is 2a maximal by Proposition 82:19. q. e. d. 82:21a. Suppose either of the equivalent conditions of the proposition -
-
is satisfied. Let F x and F y be the isotropic lines of V. Then the base x, y for V is adapted to L. Proof. We can assume that B (x, y) . 1. In the proof of the proposition
we obtained Now by Proposition 82:8 a2 b-2 (2a) ç n.L Ç 2a, hence a b-la b, hence L bx q. e. d. 82:22. Example. Suppose every fractional ideal is principal, and let a be an element of P. Consider a lattice L on the hyperbolic plane V. Then L is 2 ao-maximal if and only if (0 ci L
ce 0 Hence in this situation any two a maximal lattices on V are isometric. 82:23. Let V be an isotropic regular quadratic space and let K and L be maximal lattices on V. Then there is a splitting V = U I W in which U is a hyperbolic plane and -
L
n U)
(L n W) ,
K = (K n U)
(K n W) .
Proof. Let L be a-maximal, let K be b-maximal. For each non-zero vector x in V we let a. denote the coefficient of x with respect to L and we let b. denote the coefficient of x with respect to K. We put r=
, rx = bxfax .
Now alf S L for some non-zero a in F, hence r.S a-lo; hence we can pick an isotropic vector x in V for which r. is maximal (among all the isotropic vectors of V). By Proposition 82:20 and Corollary 82:21a there is an isotropic vector y such that B (x, y) 1 and L
(axx + av y) 1 • • •
Since ax (ivy is a-maximal its scale ax at, must be equal to -I a by Proposition 82:21. Now bx x
by is contained in K, hence bby s-7, 2 b.
Chapter VIII. Quadratic Forms over Dedekind Domains
237
Therefore rx ry S t.
Using Proposition 82:20 and Corollary 82:21a again, we find an isotropic vector z such that K= (be z + b ey) ± • • • . This time we obtain I), bv =
hence rz ry D r. Therefore
and a,a,
r xr y Ç r Ç rz ry . But r , was chosen maximal. Hence rx rv = r and so
bb y = (a.ay) (try) = Therefore b xx Hence
by has scale
1
1
a bia = b.
and volume — 4 b2 ; so it is -l b-modular.
K (b z x ± be y) ± • • • L = (a x x + (iv y) 1 • • • q. e. d. Then U=Fx+Fy gives the desired splitting of V. § 821. The lattice La Consider a lattice L in the quadratic space V, and a fractional ideal a in I. Suppose that L is regular. We define La as the sublattice
1
La= Ix E LJB(x, L) Ç a} of L. For the trivial lattice we have La= 0, so let us assume that L is not 0. We immediately have B (La, L) S a , sLa S a , and La= L <4. 6L S a . It is easily seen that La = aL#n L. In particular this shows that FLa=FL. For any splitting L=J IK we have ± Kr= PI Ka. Now consider another fractional ideal b. Then ab-1 Lb S LaÇ Lb if a S b follows directly from the definitions. If L is b-modular, then IL if b S a La = ab -'L if b2a, since La=aL4*nL—ab-lLnL.
Part Four. Arithmetic Theory of Quadratic Forms over Rings
238
§ 82J. Scaling
Consider the lattice L in the quadratic space V. Let a be a non-zero scalar. Recall that 'Pc denotes the vector space V provided with a new bilinear form BOE (x, y) = a B (x, y). We shall use LOE to denote the lattice L when it is regarded as a lattice in Va. Thus the properties of LI as a lattice are identical to those of L; the superscript a merely emphasizes that our interest has shifted from the properties of B on L to those of B1 on L. We easily see that SLOE
(6 L) , n LOE = cc (n L) , t, LOE ---
(b L)
where r denotes the rank of L. If L is a-modular in V, then La is a amodular in Va. If L is a-maximal on V, then LŒ is cx a-maximal on Vcc. If Ka. So for a lattice L L K is a representation, then so is cp: LOE on the regular non-zero space V we have (L) ---- 0 (La) , 0+ (L) = 0+ (L a ),
and cis L cis
cls+L = cls+LŒ.
If L is a free lattice with matrix M, then LI is free with matrix cc M; so here d La = ccr (dL) .
Suppose L is regular and let a be a fractional ideal. Then the notation Lcc should not be confused with the notation L. For instance, Lao denotes a certain sublattice of L in the quadratic space V, while LI denotes the original lattice L in the quadratic space VI. It is easily verified that (LŒ) a = (Lc‘ - ' a)a . § 82K. Localization
Consider the lattice L in the quadratic space V. Let p be any spot in the Dedekind set of spots S on which our ideal theory is based. Consider the quadratic space VI, (i. e. the F;, -ification Fp V of the quadratic space V) and the localization Lp of L in Vp . Take a base x1 .. , x,. for FL in which L has the form ,
L a i xi + • • • + ar xr .
Then by § 81E Lp aip + • • • ± afp x,.
so by Proposition 82:8 we have sL p
E ai;, a1 , B (xi, xj) = (E aj ai B (xi, xj)) =
Similarly with norm and volume. In other words, sL p = (s
, n L p = (n L) ;, ,
= (»L)p
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
239
holds for all p in S. In particular two lattices L and K have the same scale (or norm or volume) if and only if Lp and Kp have the same scale (or norm or volume) at all p in S. From this it follows, using the definition of a modular lattice, that L ay-modular for all p in S. is a-modular if and only if L It is also true that L is a-maximal on V if and only if Lp is ap-maximal a y-maximal for all p in S. on Vp for all p in S. First suppose that L If L is not a-maximal, there is a lattice K on V such that KDL and rtK ç a. Then by § 81E there is a of in S at which K g D L g . But nKg ç a q . This denies the maximality of L g . Conversely let us be given an a-maximal lattice L on V, and let us suppose that L g is not aq-maximal at some spot of in S. Then there is a lattice J(q) on Vq which strictly contains L g and has n j (q) C aq . By Proposition 81: 14 there is a lattice K on V with if if
p C S—q p -- q .
Here we have KD L. But nK), C. ap for all p in S. Hence nK .Ç. a. This denies the maximality of L. Chapter IX
Integral Theory of Quadratic Forms over Local Fields This chapter classifies quadratic forms under integral equivalence over local fields'. We continue using the notation of the last chapter, except that the field F is now taken to be a local field, S is the single spot p, o is the ring of integers o (p), p is the maximal ideal m (p), and u is the group of units u (p). We let i 1-= I l y be a valuation, say the normalized valuation, in the spot p. The letter a denotes a prime element of F at p. Ideal theory and lattice theory are with respect to the single spot p. The quadratic space V provided with its symmetric bilinear form B and its associated quadratic form Q is assumed to be regular and nonzero. We shall consider non-zero regular lattices L, K, J, . . . in V. Note that every line Fx in FL contains a maximal vector of L since the class number of F at p is 1. And every non-zero lattice is free. In particular the discriminant dL of § 82B is available for use.
5 91.
Generalities
Here F can be dyadic or non-dyadic. Note that L has the following property: if we take any two vectors x, y in L for which 1B (x, 3))1 is ' For the local integral theory of hermitian forms see R.
Math. (1962), pp. 441--465.
j ACODOWITZ,
Am. J.
240
Part Four. Arithmetic Theory of Quadratic Forms over Rings
largest, then B (L, L) = B(x y) = L
Similarly if we take any vector x E L for which 1Q (x)I is largest, then Q (x) 0 = n L § 91A. Maximal lattices 91:1. Theorem. V is an anisotropic quadratic space over a local field, and L is an a-maximal lattice on V. Then L ={x E VIQ (x) Ea}. In particular, all a-maximal lattices on V are equal. Proof First let us show that the set X = {x E V IQ(x) E a} is closed under addition. Consider typical vectors x, y in X. Thus Q (x), Q (y) E a and we must show that Q (x + y) c a. Suppose not. Then 2 B (x, Y) = Q (x + y) - Q (x) - Q (y) a, hence 12 B (x, y)I > [al and so Q (x)Q (Y) B
I -I •
Now the discriminant dB (x, y) is equal to
—B (x, y) 2 11
Q (x) Q (Y)1 B(r y) 2
so —dB (x, y) is a square in F by the Local Square Theorem. So Fx Fy is a hyperbolic plane. This is impossible since V is given anisotropic. Hence X is indeed closed under addition. Hence it is an o-module. Suppose X is not equal to L. Then L C X by dëfinition of X, so we can pick z E X — L. Then L oz is a lattice contained in the o-module X, so Q (L o z) Q (X) Ç a, q. e. d. z) S a. This contradicts the maximality of L. hence n (L 91:2. Theorem. Let K anti L be a-maximal lattices on the regular quadratic space V over the local field F. Then cis K = cis L. Proof. We can suppose that V is isotropic by Theorem 91: 1. By Proposition 82:23 we have a splitting V Hi 1 • • • L. H,. ± Ho in which H1, , H,. are hyperbolic planes and 1/0 is either 0 or an anisotropic space, with L (L _L - - (L Hp) (L n Ho) , 1K (Kñ H 1) 1 • • • ± (Kñ H,.) j (K H 0) .
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
241
Now L r Ho and K r Ho are either both 0 or both a-maximal on Ho, hence they are equal by Theorem 91:1. And L r Hi and K r Hi are isometric for 1 i r by Example 82:22. So for 0 I r we have E 0 (II1 ) with a (L r H1) K r H1. Then —
a=a1±*••Ictriao
is an element of 0(V) vvith L = K. Hence K E cis L. Hence cis K = cIsL. q. e. d. 91:3. Example. If L is an a-maximal lattice on a regular quadratic space V over a local field F, then Q (V) r\ a Q (L). For clearly Q (V) n a Q (L). On the other hand if oc is a non-zero element of Q (V) n a, there is a vector x in V with Q (x) = cc E a. This vector x is contained in an a-maximal lattice M on V by Propositions 82:9 and 82:18. But cis L = cis M by Theorem 91:2. Hence a E Q (L) as required. In particular if dim V 4 we must have Q (L) = a since every regular quadratic space V with dim V 4 over a local field is universal by Remark 63:18. If dim V = 3 we can use Remark 63:18 to show that V represents either all units or all prime elements, hence for any # in a we have either Q (L) 9 u or Q (L) 2 flu. —
§ 91B. The group of units of a lattice
Consider the lattice L on the regular, non-zero quadratic space V over the local field F. Let u be a maximal anisotropic vector of L. We claim that the symmetry T.. of V is in 0(L) if and only if 2 B (u , L)
Q (u)
S0.
First suppose Zu is a unit of L. Then 1-ti L = L and so 2 B (u,
Q (u)
u
x—
xEL
for all x in L. But the coefficient of u is o since u is maximal in L, hence 2E (u, x)112 (u) E o as required. Conversely if this condition is satisfied it is dear that .ru L Ç L and so ru L = L by Example 82 :12, i. e. E 0 (L) as required. 91:4. Let L be a lattice on the regular non-zero quadratic space V over the local field F. Then 0 (L) contains a symmetry of V. Proof. Take u E L with Q (u) 0 = nL. Then u is clearly a maximal vector in L; and 2 B (u, L) C28L SnL = Q(u) o , so that 2B (u, L)/Q(u) Ç o; hence x„, is a unit of L. 91:4 a. (0(L): 0-E(L)) = 2 and so cis L = cls-FL. O'Meara, Inirodlielion to quadratic forms
q. e. d. 16
242
Part Four. Arithmetic Theory of Quadratic Forms over Rings
91:5. Let L be a maximal lattice on an anisotropic quadratic space V over a local field. Then 0(L) = 0(V). Proof. We must show that a typical a E 0(V) is in 0(L). Now by Theorem 91: 1 there is a fractional ideal a such that
L {x E V IQ (x) E a} .
For any x in L we have Q (a x) = Q (x) E a since a is an isometry, hence aL CL, hence a E 0(L). q. e. d. 91:6. Let V be a regular quadratic space overa local field with dimV 3. Then 0 (0 -E(V)) = F. Proof. If dim V
4 the space is universal by Remark 63:18, hence there is a symmetry with any preassigned spinor norm, hence there is a rotation with any preassigned spinor norm. We are left with the ternary case. By Remark 63:18 the space V will represent all a in F for
which VI <--oc>
has discriminant not equal to 1. So if the discriminant of V is a unit, V will represent all prime elements and at least one unit; now every element of F is a product of exactly two such elements times a square in F; hence 0(0-F(V)) F. A similar argument applies when the discriminant of V is a prime element. q. e. d. 91:7. Let V be a regular quadratic space over a local field with dimV 3. Then there is a lattice L on V with O (0 -E(L)) . F. Proof. If V is anisotropic we take any maximal lattice L on V. Then by Propositions 91:5 and 91:6 we have 0(04-(L)) = 0(0 4- (V))
.
Hence we may assume that V is isotropic. Take a base x, y, z, . . . for V in which 1
(ox
oy)
j_ (0.Z) j_ • • • .
Then for any unit e we have ey, L)20 C0 ' Q (x ey) = 2e
2B
hence the symmetry rx+81/ is a unit of L. Now this symmetry has spinor norm 2 E F2. Hence 0(0+ (L)) D uF2. Using the same argument with z instead of x ey we obtain a symmetry ; which is a unit of L and has
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
243
spinor norm 27t 1 ; then O(ux .i.„) = Emil; hence 0(0+(L)) D nub:2 and we are through. q. e. d. 91:8. L is a maximal lattice on a regular quadratic space V over a local field with dim V 3. Then 0(0+ (L)) D u. Proof. By scaling V we can assume that L is 20-maximal. If V is anisotropic, then 0+ (L) = 0+(V) by Proposition 91:5 and so 0(0±(L)) F D u by Proposition 91:6. Now suppose that V is isotropic. By the results of § 82H there is a splitting L = U I. • • • in which U has a base U = ox + ey with matrix CI 01) . Consider any e in u. Then ey,
2 B (.7;
20
2e
+
0
hence the symmetry Toe+ , is a unit of L. Now this symmetry has spinor norm 2eP2. Hence 0 (0+ (L)) D UF2 q. e. d. § 91C. Jordan splittings
Consider a non-zero regular lattice L in the quadratic space V. We claim that L splits into 1- and 2-dimensional modular lattices. If there is an x in L with Q(x) o = L. then J = ox is an 6L-modular sublattice of L. Otherwise Q(x) o ( 8L for all x in L; in this event we can find a binary sL-modular sublattice j of L: we pick x, y in L with B(x,y)o=sL; then the vectors x, y have discriminant d (x, y) = Q (x) Q (y) — (z, y) 2 and this is not zero by the Principle of Domination; hence J = o z + 0y is a binary lattice; a direct computation shows that J
B (x, y)
sL , J= B (x, y) 2o = (54 2 , so that J is actually a binary 5L-modular sublattice of L. Hence L always contains a 1- or 2-dimensional 6L-modular sublattice J. Then L has a splitting L=JIK by Corollary 82: 15a. Now repeat the argument on K, etc. Ultimately we obtain a splitting of L into 1- or 2-dimensional modular components. This establishes our claim.
If we group the modular components of the above splitting in a suitable way we find that L has a splitting L
_L • • • _L L i
in which each component is modular and 64. • • • )54 Any such splitting is called a Jordan splitting of L. We have therefore proved that every non-zero regular lattice L in a quadtatic space V over
a local field F has at least one Jordan splitting. The rest of this chapter is really a study of the extent of the uniqueness of the Jordan splittings of L. 16*
244
Part Four. Arithmetic Theory of Quadratic Forms over Rings
We shall need the lattice La of § 821. First suppose that L is b-modular. If b = a, then La= L and so L a is a-modular. If b C a, then La= L is b-modular with b C a. And if b D a, then La = a b-1 L is a2 b-4-modular with a2b--' c a. So for a b-modular lattice L we have the following result: La is a-modular if and only if a = b, otherwise it is c-modular with c C a. What does this mean in general? Let us consider any non-zero regular lattice L in the Jordan splitting
L = L1 I • • • 1 L i .
We know that sLa S a. And we have a splitting L a DILL • • • ± Ltt' into modular components. If L i is not a-modular, then L will be cmodular with c c a. And if L i is a-modular, then Lf will be equal to L. Hence we find that s La = a if and only if there is an a-modular component in the given Jordan splitting. Otherwise &La C a. Now consider L 1 in the given splitting, and suppose it is a-modular. If we group the components of the splitting LI ± • • ± .14` we obtain a Jordan splitting of La in which the first component is Li. So we have proved that an a-modular component in a given Jordan splitting of the lattice L is the first component in some Jordan splitting of the lattice La. 91:9. Theorem. Let L be a lattice in the quadratic space V over the local field F, and let L = L1 1,•• • ± L i , L = ± • • • ± Kr
t we have be two Jordan splittings of L. Then t = T. And /or 1 A (1) 5 LA = 8K2, dim LA = dim Ka, (2) nLa sLa if and only if nKA &KA . Proof. We shall use the results of the preceding discussion in the proof. 1) Suppose there is an a-modular component in the first Jordan splitting. Then sLa a. Hence there is an a-modular component in the second splitting. Hence, on grounds of symmetry, we have t = T and 2, t. 8L 2 = sICA for 1 Consider a typical A with 1 A I. We must prove that dim LA = dim KA and that n LA=- 841 if and only if nlça = 6 K2. Now LA and KA are first components in Jordan splittings of La where a = sLA = sKA . Hence we can assume that A . 1. And by suitably scaling the bilinear form B on V we can assume that sL o. On grounds of symmetry we are therefore reduced to proving the following: given 54— 6 K1 — s L=0 , prove that dim././ dimKi , and also that n.Li = n implies niC, = 0. This we now do. 2) We shall need the projection map 99: FL„--›-FK„. This is defined as follows: each x in FL, is an element of FL and hence has a unique -
-
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
245
expression x—y+z with y C FKI and z C F(K2 1_ • _L Kg). Put ipx y. Then Tx is a well-defined element of FK1 , and it is easily seen FK1 determined by x --->— 99X is F-linear. that the map T: FL, Now when x in L 1 is expressed as z y + z in the above way we have y E K1 and z(K 2 I''• L K t . Hence 99L1 c K 1 . And for all )c, x' in L1 we have B (Tx, T xi) = B (x — z, z') B (x, x') modp since B (K2 • • • K t , L) p In other words B (99 x Tx') — B (x, z'), Q( T x) Q (x) modulo p for all x, x' in L1 . Suppose we had Tx = 0 for a non-zero vector x of FL,. Then we would have cp-x — 0 for a maximal vector x of the lattice L1 . Since L 1 is unimodular there is a vector y in L1 with B(x, y) 1 by Proposition
82:17. Then 1
B(x, y)
B (Tx, Ty) — 0 mod p
,
and this is absurd. Hence 99 is an isomorphism of FL, into FKi . Hence
dim L1 = dimFLI dimFK1 — dimKi , as required. 3) Finally we have to prove that nK/ — if nL1 = v. Since nil find a vector x in L 1 with Q (x) s where e is in u. Then wecan
Q( T x) Q (x) e modp Hence Q( T x) C u. Hence nKI D o. Hence nKi q. e. d. Consider non-zero regular lattices L and K in quadratic spaces V and U over the same field F. Let
L=
_L • • L t , K =
• - • 1. KT
be Jordan splittings of L and K. We say that these _Jordan splittings are of the same type if t = T and, whenever 1 t, we have
eLA 6K2 ,
dimL2 = dim KA
and !IL A --= 6L 2
if and only if
n KA
= 6 Kz
.
We know from Theorem 91:9 that any two Jordan splittings of L are of the same type. And the same with K. We say that the lattices L and K are of the same Jordan type if their Jordan splittings are of the same type. Isometric lattices are of the same Jordan type. 91:10. Notation. Given a lattice L and a Jordan splitting L -•• L I , we put
L (i)
L1
--
Li
246
Part Four. Arithmetic Theory of Quadratic Forms over Rings
and we call L(1) C L(2) C • • • CL()
the Jordan chain associated with the given splitting. We put = L.
•••_L
and we call LA ) D L r2) D • • • D / (t )
the inverse Jordan chain associated with the given splitting. Clearly L = L( 1) L 4 4. 1) .
A Jordan chain is determined by one and only one Jordan splitting of L. 91:11. Example. Let L(1) C • L. (t) and K(1) C» • C K(0 be Jordan chains of lattices L and K of the same Jordan type. Then L (i) and Km are lattices of the same Jordan type. Also )3 L(0 = K. Hence there is a unit s such that d LtOld K = S.
The same applies to inverse Jordan chains of L and K. § 92. Classification of lattices over non - dyadic fields Throughout this paragraph we assume that the local field under discussion is non-dyadic. We consider a non-zero regular lattice L in the quadratic space V. We know from § 82E that 2 el. Ç nL S EL. But 2 is a unit in o since F is non-dyadic. Hence nL = eL, i. e. norm and scale are equal over non-dyadic fields. We can therefore pick x E L with Q(x) o = ttL = 5 L . Then J = ox is an EL-modular sublattice of L, hence L has a splitting L =J _L K. If we repeat on K, etc. we ultimately find a splitting L = oxi _L • • • 1, o x, . In other words, in the non-dyadic case every non-zero regular lattice has an orthogonal base. 92: 1. Let L be a unimodular lattice on the quadratic space V over the non-dyadic local field F. Then there is a unit 8 such that L.r. <1> ••• ± <1> _L <E> .
Proof. Since L has an orthogonal base we can write L 2-4 <8 1 > _L • • • I <811> (Si E u) . Put s = e en . Then <1> • • • <1> ± <8> by the criterion of Theorem 63:20 in virtue of the fact that the Hilbert symbol `3' ) is 1 whenever 6, 6' are units in a non-dyadic local field. FL
(-!—
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
247
So there is a lattice K over V with K <1> 1 • • • 1 <1> J.. <e>. Now K and L are 0-maximal on V by Proposition 82:19, hence they are isometric by Theorem 91:2. q. e. d. 92 : 1 a. There are essentially two unim,odular lattices of given dimension over a given non-dyadic local field. 92: lb. Q (L) u if dimL = 2. And Q (L) = o if dimL 3. Proof. If dimL = 2 we apply the proposition twice to obtain
L <1> I <e>
Therefore d B (47) . . . , x,.) is a unit. Hence D (974) 9)L1 = K1 by Corollary 82:11a. So K1 has the base
o
—» Kr So
0 (ço xi) + • • • + 0 (99 x,.) . 1
O. T.
A similar theorem holds for representations in the non-dyadic case. See O'MEARA, Am. J. Math. (1958), p. 850.
248
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Hence dK, s (1 + a) with cc in p. Hence dKi s by the Local Square Theorem. Hence dK, = 2) Conversely suppose that dL i = dfCi for 1 i t. Then it follows easily from Proposition 92: 1 that there is an isometry ai of FLi onto FICI such that o•i L i = Ki . Put = • • • I at . Then a is an element of 0 (V) such that cf. = K. Hence K E cIsL. Hence q. e. d. clsK = clsL t. 92:2a. Corollary. cIsL = clsK if and only if FLi F Ki for 1 92: 2b. Corollary. Let L(, ) C» • • C L(0 and K(1) C • • • C Kw be the Jordan chains associated with the given splittings of L and K. Then clsL clsK if and only F K(i) for 1 I t. 92:3. Theorem. Let L and K be isometric lattices on the regular quadratic spaces V and U over the non-dyadic local field F. Suppose there are splittings L = L' I. L" and K = K' J. K" with L' isometric to K'. Then L" is
isometric to K". Proof. This is an easy application of Theorem 92:2 and is left as an q. e. d. exercise to the reader,
92:4. Theorem. Let L be a lattice on a regular n-ary quadratic space over a non-dyadic local field. Then every element of 0 (L) is a product of at most 2n —1 symmetries in 0 (L). Proof. Let V be the quadratic space in question. The proof is by induction to n. The case n = 1 is trivial', so we assume that n> 1. By suitably scaling V we can assume that sL = 0. Consider a typical ci in O (L) . This a must be expressed as a product of symmetries in 0 (L). Fix y in L with Q (y) equal to some unit e. Then (31 - r + + crY) = 48 and this is a unit since the field is non-dyadic. Hence either Q(y — cry) or Q (y cry) is a unit. In the first instance the criterion of § 91B shows that the symmetry r , , is a unit of L, and we have Tv–avY =
crY •
In the second instance ri, and ry +„ are elements of 0(L), and we have r11+citf
TO/
cry.
So in either case there is an element p which is a product of one or two symmetries in 0(L) such that py = ci y. By Corollary 82:15a we have a splitting L = oy I K. Then p-la induces an isometry on FK; this isometry is a unit of K and hence by the inductive hypothesis it is a J. Starting the induction at n = 2 will show that at most 2n —2 symmetries are needed when n 2.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
249
product of at most 2n — 3 symmetries in 0 (K). Then g-lcr = (1 ± xi) (1 I r2) . .
Hence a=
e ( 1 I Ti) ( 1 ±x2 ) • •
is a product of at most 2n — 1 symmetries in 0(L). q. e. d. 92 : 5. Let L be a modular lattice on a quadratic space over a non-dyadic local field with dimL
2. Then 0(0+ (L)) = uF.2 . Proof. By suitably scaling the space we can assume without loss of generality that L is unimodular. Take a typical symmetry in 0 (L). This symmetry has the form Ty with y a maximal anisotropic vector in L. Since r is a unit of L we have B (y, L) Ç Q (y) 0 by the criterion of 0 since L is unimodular. Hence Q (y) E u. So § 91B. But B (y , L) O (r) Ç 14'2. Therefore (0± (L))
We then obtain equality here by applying Corollary 92: lb to find a symmetry -t- in 0 (L) with 0 (T) = EF2 for any e in u. q. e. d. 92:6. Example. Let L be a lattice on the regular quadratic space V over the non-dyadic local field F. Consider a maximal anisotropic vector y of L. The criterion of § 91B says that the symmetry is a unit of L if and only if B (oy, L) Ç Q (y) o. By Proposition 82:15 we know that this condition is equivalent to saying that oy splits L since v y is Q (y)modular. Hence for any maximal anisotropie vector y in L we have Tv
E
(L)
oy splits L.
92:7. Example. Suppose V is a regular quadratic space over a nondyadic local field with invariants dV = 1 and S V = 1. We claim that there is a lattice L on V with 0 (0 (L)) =
By Theorem 63:20 we have V=-,- <1> ± • ± <1> .
Hence there is a lattice L on V with Lc <1> j. <2>
...
< 212 (n-1)>,
By Theorem 92:2 we know that if oy splits L, then Q (y) ( 21 U2 for some i (0 i n 1). Hence by Example 92:6 all symmetries in O (L) have spinor norm .P 2. Hence by Theorem 92:4 all elements of 0(L) have spinor norm F 2.
Part Four. Arithmetic Theory of Quadratic Forms over Rings
250
92:8. Example. Suppose V satisfies the conditions of Example 92:7, and suppose further that dim V is even. Let e be any unit in o. Then there is a lattice L on V with 0(0- (L))
e 1 "2
.
92:9. Example. Let L be a lattice on a regular n-ary quadratic space
over a non-dyadic local field, with 1
8L (O
n(n-1)
and bLpr
We claim that 0(0.1- (L)) u .
This follows immediately if we can prove that the number t in the Jordan splitting L= I • • • _L L t
is less than n: for then we will have dim L i 2 for at least one 1(1 i t) and we can apply Proposition 92:5 to this L i . Suppose if possible that = n. Then b Li
= sL i S
,
hence „ -2 n(n-1) —
and this is contrary to hypothesis.
§ 93. Classification of lattices over dyadic fields Throughout this paragraph F is a dyadic local field. Thus F has characteristic 0 2 and the residue class field of F at p is a finite field of characteristic 2. So the residue class field is perfect, and the congruence e'
ei52 modn
has a solution i5 for any given units e and e'. For any given a, 13 in 1 we shall write ocfi
if 413 is an element of u2. This defines an equivalence relation. And a 13 if and only if affl is a unit with quadratic defect b (4/3) = 0. For any fractional ideal a we write =.-.P 13 mod a if a/fl is a unit and cc = 13e2 moda for some unit e. This also defines an equivalence relation. And a fi moda if and only if al 13 is a unit with 1 See C. H. SA11, Am. J. Math. (1960), pp. 812-830, for the integral theory over local fields of characteristic 2.
Chapter 1X. Integr -.-1 Theory of Quadratic Forms over Local Fields
251
b (a/fl) Ç a/ 13. In virtue of the perfectness of the residue .class field we have cc 13 mod a p when ai i3 E u. The letter A will denote a fixed unit of quadratic defect 4o. It is assumed that 4 has the form A —1 + 4o for some fixed unit e in F. 1 mod40. Of course A V will be a regular n-ary quadratic space over F, L will be a non-zero regular lattice in V. As usual 2(6L) Ç nL sL But we now have 2(5L) Cs/. , and it is this strict inclusion that makes the dyadic theory of quadratic forms distinctly different from the non-dyadic theory of the last paragraph. 93:1. Notation. We let o 2 denote the set of squares of elements of 0. The symbol 0 2 already has a meaning in the sense of ideal theory, namely oi2 is the product of the fractional ideal o with itself. However this product is equal to o so there is never any need to use the symbol 0 2 in this sense. For us then 0 2 will be the set 02
cx2
0}
93:2. Notation. Given a non-zero scalar a and a fractional ideal a we shall write a mod a dL a mod a. This is the same if dL = 13 for some /3 E F which satisfies 13 in which as saying that L has a base x1 , . .
dB
a mod a .
. . , xn)
Given two lattices L and K we write
dLidK
a mod a
if there are non-zero scalars 13, y such that dL = 13, dK = y, and , #/y a mod a. This is equivalent to saying that there are bases and y•• ym for L and K such that
dB (xj., . . . , x n)IdB (yi, . .
ym )
a mod a .
§ 93A. The norm group gL and the weight tvii It is easily seen that the set Q (L) + 2(6 L) is an additive subgroup ofF. We shall call this subgroup the norm group of L and we shall write it
L Q (L) + 2 (6 L) The norm group gL is a finer object than the norm ideal nL which it generates. We have 2(6L) qL ( nL ,
252
Part Four. Arithmetic Theory of Quadratic Forms over Rings
and nL = ao holds for any element a of gL with largest value. Given two regular non-zero lattices K and L in V, then K L gK gL , and g(K L) = gK + gL . The sets Q(L) - and gL don't have to be equal. For instance if L = o x, then Q (L) = 0 2 Q (x) contains no fractional ideals and so it cannot be all of gL. On the other hand, if L is any lattice in V with L o and if L 0 contains a sublattice H of the form H ( 1 0) then we do indeed have gL = Q (L) . In order to prove this we must show that a typical element Q (x) 1 2 (x EL, a ( 0) of g L is also in Q (L). Now we have a splitting L = H j . K since H is unimodular and sL = 0, hence we can write x h k with liEH,kE K. Since Q (H) =--- 20 we can find h' E H with Q (h') = Q (h) + 2a. Then - -
Q (151 + k) Q (h') + Q (k) = Q (h) + 2 oc Q (k) = Q (x) + 2oe . Hence Q (x) + 2 cc is in Q (L) as required. So Q (L) and gL are sometimes equal. As a matter of fact the result that we have just proved will enable us to arrange gL = Q(L) whenever we please. 93 : 3. Let L be a lattice of scale 0 on a hyperbolic space V over a dyadic local field. Then there is a unimodular lattice K on V with LSKSV such that gK gL. Proof. As we ascend a tower of lattices on V we obtain an ascending tower of volumes in F. Hence we can find a lattice K on V with LSK, sK=o, gK=gL,
and such that there is no lattice K' on V which has these properties and strictly contains K. This K will be the lattice required by the proposition. We assert that every isotropic vector x in V which satisfies B (x, K) So is actually in K. For consider K' ox + K. Then sK' = 0 and Q (K') Q (K) + 20 = gK, so that
L S K' , 8K 1 = o , gK' = gL Hence K' = K by choice of K. Therefore x is in K as asserted. Hence every maximal isotropic vector x of K satisfies the equation B (x, K) = o. Pick y E K with B (x, y) = 1. Then ox + oy is a unimodular sublattice of K, so we have a splitting K =- x oy) j.. J. But F J is isotropic since F K is hyperbolic. Hence J has a splitting J = (o x' + o y')1_ J' with x' isotropic and ox' oy' unimodular. Repeat. Ultimately we
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
253
obtain a splitting of K into binary unimodular lattices. So K is uniq. e. d. modular. 'We let mL denote the largest fractional ideal contained in the norm group g L. Thus 2(sL) SmL S gL SnL.
Let us show that is even. Suppose not. Pick a E gL with al) = nL and write mL = a Or + with an r O. We claim that ap2r S gL. We have to show that any element of the form a s 7L2r with e E u is in g L. By the perfectness of the residue class field we can solve the congruence ordp nL ordp mL
a 8 7c 2r
a 62 7g2r mod
a p2r +1
for some unit 6. But a p2r +1 Ç gL by definition of m L. And a (a a ) 2 is in g L. Hence ac 7c2r is in g L. So ordp nL ordp mL is even. We define the weight toL by the equation ti, L = p (m L) ± 2 (s L) . So mL depends only on gL, while troL depends on L. We have mL ,
p(mL)
2(sL) cwLnL
.
Also ordp nL ordp tvL
is odd
tr•L = p (mL)
and ordp nL ordp toL
is even = toL = mL = 2 (sL) .
Hence ito L = n L =
2 (s L)
YLL
It is easily seen that KL=
tr•IC SID/.
and that
gK =gL with sK=sL => toK=toL. We call the scalar a a norm generator of L if it is a scalar of largest value in gL. Thus a is a norm generator of L if and only if aEgLSao, j. e. if and only if a E gL with ao = nL We call the scalar b a weight generator of L if it is a scalar of largest value in to L. Thus b is a weight generator of L if and only if bo = wL . If a and b are norm and weight generators of L, then 1b1 < 154 , 1b1 lai , 12(L)1 ordp a + ordp b is even =1b1 = 12(641 lbl = lai <=> la' .12(5L)I .
254
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Let us prove that every element of gL whose order has opposite parity to the order of nL is in to L. In other words, if a is a norm generator of L, and if fi E gL with ordp a + ordp 46 odd, then /3 E ito.L. It is enough to prove that fib S g L, for then flo S m L by definition of mL, hence S p (mL) S toL. Consider a typical non-zero y in -46o. We have to show that y E g L. Now by Proposition 63: 11 we have E F such that + /3 772 mod4 y. By the Principle of Domination both and 12 are in 0, since y E /3 o S ao. But gL stands multiplication by elements of 02, by definition of g L. And 4 y oLS 4 (nL) S 2 (s L). Hence y E g L, as asserted. 93:4. Let L be a non-zero regular lattice over a dyadic local field with norm and weight generators a and b. Then
gL a02 bo Proof. The set gL stands multiplication by elements of o2, hence ao2 ç g L, hence ao2 roL is contained in g L. Conversely consider a typical element oc of g L. We wish to express a as an element of a 0 2 + to L. By Proposition 63: ii we have scalars and 21 such that a + a n 272 mod4 . By the Principle of Domination we see that E 0, since a E ao. Then oc and aV are in g L, and 4 a o S 2 (sL) S g L, hence an n 2 in in gL, hence it is in to L since its order has opposite parity to the order of nL. q. e. d. 93:5. Example. Let us give a general method for finding norm and weight generators for L (computable methods will be given later in § 94). To find a norm generator simply take any element a of largest value in Q (L). To find a weight generator first take 1,0 E Q (L) of largest value
such that
ordi, a + ordr,b0 is odd.
If boo D 2 (s 4, then to L = boo and bo is a weight generator of L. If boo ç 2 (s L), or if bo does not exist, then to L = 2(s L) and we can take any b for which bo = 2 (s L). 93:6. Example. Let a be a norm generator of L, and let a' be some other scalar. Then a' is a norm generator of L if and only if a a' mod toL. 93:7. Example. Consider a non-zero scalar oc. Then g (OEL) -= (x 2 (g L). in (cc L) cc 2 (m L) , and to (cc L) = a2 01,4 If a is a norm generator of L, then a2 a is a norm generator of (xL. 93:8. Example. What happens when we scale the quadratic space by a non-zero scalar oc? We obtain g (D‘) = oc( L), in (D') = cc (m L) , and to (Di) = oc (to L) . liais a norm generator of L, then ct a is a norm generator of LI. If oc = 1 we obtain gL-i = gL , = tilL , tvL - 1 ttIFL , —
and a is a norm generator for L --1 as well as for L.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
255
§ 93B. The matrix A (x, /3) We shall use the symbol A (cc, 13) to denote the 2 x 2 matrix A (a, fl) =
1,3) 1
whenever a, fi are scalars which satisfy the conditions
ac,flEo,
—I -1-afiEu.
These conditions simply guarantee that the matrix A (of, 13) is unimodular. Whenever the symbol A (a, fi) appears it will be understood that a, fi satisfy the above conditions, even if this is not explicitly stated at the time. We use A (a, fl) to denote ordinary multiplication of the matrix A (oc, 13) by the scalar Thus
93:9. Example. If L A (a, 0), then L A (a+2 f3, 0) for any )3 in o. 'For if we take a base L ox oy in which L has the matrix A (oc, 0), then L = o (x + fly) 4- o y also gives a base for L and the matrix of L in this base is A (a+2fi , 0). 93:10. Example. Let L be a binary unimodular lattice and let a be a norm generator of L that is also in Q (L) . We claim that (i) L A (a, fi) for some /3 E tvL, (ii) if n, 2o then the /3 in the above matrix for L is a weight generator of L. To prove this we pick any x c L with Q (x) — a. Then x is a maximal vector in L, so there is a vector y in L with L= ox io y . Now a is a norm generator of L, hence by Proposition 93:4 there is a in o and an a/ in tr•L with Q (y) Then L ex (y ± x) and ?I .
—
71 ) + (va) n B (y , x) E tvL Q (y + x) — (a But mi.( sL o, hence B (x , y ± x) is a unit since L is unimodular. Put z = (y x)1B (x, y + • x). Then L has a matrix of the desired type in the base L =ox + oz. This proves the first part of our contention. Now the second part. Here we are given tvL D 2e. We recall from Example 93:5 that there is a number b E Q (L) such that be = tvi, with ordp a ± ordi, b odd. Then ibl, and 12f < ibl, and
Ifli
b = a42 -I- n
+ #272
with E o. If we had fil < 1bl we would have Ibi = laVi by the Principle of Domination, and this is absurd since ordr a ordp b is odd. Hence lb!, so Iv", fie and f3 is a weight generator of L. 93:11. Example. Let L be a binary unimodular lattice with nL Ç 2o. We say that L A (0, 0) or L A (2, 2o) .
im
We know that L is 2o-maximal by Proposition 82: 19. So if FL is isotropic we will have L A (0, 0) by Example 82:22. Now suppose that
256
Part Four. Arithmetic Theory of Quadratic Forms over Rings
A (2a, 2 /3) by Example 93:10, and both FL is not isotropic. Then L cc and /3 will have to be units since otherwise we would have d (FL) =—1
by the Local Square Theorem. Construct a lattice K on a quadratic space F K with K A (2, 2e). Then a direct calculation of Hasse symbols shows that Sp (FL) = S (FK), hence FL and F K are isometric by Theorem 63:20, hence L and K are isometric by Theorem 91:2, hence L
A (2, 2e). 9 3 : 12. Let L be a lattice on a regular quadratic space V over a dyadic local field. Suppose that L has a splitting L = J I K and that J has a base J = x oy in which J A (a, 0) with a E 0. Put
r
+ +0 y
with
zE
Then there is a splitting L = J' K'. And K' is isometric to K. Proof. J' is unimodular and B (J', L) c o since z c K°. Hence we have a splitting L = I K' by Proposition 82:15. For any u EFK we define
Tu=u
B (u, z) y Then Q (9)u) = Q (u) since Q (y) = 0 and B (u, y) 0, hence 97: F K -4– V is a representation; but FK is regular; hence 99: FK>--V is an —
isometry by Proposition 42:7. Now B (9) u , x z) = B(u — B (u, z) y, x z) = B (u, z) — B (u, z) B (y, x) = 0. And similarly B (gm, y) = 0. Hence B(9) (F K), F J') = 0 and we have FK'. an isometry ep: FK Now B(u,z) E o whenever u E K since z E K°. Hence 9).K S L. Hence S F K' r L = K'. But 99 preserves volumes since it is an isometry.. q. e. d. Hence 99K K'. So we have found an isometry of K onto K'. 93:13. Example. Suppose the lattice L on the regular quadratic space V has a splitting L A (o 0)) K
a
with
E F and
cc E 0. Let /3 be any element of gi.K4 °. Then L<M(a+ 6-1 13,0)> _LK.
This is obtained from Proposition 93:12 by scaling. In particular, if L has a splitting L
K
with oc E o and sK S o, then for any fl in Q (K). § 93C. Two cancellation laws 93:14. Theorem. A lattice L on a regular quadratic space over a dyadic local field has splittings L= JIK and L =J' J. K' with J isometric to J' and J A (0, 0). Then K is isometric to K'.'
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
257
Proof. 1) Take bases J == ox + y and J' =ox' + Dy' in which A (0, 0) and j' A (0, 0). Note that j and j' are contained in L°. Hence j + j' L° and B( j, j') o. 2) First we do the special case where x= x'. We can express y' in the form y' = ocx + y + z (oc E 0, z since y' is in L =J j K and B (x, y') = 1. Then z = y' — ocx y is in (J + J') K Ç L° r K, hence z is in K°. We can therefore apply Proposition 93:12 to the sublattices =ox±o(y -Fax) J' =ox + o (y + cxx + z) of L. This gives us an isometry of K' onto K. 3) Next we do the case where B (J, J') = o. We can suppose that B (x, y'), say, is a unit. Making a slight change to x', y' allows us to assume that B(x, y') = 1. Put J". ox + Dy'. Then ra.-- A (0, 0) is a unimodular sublattice of J J' S L*, hence there is a splitting L = J" ±K". But we can apply the special result of step 2) to J and J", and also to J" and J'. Hence K K" and K" K'. Hence K K'. 4) Finally we consider B (J, J') S p. Here we put j"= ox +0 (y+y'). Then J" is a unimodular sublattice of L° with n J" S 2o. Hence J" A (0, 0) by Example 93:11. And we have a splitting L J" j K". But here B (J, J") o and B (J", J') = 0. Hence K K" and K" K' by step 3). So K K'. q. e. d. By a hyperbolic adjunction to a lattice L on a quadratic space we mean the adjunction of a lattice J of the form
T
al A (0, > 1 • • • _L
(0, 0))
E -fr)
If = a for each i, then J is a-modular and we call the adjunction of J an a-modular hyperbolic adjunction. 93: 14 a. Corollary. A lattice L on a regular quadratic space over a dyadic local field has splittings L = J J K and L J j K1 with J isometric to J1. Suppose that J is a-modular with ..
giSg Ka and gJ1 ÇgK. Then K is isometric to
Proof. By scaling we can assume that a --- 0. Adjoining j-1 to L shows that we may assume, without loss of generality, that J and L. are isometric unimodular lattices on hyperbolic spaces with gjSgK° and g S gig'. Now by Proposition 82:17 a unimodular lattice on a hyperbolic space has the form (oci, 0)) ± • - • 1
17
258
Part Four. Arithmetic Theory of Quadratic Forms over Rings
hence we have a splitting L
(A (cri, 0)) ± • • • ± (A (a,.,0)) j_ K
in which the ai are in g f and hence in g K°. Successive application of Example 93:13 now gives L
(21(0, 0)) ± • • • ± (A (0, 0)) K
Similarly (A (0, 0)) j • • • ± (A (0, 0) > j. K1 .
q. e. d. K1 by Theorem 93:14. The general cancellation law for lattices on quadratic spaces over non-dyadic fields (Theorem 92:3) does not extend to the dyadic case. For instance there is an isometry (A (0,0)) ± (1 ) (A (1,0)) ± <1>
Then K
by Example 93:13; but (A (0,0)) and (A (1,0)> are not isometric since their norms are not equal.
§ 93D. Unimodular lattices 93:15. A unimodular lattice L in a quadratic space over a dyadic local field has an orthogonal base if and only if nL = sL. If nL c sL, then L is an orthogonal sum of binary sublattices. Proof. If L has an orthogonal base it contains a 1-dimensional unimodular lattice; any such lattice has norm o; hence nL= o = sL. Conversely let us suppose that nL =sL=o. Then there is an xinL with Q (x) in u. The lattice ox is a unimodular sublattice of L and therefore splits L. Hence we have a splitting <4> 1_ • • • 1
<Er> ±
(oh, 181)> .L • ' • 1 (A (oct,
13t) >
in which r 1 and t 0, by § 9IC and Example 93:10. Consider the 3-dimensional sublattice K (8) I (A (a, (I)). If neither or nor 18 is a unit, then A (a ± e, fl) is unimodular and we obtain a new splitting K <e'> j (A (a ± e, 13)) in which a + e is a unit.We may therefore assume that cc, say, is a unit.Then C
(8)
<Œ>
C•r5 >
By successively applying the 3-dimensional case to L we ultimately obtain (el ) as required.
• • • _L
<en>
To prove the last part we take a splitting of L into 1- and 2-dimensional components. If a 1-dimensional component appeared in this splitting it would be unimodular and hence would have norm 0. This is impossible since nLc sL = o. Hence L splits into binary components.
q. e. d.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
259
9 3:16. Theorem. Let L and K be unimodular lattices on the same quadratic space over a dyadic local field. Then clsL = clsK if and only if g L g K. 1 Proof. By making the same unimodular hyperbolic adjunction to L and K we can assume, in virtue of Theorem 93:14 and § 93A, that Q (L) = gL = gK = Q (K). Now adjoin the lattice L ± L -4 to each of L and K, and let the resulting lattices be denoted L' and K' respectively. So =gL=gK. Hence by Corollary 93: 14a it will be enough Now g (L 1 to prove that L' is isometric to K'. But the component L --1.1 K of K' is a unimodular lattice on a hyperbolic space, hence K' has a splitting K' = L
(A (al, 0) )1 • • • j (A (cc„., 0))
with all ai in gL = Q (L). Then L L (A (0, 0)) ± • • • j. (A (0, 0)) by Example 93:13. Similarly we find K'
L'
L j.. (A (0, 0)) 1 • • • 1 CA (0, 0)) • q. e. d. Hence K' L'. Hence K L. In other words, clsK = cls L. 9 3:17. Example. Let L be a binary unimodular lattice on a quadratic space over a dyadic local field, let a be a norm generator of L that is in Q (L), let b be a weight generator of L, let the discriminant dL be written in the form dL = — (1 ± oc) with b (1 + oc) = oco. We claim that aa--1 (bo
and
L
A (a, —cca-1).
By Example 93:10 we have L A (a, b2) with 2 E 0, hence —(1 — ab 2) = —62 (1 ± oc) with e E u, hence
txo = b (1 + a) = b (1 — ab S ab So oc a-1 E bo. We must prove that L K where K is a lattice with K A (a, —oca-1). Now , (—a(1 FL (a) and a is a norm generator of both L and K, hence by Theorem 93:16 it suffices to prove that wL = wK. If toL = 2o, then oca-1 E 2o S roK and so tr•K = 2o by Example 93:10. If w L )20, then A is a unit by Example 93:10. We have b(1 — ab 2) = ab Ao since here ordp a + ordp b is odd with 'al p > iblp > 121 r, and so oca-10 = bo. Hence wK aa --1 0 20. Hence toK = cca-lo by Example 93:10. So wK = la- as required. Hence clsL clsK if and only if Q (L) = Q (K), i. e. if and only if L and K represent the same numbers. 1
17*
260
Part Four. Arithmetic Theory of Quadratic Forms over Rings
93:18. Example. (i) Let us describe the unimodular lattices of dimension 3. We consider a unimodular lattice L on a quadratic
space V over a dyadic local field. Let a be any norm generator, let b be any weight generator, let d be the discriminant of L. Regard d as an element of u. Note that bo = 2o when ordp a ordp b is even, by § 93A. (ii) First we dispose of all cases with dim L 3 and ordp a ± ordp b even. We claim that (A (0-, 0)) ± ••• .
To see this we take a splitting L—J±K in which / is binary and nK = nL. Let ai be an element of Q (K) such that aio = nK. Then al a norm generator of L and so gL = a1o2 ± 2o. Hence by Example is 93:10 we have a splitting L (a1 61 271, 2 ) )1 K (A (27?', 2C)) 1. K' with 77, c, n P in o. But K' represents an element a2 such that a2o = nL, hence it represents an element of the form 28 with e a unit, hence by the perfectness of the residue class field we can write L
with n in o. Hence by Example 93:11, "
L
(iii) Next we consider dimL = 4 with ordp a ordp b odd. Here we suppose that d has been expressed in the form d = 1 + cc with b (d) = oco. It follows from Example 93:10 that we must have a Eab 0. Take lattices J, K on quadratic spaces over F with .1 (A (b, 0)) K (A (a, — (oc — 4e) a-4)) 1 (A (b, 4e b-')>. Let us prove that L is isometric to J or to K, but not to both. First we note that every number represented by J is in ao2 bo g L, hence g/ C gL; but a is a norm generator of J and J represents the number b with ordp a ordp b odd, so b E wJ , hence g j = ao 2 to f 2 g L . Similarly with K. Hence we have proved that g/=gL.gK.
It is easily checked that di = dL dK
If we compute Hasse symbols using the rules of § 63B we obtain Sp (F j) — S (F . This shows first of all that L cannot be isometric to both J and K.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
261
Secondly, it shows that FL is isometric to one of F j or F K, by Theorem 63:20, hence L is isometric to j or K by Theorem 93:16. (iv) Now let dimL = 3 with ord p a 4- ord p b odd. We claim tht
L
(A (b, 0)) ± (— d>
or
L
(A (b , 4 b-'))
(— d (1 — 4 ) > ,
but not both. This can be derived from the 4-dimensional case. Take a unit e in Q (L) and adjoin <e> to L to obtain a lattice K with (e) j. L. Then gL = gK and mi. tr•K so that
K
(A (e, . .
± (A (b, 0) >
or
K
(A (e, . . ) > _L (A (b, 4 e
Therefore
(e)
j
L
<e> J (. . . >1 (A (b, 0))
or
<e> 1 L
<e>
(. . . >
>
By Corollary 93:14a we can cancel <e>, so L has the desired form. A computation with Hasse symbols shows that L cannot have both the given forms. (v) If dimL 5 we claim that
L
(A (0 , 0) > 1 . • • , and so gL = Q (L). We can assume that ordp a 4- ordp b is odd. First we take a quaternary unimodular sublattice of L which has norm a o. If this quaternary lattice has weight b p 2 r + 1 (r 0) it will have the form (A (0, 0)) I . . . by (ii) and we will be through. Otherwise it will have weight b p2r (r 0) in which case we obtain L (A (b 0.1 2r 2 co > K for some a in o by (iii). Here nK is still ao. If we repeat the above procedure on a suitable ternary or quaternary sublattice of K we either obtain IC (A (0, 0) > ± • - • , or else L (A (b n2r , 2x)) J (A (1)7E2' , 2 /3)) I • • for some I 0 and some f3 in o. In the first event we are through. Otherwise we can assume that r t, say. Then L (A (2 b ;Or, 2 a) ) _L - • • L". (A (0 , 0) > I • • • . (vi) As a special example let us consider the case dim L 4 with d L = 1. Then FL is isotropic if ordp a + ordp b is even, and we actually have
L
(A
(a, 0)>I
262
Part Four. Arithmetic Theory of Quadratic Forms over Rings
by (ii) and Example 93:9. If ordp a + ordp b is odd, then L (A (a, O)) J
(A (a, 4a-le)) j_
(b, 4 b-i e) .
The first of these is clearly isotropic, and the second is not by Example 1, then 63 :15 (i). So we have the following result: if dimL = 4 and dL L
(A (a, 4a-1 2)) ±
with A = 0 when FL is isotropic (in particular when ordt, a ± ordr b is even) and with A = e when FL is not isotropic. 93: 19. Let L be a modular lattice in a quadratic space over a dyadic 3. Consider any 2 E g L. Then local field with sL S p and dim L (A (a, 13)) 1 L (21(a + 2, ()) K with gL c gK. Proof. Put sL = o with E F. Then by Examples 93:13 and 93:18 we have (v) (A (a, fl)) 1 aA(0,0)>IL L-4 (A (a, p)) 1 (EA ( -1. 2, 0) )1. L .)) (A (a + 2, 13))
We have gL S
(A (a + 2, )> ± ( A (0, 0)) K gK = g K and, by Theorem 93:14, (A (a, ,8)) L (A (a + 2, 13)) K
q. e. d. 93 : 2 O. Example. Let L be a modular lattice on the quadratic space V over the dyadic local field F, and suppose that dim V 3. We claim that 0 (01 -
2 u.
By scaling V we can assume that L is unimodular. Let b denote any weight generator of L. By examining the different cases in Example 93:18 we see that there is always a maximal vector y in L with Q(y) = b. This vector satisfies 2 B (y, L)
(Y)
b
°
hence the symmetry T., is a unit of L by § 91B. So there is always a symmetry in 0 (L) with spinor norm bF2. But eb is also a weight generator of L, for any e in u. Hence e E 0 (0+ (L)). Hence 0(0+ (L)) D u. § 93E. The fundamental invariants Consider the non-zero regular lattice L in the quadratic space V over a dyadic local field. Let L have the Jordan splitting L=
• • • ± Lt
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
263
Put s i = sL i
for 1
i t. Thus s L s. And L i is s i-modular with 1
Note that L 6i
) 2) • • • D st.
s i by § 91C. We define g 1 =gL, w i wPi
Thus g 1 -1)- g 2
•'•
gt
1:D2 ----) • • • :2 rvt
since L 5i D L 6-I when si fix it. Thus we have
D
si. We take a norm generator ai for Pi and we
ao
• • • _D
at ])
and also = a10 2 +
Other relations among ai, si, wi can be deduced from § 93A. The invariants t, dim L i , s i, wi, a i t) will be called the fundamental invariants of the lattice L. The number t is, of course, the number of modular components of any Jordan splitting of L. We shall call the dim L i the fundamental dimensions, the s i the fundamental scales, etc., of the lattice L. The norm group of L is equal to the first fundamental norm group gi of L. All the fundamental invariants other than the ai are unique for a given L. By Example a; will be fundamental norm generators for L if 93:6, scalars a;, and only if ai mod tv i for 1 i t
(1
i
.
Now consider another lattice L' in a quadratic space V' over the same field. Let L' = L I • • •'1_. L, denote a Jordan splitting of L', and let t', dimg, s;, rv'i , a;
be a set of fundamental invariants of L'. Let g i' be the fundamental norm groups of L'. We say that L and L' are of the same fundamental type if t = t', dimL i = dim4 6 i = gi
264
Part Four. Arithmetic Theory of Quadratic Forms over Rings
t. This is equivalent to
for 1 _‹ I
t = t', dimL i =dimL,s i =
wi
,
modw
ai
for 1 I z t. It is clear that isometric lattices are of the same fundamental type and that an isometry preserves the fundamental invariants. Suppose L has the same fundamental type as L'. Then the fundamental norm generators satisfy a1 .-1-2 a modwl and so {a1,. . ., a} can be regarded as a set of fundamental norm generators of L'. When this is done we say that we have chosen the same set of fundamental invariants for L and L', or simply that L and L' have the same set of fundamental invariants. The lattice L i is the first component in some Jordan splitting of L 6i. Hence nLi = s i if and only if ni.61= s i, i. e. if and only if a i o = 6i . So lattices of the same fundamental type are also of the same Jordan type. Let us introduce some additional notation. We put si = ordp s i, u i = ordp ai for 1 j t. These quantities clearly depend just on the fundamental type. We define fractional ideals fi for 1
with a E gi , E gi+1 when ui ui+1 is odd, and we put
cui+ tii+0+ 8i ±2p 2 5 ffi = b (cc fi) with a C gi 13 C g11 when u i ui±i is even. These ideals fi also depend ,
just on the fundamental type. It is clear that in all cases fi
The modular lattice L i is the first component in some Jordan splitting = Li ± • • •
of L 6i. Hence we have
g L i gL6'= g i . We shall call the given Jordan splitting L = L1 I • • • I L i saturated if
gLi gi for
1
t
If L = L1 ± • • ± L i is saturated, then ai is a norm generator of L i and
wi —wL i , since L i and L 6i have the same norm group and the same scale.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
265
93:21. Suppose L has a Jordan splitting L = L 1 1 • • • I L t in which dim L1 3 /or 1 i t. Then L has a saturated Jordan splitting. Proof. 1) First let us prove the result in the case where all dim Li 7. Take a set of fundamental norm generators a t and a set of fundamental weight generators b1, . . bt. So
for 1
g i = aio2 bi o
Consider a single index i and put 93:18 L i has a splitting L.
o
a A (0, 0)) ±
with in F. Then by Example
aA (0, O)> ± • - •
and this induces a splitting L
<EA (0, O)> ±
Now
aA (0, O)> j. J .
hence ai, bi are in gr°. If we now apply Example 93:13 twice we find L
aA(
-3. at , 0)) ± a A (t bi, 0)) I
I•
Hence we have the following: if we start with a Jordan splitting L = L 1 ± • • • ± Lt there is another Jordan splitting L = K1 1 • ± Kt in which ai, bi E gKi and K 5 n-2 L i for j +1. It then follows from the properties of the weight (§ 93A) that bi o ( gK i. Hence gi = ai 02 b io gK i R gi ,
and so gi = g Ki . Do all this for î = 1, then for i = 2, etc. Ultimately we obtain a saturated Jordan splitting for L. 2) Now assume that dim L 1 3 for 1 i t. Make a 4-dimensional s -modular hyperbolic adjunction of a lattice H1 to L. Then repeat with 62, etc. We obtain a lattice L with the same fundamental norm groups as L and a Jordan splitting L =(H1 1_ Li) ± • • • 1 (Ht Lt)
with dim (Hi ± L i) splitting
7 for 1 < i t. But L has a saturated Jordan j_ • • •
Kt
by step 1). Now K1 H. ± ji by Example 93:18 since dimHi =4 and dim K1 7. Hence we have a splitting L (H1
JD ± • • I_ (He ± Jt)
with g ji == gK1 g i. By Theorem 93:14 L-=' A
Hence L has a saturated splitting.
•
-LIt •
q. e. d.
266
Part Four. Arithmetic Theory of Quadratic Forms over Rings
9 3 :22. Example. Suppose the Jordan splitting L = L1 ± • • • I L t has dim L1 3 for 1 S t. Then it follows from the definition of a saturated splitting and from Proposition 93:21 that the given Jordan splitting •• • ±L't is saturated if and only if it has the following property: if L. is any Jordan splitting of L, then gL t g./.4 for 1 t. So in this case a saturated splitting is one that maximizes all the groups g L i. 9 3 :23. Example. What happens to the fundamental invariants under scaling? If L has the Jordan splitting L • • • ± L t , then LŒ has the Jordan splitting Lee = L J • • L. Hence the new fundamental ...
-
scales are ael. )•••Dast .
We have g (Lec ) e4 = g (L)Œ = CC gLel = a gi , hence the new fundamental norm groups are otgi-•••2 0cgt-
Similarly with weights. So we obtain a new set of fundamental invariants t, dimLi, cc si, cc w., cc cti for La. It follows from the definition that the ideals fi'.. , fi _l are the same for La as for L. A Jordan splitting is saturated for L if and only if it is saturated for L. 93:24. Example. What happens to the fundamental invariants as we pass from L to L#? Consider a Jordan splitting L =-- L1 1 • • 1 Lt. Then by § 82F and Proposition 82:14 we get a Jordan splitting L4L--
••-
Lr=
x-sgL t i • • • ± x-8.4
in which the fundamental scales of L# are The fundamental norm groups and weights of L and L# are not altered by making si-modulas hyperbolic adjunctions to L. Let us therefore assume that dim L i 3 for 1 I t. Suppose the given Jordan splitting of L is saturated; then it follows easily from Example 93:22 and the fact that L## = L that the Jordan splitting 7r-8, L t ± • • • I
is saturated. Hence by Example 93:7 we have —
i+1
or
-2 9 _, wt-i+i 2
alt=
where the it symbol refers to the invariants of LA. Hence for 1 i we have f *- f 1— • .
t-1
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
267
It follows from Example 93:24 that the fundamental norm groups and weights satisfy 93:25. Example.
n -28: gt
•
a -28, 111.
71 --2 8tm t
Hence
6r 2 at
61 2
ideals fi can be expressed in terms of the fundamental invariants. Consider the lattice L with fundamental 93:26. Example. The
invariants
t, dim L 1 ,6i ,w 1 , ai .
Consider any i with 1
i < t— 1. Then it is clear by inspection that = ai ai+l o
when u • u i 41 is odd. And a straightforward computation using the fact that ai ci-2 ç a11 si-±21 will show that (Ug+ Ui+i) 8i
5F fi (aiai+i) + + ai+1w2: + 2 p" when ui ui +1 is even. 93:27. Example. Let us continue with the preceding example and compute sff i in certain special cases that will be needed later. We know that u, 1 ui since g 1+1 Ç g i. We claim that { ai±troi 6i2 fi .
ai ai +1 o
a1 w 11
if if if
U i +1
U1+1
= u. Ui
± 1
Ui
+ 2 si+1 = si +
1.
The first two computations are direct. Let us do the third. We know from Example 93:25 that n-28€-Fxg 1+1 D n-28i g i , so in the present situation we have n2 gi gi +1 cg1, hence n2 a1 may be taken as an (i V h fundamental norm generator, in other words we may assume that ci1+1 = 7r2 a1 . We also have n2 W 1
w i +1
Now substitute these values in the second formula of Example 93:26. We find that sf fi = ai w i 44 as required. § 93F. Determination of the class 93:28. Theoreml. Let L and K be lattices on a regular quadratic space over a dyadic local field F, suppose that L and K have the same fundamental I A different classification involving the so-called "Gauss sums" can be found in O. T. O'MEARA, Am. J. Math. (1957), pp. 687-709.
268
Part Four. Arithmetic Theory of Quadratic Forms over Rings
invariants, and let
Lq) ( • • • C L( i) and K( .1) C- • • (K be Jordan chains for L and K. Then clsL = clsK if and only if the following conditions hold for 1 I t 1:
(i) dL( i)IdK( i) 1 modfi (ii) FL(i) FK(i) (iii) FL(1) F K( i) j..
when when
fic 4 ai +1 tv-ai
fi C 4ai wi71 Proof. Let V denote the quadratic space in question. Let L=
-•-
Li , K
1. • • • 1. K t
denote Jordan splittings which determine the given Jordan chains. The fuhdamental invariants t, dim Li, ai refer to both L and K since it is given that L and K have the same fundamental invariants. Recall the auxiliarly notation si = ordp 6i and ui = ordp ai. Let b, .. bt denote a set of fundamental weight generators. Proof of the necessity. Here we are given that clsL clsK. Hence there is an isometry a E O (V) such that K = a L. Now a preserves the fundamental invariants, and if (i), (ii), (iii) are established for K and a L they can be carried back to K and L by a-1, hence we can assume that K = L. The above Jordan chains and splittings now refer to the same lattice L. 1) The first step is to prove (i) and (iii) for the following special case: — 1, FL 4-dimensional hyperbolic, LI unimodular, gLi gK, = If u1 = u2 we have h = a w1 by Example 93:27; but dL i. = 1 and dK, c-2. 1 moda , eh by Example 93:10; hence dL/friKi 1 modh and we have (i); condition (iii) is vacuous since w1 D 261 = 2e implies that ,
,
,
4 ai
Hence the special case is proved when u1 = u2. We therefore assume that u1 < u2. We now choose a base (exi ex2) j. (0x3 + in which <2,1(a 0)> 1, where a; and ai' are norm generators of PI; the existence of such a base follows easily from the fact that FL" is hyperbolic and that LI has norm aio. Now each of the vectors xi (1 i s-7. 4) is in L = K, j . Kr2) , hence we can find vectors yEK,, E Kr2)
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
for 1
i
269
4 such that xi = Y
Then Yi E Lel and so f B (xi, xi) = B(31 , 31,) mod 5 2 Q (xi) Q (y e) mod g2
for 1 S I
4. In particular we have
4 and 1 j
B (xi , xi) = B (yi , 5/I ) modp so dn(y1 , • y4) is a unit, hence oh + • • • + oh is a 4-dimensional sublattice of K1 with volume equal to o, hence = 0y1 + • • • -}-0y4.
Now solve the pair of equations
B 3/ 2) + B (y2, y 2) + for E3, 773. We have {
Q(y 1 ) + 773 B (yi, y2) = 0 B Y) + 273 Q (Y =
(311) Q(y2) — B (y1, Y2) 2 EU,
and so the solutions 3, 713 are in s2. Then 34 = Y3 + Eah.
2702
is a vector of K1 which is orthogonal to both Yi and y2. Similarly obtain 3/4 = Y4 + .1)11
orthogonal to y 1 and y2 with = ( 11
174Y2
and 774 in 62. Then 0 Y2) -I- (0 )4 + *A) •
We find that we still have {
for 3
B (xi, xi) B yi) mod6 2 Q (xi) Q (y;) mod g2
4 and 3 j of f we find that
4. By a direct calculation using the definition
dB (yi., Y2)1 modh ,
dB
—1 modh ,
hence &K.,. 1 mod f1. Hence dL ifeiK i. , 1 mod h. This proves (i) in the special case under discussion at the moment. We must prove (Hi). We assume that h C 4 al toTl. We have Q(y i )o = = n.K i since Q Q(x 1) modg2, and since g 2 Ç al p by the assumption that u1 < u2. Similarly with A. So we have just found a splitting = J j_ J'
270
Part Four. Arithmetic Theory of Quadratic Forms over Rings
in which J and J' have norm ale and
dJ
—1 mod
, d
—1 mod fi .
Then by Example 93:17,
.T=. 21 (aicc2 + fl, 71) with a E v, flEw1, vE aT'f1 c 4wr'p. Hence
A> I r 1_
Now
J'J
(2% 0C2 + fl) y E (w1) (41V/71) hence is isotropic by the Local Square Theorem and Proposition 42:9, hence
<%) =. J r L
.)>
<•••>.
Hence F.Kl
I. < • • > .
Hence FL/ ±
. Li-1 ± • • 1. Li4 allows us to assume that each space FL i (but not necessarily each TWO is hyperbolic. Suppose we wish to verify (i) and (ii) for a specific value of i. By suitably scaling V we can assume that Li+, and K i +1 are unimodular. Now Lti 1) is a hyperbolic space. Hence by Proposition 93:3 there is a
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
271
unimodular lattice J on FLit + » with Lri+i) _Ç J FLA+ » and g J =g Put L' = L + J. By considering the Jordan splitting L' = L1 ± • • ± L j. J it is easily verified that the first i -I- 1 fundamental invariants other than the fundamental dimensions are the same for L and L'. Obviously the first i fundamental dimensions also agree. So the above splitting for L' is saturated. If 1 A j the lattice KA is 62-modular with $2 D o, and B (KA, L') cC B (KA, L) B (KA , J) ç B (KA, L) B (L', sa Hence KA splits L' by Proposition 82:15. Hence by Proposition 82:7
there is a splitting L'
± • • • 1.
±I.
It is easily seen that this splitting is saturated. If we can prove (i) and (ii) for the lattice L' at i, then we shall have these conditions for L at i. In effect this means that we can make the further assumption that i = t —1 with L t unimodular. By Example 93:18 we have a splitting L e = L j . Lr with dim/4 = 4 and Lr
.
ft
_,
dL eldK t L, .-- 1 modf? .
by Example 93:24, and dL( t _i) • dL t dK( t _i) dK t, 1 modft _i .
We have therefore completely proved (i). Now (ii). We suppose that h-1C 4 at WTI. By Example 93:24 this condition translates into C 4 ar(w4t) -1 . So by step 1), FL FK j <—ar>. Hence FL<•••> KTI I
272
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Hence FL(t _1) F 4_0 ±
F.Kil+1)
4)
J
.
Adjoining V to each side gives FL(i)
j
FL;+1) FK'4 ÷/) J < • • >
FK(i) j FKri+i) F ..
Hence FL(i) FK( i) J
A similar argument involving the adjunction of the lattice L -1 Lr 1 ± • • • I LT1
allows us to assume that each space FLi (but not necessarily each FKi) is hyperbolic. Now by Example 93:18 we know that each Li (or Ki) is of the form L1 A (O, --A (O, with dimg = 4. By making suitable cancellations to the L i and K i we see that we can assume that dimLi dimKi 4 for 1 I t. Finally we scale V to make L1 and K1 unimodular.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
273
So we have reduced things to the following: we are given lattices L and K with the same fundamental invariants in saturated Jordan splittings
L=
K=
••• I K t
which satisfy conditions (i)—(iii), the first components L1 and K1 are unimodular, and all spaces FL are 4-dimensional hyperbolic, and we must prove, using the inductive hypothesis, that L and K are isometric. Note that Example 93: 18 (vi) asserts that L1 has the form L1 ,-_1-4, (A (al, C)> I . 3) The splitting L = L1 I•••IL t will be left fixed throughout. The computational part of the proof consists in successively changing the Jordan splitting K1. I K2 until a new splitting is obtained in which K1 is isometric to L1 . It is understood that each splitting employed in this procedure induces a saturated splitting on K. We shall use (without further referênce) the rules of Example 93:13 and Proposition 93:19 to effect these changes to K1 I K2 . In connection with this matter we must mention that conditions (i)—(iii) will continue to hold if the given splitting K1 I • • • I K t is replaced by any other Jordan splitting K I • I K. For dL( i)/dK( i) ,- 1 mod fi by hypothesis, and dK (i)/c/Kr(i) 1 mod fi by the necessity of this theorem, hence dL (i)/c/K i) 1 modfi and so (i) holds in the new splitting. Let us do (ii). Assume we have ft c 4 a t+1 tv,71 11 . Then FL (t) FK(i) I
18
274
Part Four. Arithmetic Theory of Quadratic Forms over Rings
saturated. Hence Lr2) and KA) satisfy conditions (i)—(iii), hence they are isometric by the inductive hypothesis. Of course this implies the isometry of L and K. 4) Now we start the computational part of the proof. First suppose that u2 = u1 . So a2 o = ai o. In this event we have fi a2tv/ by Example 93:27. We know that dLi. = 1. Let us arrange dif/ 1. By condition (i) we have c/K/ 1 modfl, hence d.K 1 = 1 + ad. for some A E w1 . Then
.
K2' 1.
± K 1 ± K _L K _L K.
Hence K1 ± K2 has a saturated splitting K1 _L .1q in which c/K1 = 1. In other words we can assume that difi = 1. By Example 93:18 (vi) there is a splitting
I. in which A = 0 when FK/ is isotropic and A. = e when FK/ is not isotropic. If w2 c 4tr•i-I, then fi = a2w1 (4 a 2 w' and condition (ii) holds. Hence in this case we have FL1 FK/, so FK/ is isotropic, therefore = O. Otherwise tv2 4w11 = 4 WI). Then
_L
.
So we can take A. = 0 in all cases. Then K1 LI. Hence K L by the observation at the end of step 3). 5) Next suppose aio = 2o. By step 4) we can assume that a2o S 2p. Hence f/ c. 4p. So d.KI 1 mod4 p, hence dKI = 1 by the Local Square Theorem. Hence by Example 93:18 K i has the form
. But Li has the same form. Therefore L1 K1 and L K by the observation at the end of step 3). 6) Now let u2 = U l + 1. We can assume that ao D 2o by step 5). Then a2 E w1 since a2 E g2 S g/ with ordp al ordp a2 odd, hence al p = a2o is contained in g hence ai o is the maximal ideal in gi , hence g/ = And bl o tv1 a2o since ai o D 2o. Now f/ al a2 c' by Example 93:27, so we can proceed as in step 4) to get a new splitting K1 ± K2 in which //K t - 1. By Example 93:18 there is therefore a splitting ,
f-e. _L
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
275
in which A = 0 if FK„ is isotropic and A = e if FK„ is not isotropic. If i 1 0 we have fi = a1 a2 o (4a2 toy' and condition (ii) holds. This tv2 C 4a— irnplies that FL, FK„ and so A = O. If tvi D4ar1P we can proceed as in step 4) to get a new splitting K11 K2 with A = O. So in any case we can arrange A — O. Then Li K1 . Hence L K. So the case u2 = u, 1 is proved. 7) Next we settle all cases in which .S2 = 1, j. e. in which the second component is p-modular. By steps 4), 5), 6) we can assume that a 2 o Ça1 p 2 and al o 20. Then in fact a 2 0 = a1 p 2 since a2 g 1 c g2 c g„. We have a1 tv2 by Example 93:27, so we can arrange the splitting K1 I K2 to be such that dK„ = 1 as in step 4). Continuing as in step 4) we obtain a new splitting ± K2 (A (a1 , C)> ± (A (b1, OD 1 q and then K L. (There are two variations in this part of the proof: first, one uses (iii) instead of (ii), secondly one needs the inclusion a2 o a 1 p2 4aTio .) 8) Finally we consider the following situation: s2 > 1, a2 oÇ a1 p 2 , al p D 2o. Perform the adjunction of the same lattice J to L and K, where <7. cA (0, 0) > . Put L' = L I J and K' K I J. Suppose we can show that L' and K' satisfy the conditions of the theorem. Then s; si st s, and 4 1, so L' and K' will be isometric by steps 2) and 7). Hence L and K will be isometric by Theorem 93:14. The rest of the proof therefore concerns itself with showing that L' and K' satisfy the conditions of the theorem. We have Jordan splittings —
L'— L 1 1 L i t,
L2 1 • • • 1
K1 1 K 1 1 12 1 K 2 1 • • • 1 K t
Lt ,
in which L 1 1 1 J = K 1 1 1 . It is easily seen that L' and K' have the same fundamental invariants. And it is clear what happens to the invariants t, dimLi , 5 i as we pass from L to L'. What about the new fundamental norm groups, weights, and norm generators ? We find that we can take
(i = 1, 2, . .
t) ,
hence from the definition
f;
fi
(i
-
2,
•
• , t) •
We have (L')P =
j
L2 1 ••• ILt,
hence n (L')P — :7 2 ai ø and we can take ali 12 = a2 a1 as a fundamental norm generator. By Example 93:25, :72 gi C g 11 ç g1
gr2 w1 18*
276
Part Four. Arithmetic Theory of Quadratic Forms over Rings
and g2
g 1 1 /2 P W 2 -C W1 11 2 •
Hence from the definition
-
fi
Ti fi ,/ It is now obvious that L' and K' satisfy (i) for all i. Also that they satisfy (ii) and (iii) for i — 2, . . . , t. Also that they satisfy (ii) for i = 1'/ 2 . We see that (ii) is vacuous for i 1 since then 2
wit h fi = a1 wIt h with 7 4 al p 2 = 4 ait h . We leave the verification of (iii) for 1 — 1 and i= 1'12 to the reader. q. e. d. § 93G. The 2-adic case The ideal 21) in a dyadic local field F is always contained in the maximal ideal p since the residue class field has characteristic 2. Let us suppose for the rest of this subparagraph that we actually have 20 = p, or equivalently that 2 is a prime element in F. We shall call such a field a 2-adic local field in order to signify that 2 is one of its prime elements. (The field of 2-adic numbers Q2 is an example of a 2-adic local field.) Otherwise let the situation be the one already under discussion in this paragraph. In particular let L be a lattice in a regular quadratic space V over F with the Jordan splitting L
j_ • • •
Lt
If a is any fractional ideal in F, then there are no fractional ideals properly situated between 2a and a. Hence n L is either 5L or 2(L). So by Theorem 91:9 the ideal tt L i is independent of the Jordan splitting L 1- • "L t. We define
t ni n L i for 1 (This definition applies only in the 2-adic case since otherwise the ni not uniquely determined by the Jordan splitting.) We call the are quantities t, dim L1 , s i , ni the Jordan invariants of L. It is clear that two lattices have the same Jordan invariants if and only if they are of the same Jordan type. We put
ui ordp ni We have
nL1+1 ç sL i ç 2(sL 1) ç nL i and so th 7 n2
•••
.
What is the connection between the Jordan invariants and the fundamental invariants of L? The maximal ideal in L in the norm group
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
277
gL must be all of nL since 2 (BL) ( na SnL with ordpmL + ordp nL even. Hence gL =- nL and toL = 2(sL) . Then gLei = nLei = nL i and so gi — ni , tvi = 2si for l
it.
In particular this shows that the quantity ui = ordp ni is equal to the quantity ui = ordp ai defined in § 93E. We can take any element of order ui as a fundamental norm generator, in particular we can take ai . Th. The lattice L now has the following set of fundamental invariants: t, dimL i, si , 2,
This shows that any two lattices with the same Jordan invariants (over a 2-adic local field) have the same fundamental invariants. 93:29. Theorem.' Let L and K be lattices on a regular quadratic space over a 2-adic local field F, suppose that L and K have the same Jordan invariants, and let L6.) C • • • C L( t) and Kb) C • - • (K be Jordan chains for L and K. Then clsL = clsK if and only if the following t — 1: conditions hold for 1 i
(i)
dL( i)IdK( i) e. 1 modni ni _ols., (ii) FL(i) -4— FIC(i) i (2ui> when n1+1 (4n. Proof. 1) We obtain the necessity by suitably interpreting Theorem 93:28. Take a i = 2" for 1 ._ i_.< t. Then it follows by direct calculation from Example 93:26 that s1 fi D ni ni+„ and that fi ç 80. 61 fi D ni ni+1 Hence (i) is a consequence of condition (i) of Theorem 93:28 and the Local Square Theorem. We find that n i +1 S 4ni implies that fi c 4ni triTi, therefore (ii) is a consequence of condition (iii) of Theorem 93:28. 2) The sufficiency is proved in essentially the same way as the sufficiency of Theorem 93:28. Let V denote the quadratic space in question. Let L = LI _L • • - _L L t , K . K i _L - - • _L K t denote Jordan splittings which determine the given Jordan chains. We shall prove the sufficiency by induction on the quantity t. If t = 1 we See O. T. O'MEARA, Am. J. Math. (1958), pp. 843-878, for the solution of the representation problem over 2-adic fields. A general solution over arbitrary dyadic fields is not known at present, but much progress has been made on this difficult problem by C. RIEHM, On the integral representations of quadratic forms over local fields (Princeton University thesis, 1961). 1
278
Part Four. Arithmetic Theory of Quadratic Forms over Rings
have gL1 n1 gKi and so clsL = clsK by Theorem 93:16. We therefore assume that t> 1 and proceed with the induction. By making suitable hyperbolic adjunctions we can assume that i t, in virtue of Theorem 93:14. By adjoining dimLi 3 for 1 the lattice = -LT1. we see that we can assume that each space FL, is hyperbolic, in virtue of Corollary 93:14a. By making suitable hyperbolic cancellations we i t. By scaling V we can can assume that dimLi dimKi = 4 for 1 assume that L1 and K1 are unimod-ular. It is enough to find a new Jordan splitting K .1q If3 I •• I K t L1. For then we can assume that K1 = L1 by Witt's in which K theorem. So the lattices L2 ± L 3 ± • • j_ L, and Iq K3 ' • • I K t are on the same quadratic space, and they clearly satisfy the conditions of the theorem. The inductive hypothesis then asserts that these lattices are isometric. Hence L K. First consider the case û2 = 2o. Then Li . But
2(70> ± (A (2, 2 i3)> with a, 9 E 0, hence by Proposition 93:19 K1 ± K2 (.11 (2u1, j_ with 2. equal to 0 or e and the same with #. But n2 = 4o Ç 4n1, hence ,
FL,
,
<s)
± <1>
where. e is the unit (1-42.) (1 - 4 It). A computation of Hasse symbols, say, will show that # = 0. Then
± K2
j_ j_ .1q
since g K2 = n2 4o. But
j_ .
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
279
Finally let n2 S 80. Then dL I/dK i ._—_.t 1 mod8o, hence difi = dLi = 1. But n2 S 8o S 4n1, so that FL, ± (1>FK1 ± <1>, hence F Ki is isotropic. Hence _L q. e. d. § 94. Effective determination of the invariants Let us show how to compute the invariants of the local integral theory. We suppose that L is a regular lattice given in the base L = o - - • + ox, and that the symmetric bilinear form B is determined by specifying the values B (xi , xj) = aij . (In practice it is just the symmetric matrix (aii) that is given.) To find the scale we simply take the largest of the ideals B (xi, xj ) o, i. e. 5L = f (a • • o) •
To find the norm we take the largest of the ideals Q (xi) o and 2 (L), i. e. (a i o) 2 (5 L) . nL These rules are justified by Proposition 82:8. Next we show how to obtain a Jordan splitting from the given base. First the non-dyadic case. If there is no xi with Q (xi) o = sL, then there will be two distinct basis vectors, x1 and x2 say, with B (xi, x2) o applying the Principle of Domination to the equation Q (x1 + x2) =Q (xi) Q (x2) + 2B (xi, x2) shows that Q (xi + x2) o = 5 L. (Note that this will not work in -the dyadic case.) Hence a minor adjustment to the given base allows us to assume that Q (x1) o = BL. All the coefficients B (xi, xi)fQ (xi) are now in o, hence the vectors xi, y2, . . y. with Yi = xi
B (xi, x1)
form a new base for L and in fact we have a splitting L= ox1 ± (oy2 + • • • + oyn) . Repeat all this on o y2 + - • + oyn, etc. Ultimately we arrive at an orthogonal base for L, and a Jordan splitting is formed by suitably grouping these basis vectors. The dyadic procedure is somewhat longer. If there is a vector, say xi, with Q(x1) o = sL, use the non-dyadic method to split off o xi. Otherwise Q (xi) oc sL for 1 i n. Then B(x1 x2) o=sL, Q (xi) oCeL ,
280
Part Four. Arithmetic Theory of Quadratic Forms over Rings
say. This implies that the equations
B (x1, xi) + Ei Q (x1) + B (x1, x2) = 0 B (x2, xi) + Et B (xi, x2) + 1h Q (x2) = 0 have integral solutions X1, x2, y 3 , • • • , yn with
Ei ni for ,
3 I
n. Hence the vectors
yi = xi + ni x2 form a base for L, and in fact there is a splitting
L (o + o x 2) ± (0y3 + • • • + oy.„) with O X1 + O x2 modular. Repeat on oy3 + • • • + oyn, etc. Ultimately we obtain a Jordan splitting for L. The Jordan splitting is all that is needed in the non-dyadic theory. The Jordan invariants are also needed in the 2-adic theory, but these can be read off from the Jordan splitting. So Theorems 92:2 and 93:29 can be applied in practice. The general dyadic.theory requires a knowledge of the fundamental invariants ai and wi (the invariants fi are obtained from the fundamental invariants using Example 93:26). First let us find a norm generator a and the weight t0 of L. The norm generator is easy: if nL = 2 (s L) take any a with ao ---- 2 (s L) , otherwise take a = Q (xi) for any value of I which makes Q (xi) o largest. The weight is then given by the formula w=f ab(Q (xj)Ia) + 2 (s L) .
We leave the verification of this formula to the reader. In order to find , w t we apply this the fundamental invariants a1,.. ai and w1, Pg. These lattices ' can be procedure to each of the lattices Lei, expressed in terms of the Jordan splitting L L 1 I • • • ± L t by the formula (grgi--814) -••± Li _i) ± L i ± • • .
§ 95. Special subgroups of 0.(V) We conclude this chapter by giving the structure of the groups
n Z,,, On'
, 0„410„'
of a regular n-ary quadratic space V over an arbitrary local field. Recall that we first raised this question over a general field in § 56, and that we have already described these groups over complete archimedean fields and over finite fields in §§ 61 and 62. 95:1. Let V be a regular n-ary quadratic spq.ce over a local field F. Then 0;, .Q.„ with the following exception: V is a quaternary anisotropie space over a non-dyadic field. In the exceptional case we have (0'4 : D.4) - -- 2.
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
281
Proof. 1) If V is isotropic, in particular if n 5, then On = 12„ by Corollary 55:6a. If 1 3 we have On,' = Q Proposition 55:5. So the quaternary anisotropic case remains, and we must prove that 04 = Q4 if F is dyadic and that (OI: Q4) — 2 if F is non-dyadic. Note that d V = 1 by Proposition 63:17. If cc is a non-zero scalar in Q (V) we let .r. „› stand for a symmetry with respect to a vector x with Q (x) = oc. This is not a precise notation since -Q,c> is not uniquely determined by oc. However the coset -r <,(> Q4 is well-defined in virtue of Proposition 55:3. In fact we can write TC1x) - • • r
( ai
E 17)
without fear of ambiguity, and this coset is even independent of the order in which the cci appear. 2) We are ready to do the dyadic case. We must prove that a typical element a of 04 is actually in Q4. Express a as a product of four symmetries, say T(Œ1)..
• T OO
(az
C
Here oc1oc2oc3oc4 is in F2 since a is in 0,'4 .Clearly oc1 oc2 , oc3, cc4 fall in at . . most four distinct cosets of I; modulo F2 . On the other hand (F:F2 ) 8 by Proposition 63:9. Hence there is a e in F such that eocl , eoc2 , eoc3 , ecc4 are non-squares in F. Now V represents —e since V is universal, hence there is a ternary subspace U of V such that V (—e) I U. Here we have dU = —e since d V = 1, hence the discriminant of the space U is a non-square, hence this space is isotropic, hence U represents ocl, oc2 , oc3 , oc4. We can therefore write ,
cr•Q4 —
T(cc i> • • • T(cc4 > Q4
in such a way that each of the symmetries appearing in this equation is with respect to a line in U. But 0(U) Q3 (U). Hence a.(24 = Q4 . SO a E .f24, as required. 3) From now on we can assume that F is non-dyadic. Let A denote a fixed non-square unit in F. Take a typical a in 04 and express it in the form r
> r<> r
Otherwise we can write Cr 124
TO)
r
282
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Hence 0 ri consists of at most two cosets of Q4, namely Q4 and To> T(A> T(,> Tot A>Q4.
Hence (0'4 : Q4) = 2. 4) Finally we must prove that (0,1: Q4) = 2 for the non-dyadic case now under discussion. In order to do this it will be enough to produce an element of 0 41 which is not in Q4. Since V is anisotropic there is exactly one o-maximal lattice on V by Theorem 91: 1. Let L denote this lattice. By Proposition 63:17 we can take a lattice L1 J L 2 on V with
<1> I (--/1>
and
L2
(n> 1 <----r/1).
Then L1 I L 2 Ç L. But L cannot be unimodular, for if it were the space V would be isotropic by Example 63:14. Hence by a volume argument
L
L i j_ L 2 .
We shall need the lattice LP defined in § 821. Note that LP
p
I L 2 pL
Since D' is an additive subgroup of L we can form the factor group L/LP. Let l denote this factor group and let bar denote the natural group homomorphism
L --> = LILP Also let bar denote the natural ring homomorphism of o onto the residue class field F of F. We put a scalar multiplication on V by defining
ax
(c
, x E L) .
This is well-defined since p L Ç I)', and it makes V into a vector space over F. It is easily seen that any base for the lattice L1 becomes a base for the vector space F , hence dim V - 2. Put a symmetric bilinear form on V by defining B(, )7) B (x , y) V x, y E L This is well-defined and it makes V into a regular binary space with discriminant —,71. Each a in 04 is a unit of L by Proposition 91:5; and aL = D' by definition of D'; so a induces a mapping 5: V -->- V by means of the equation ax V x EL;
r;
it is easily seen that 5 is an isometry of V onto and that 19-1cr2 =51 52 for all cri , a2 in 04. Hence we have a natural homomorphism
04 (V)
02 (F) .
By Corollary 92: lb we can find vectors x, y (x) -- 1 , 42 (y) = A ,
(14) = T
L1 and u, V E. L 2 with (v)
Chapter IX. Integral Theory of Quadratic Forms over Local Fields
283
Define = Tx Ty Tu Tv E
.
We assert that a is not in Dd. Once we have proved this we shall be through. Suppose if possible that a E Dd. Then o E 122 (7) 1%(7'). On the other hand it is easily seen from the defining equation of a symmetry that O = Ty . Hence (9 (a) = Q Q(37) = J. But 71 is not a square in P by the Local Square Theorem. So O is not in 49 (F). This is absurd. Hence (0 D4) =-- 2 as required. q. e. d. 95: la. In the exceptional case TO> r 'V W "r 124 and DA are the two distinct cosets of al modulo Dd. Here Li can be any non-square unit of F.
95:2. Let V be a regular n-ary quadratic space over a local field with n 3, and let a be an element of 1.2n of the form a = "Coo . . . Too with r < n and all ai in F. Then there is a regular ternary subspace of V which represents
Proof. If V is isotropic we take any regular ternary subspace which contains an isotropic vector of V. We may therefore assume that V is a quaternary anisotropic space and that r = 4. In the dyadic case we can use the argument used in step 2) of the proof of Proposition 95:1 to find a regular ternary subspace of V which represents x1,. cc4. So let us suppose that F is non-dyadic. If xi , . 24 fell in distinct cosets of F modulo F2 we could not have ci in 124, by Corollary 95: la. Hence we can assume that 2ia2 E P2 and o 324 E F2, say. There is clearly a ternary (in fact a binary) subspace of V that represents al and as. This space represents al, as. q. e. d. If V is any regular quadratic space over a local field other than an anisotropic quaternary space over a non-dyadic local field, then the condition for 1-4 n Zn to be {1 l v} is the same as the condition for —1 v to be in 49;„ namely dV = 1 with n even. Let us settle the exceptional case. We claim that in the exceptional case we will have Dd n Zd = ± 1 v} if and only if —1 is a non-square in F. For suppose that —1 is a non-square in F. Then —4 E ff'2. Hence
V <1> ± <1>
± ,
therefore , iv =. TO> TO> .r<x> 2.<,1 > E and so 124 n Z4 = { ± l v}. Conversely suppose that —1 v E124. If-1 were —
284
Part Four. Arithmetic Theory of Quadratic Forms over Rings
a square we would have l'o.) t(4)
T(s) T(vrA) 94 T(1) 1..(—A) T r) T(---svA) 94 — (
94 = 94
and this would contradict Corollary 95: la. So —1 is a non-square. Finally we recall from Proposition 91:6 that the group the form
has
for n 3. The same applies if n = 2 with V isotropic. If n = 2 with V anisotropic, then one can use Example 63:15 to show that 0/0,,,' is isomorphic to a subgroup of index 2 in
We have therefore fully described the groups rA Z, 0114, 0110;
over local fields. Chapter X
Integral Theory of Quadratic Forms over Global Fields We conclude this book by introducing the genus and the spinor genus of a lattice on a quadratic space over a global field, and by studying the relation between these two new objects and the class. We shall use these relations to obtain sufficient conditions under which two lattices axe in the same class. We continue with the notation of Chapter VIII, except that the field F is now a global field and S is a Dedekind set of spots which consists of almost all spots on F. We let o be the ring of integers o (S), u the group of units u (S). As usual we let p stand either for a prime spot in S or for the prime ideal which it determines in o. (There will be one exception to this notation: in § 101A we shall let F denote an arbitrary valuated field.) Ar or Q will stand for the set of all non-trivial spots on F, I 1 p will be the normalized valuation on Fp at a spot p in Q. If p is discrete we let op, up, mp stand for the ring of integers, the group of units, and the maximal ideal of Fp at p. ap will be the localization at p of the fractional ideal a of F at S. V will be a regular non-zero n-ary quadratic space over F with symmetric bilinear form B and associated quadratic form Q. We shall consider lattices L, K, ... on V, always with respect to the underlying set of spots S. We let Vp denote a fixed localization of the quadratic space V at a spot p in Q. The lattice L p will be the localization of L in Vp at any spot p. in S. The notation 0 (V), 0+ (V),. .. for the subgroups of the orthogonal group will be carried over from Chapters IV and V.
Chapter X. Integral Theory of Quadratic Forms over Global Fields
285
§ 101. Elementary properties of the orthogonal group over
arithmetic fields § 101A. The orthogonal group over valuated fields In this subparagraph F denotes an arbitrary valuated field, not necessarily the global field F under discussion in this chapter. Let 1 I or I lp be the given valuation on F, and let p be the spot which it determines. V is an n-dimensional vector space over F. A base x1,. . . , x„ is taken and fixed for V. The norms II II which we are about to define are with respect to the same base xj., . . . , x,,, unless otherwise stated. Recall our earlier notation: 1,7 (V) denotes the algebra of linear transformations of V into itself, and /12(F) denotes the algebra of n x n matrices over F. Practically no proofs will be given here. All assertions can be verified either by inspection or by simple direct calculation. First we define the norm on V. Given any vector x in V, express it in the form x — oci xi + • • • + cc„x,,, (cci E F)
and then define the norm of x by the equation
11xIl p = max lai lp . i Use II II instead of 11 Il y whenever convenient. So II 11 is a real-valued function with the following properties: (1) 11x11 > 0 if x E 1,., and 11 0 11 = 0
V a EF, x E V (2) Ilaxil = 1 0:1 11x11 (3) 11x ± Yll .- 11x11 ± 11Y11 V x, y E V. In other words 1 II is a norm in the sense of § 11 G. And we can make V into a metric (topological) space by defining the distance between the vectors x and y to be IIx — yll. As usual,
1 1 14 - 11311 .__ 1 x In the case of a non-archimedean field we have
11x ± Yll 5 max (11x11 , bill) V x, y E V with
lix ± yll = max (Ix°, 11Y11) if 11x11 * 11Y11 • In particular, in the non-archimedean case there is a neighborhood of any given point x0 + 0 throughout which
11x1 1 - 1 x011 .
Part Four. Arithmetic Theory of Quadratic Forms over Rings
286
Each of the mappings (x, y) ->- x ± y of V x V into V, x of V x into V, (a, x) ax of F x V into V, is continuous. This means, to use the language of topological groups, that V is a topological vector space over the topological field F. The map (Y1 , • • • , Yr) -›- Yi + ' • ' + of Vx•••x V into V is continuous. So is the map x ->- 11x11 of V into R. Now do the same thing with LE (V). Consider a typical ci in LE (V), write axi = E oci; xi (mu EF)
for 1 j
n and define the norm of a by the equation
lloll
= max
1 = max Ilax5 11 p .
lip whenever convenient. Then 11 11 makes L(V) into a normed vector space, i. e. we have (1) IJciJ > 0 if crEL E (V) with ci== 0, and 110II = 0 V aEF, ci E (V) (2) 11'2 4 = 1 01 11011 Use II II instead of
(3) 1 0' + Tli
1 011 + liT11
V a, 'r E LE (V).
And L (V) is provided with a metric topology in which the distance between a and r is defined to be Ila — T11 . As usual, I 11 0II — II TII I -5 2.11 • In the case of a non-archimedean field we have TII
with
max (11 4 114)
V a, T
E LE (V)
1 0' + Til = max (11011 , linii)
if 1 011 I 114 In particular, in the non-archimedean case there is a neighborhood of any given co == 0 throughout which liall coil • We again have continuity of addition, of taking negatives, and of scalar multiplication, so that LE (V) is also a topological vector space over the given topological field. All this parallels the discussion for V. But we also have multiplicative laws to consider. We find that
lic xil
n 11 011 14 1 01 114
in general if non-archimedean,
Chapter X. Integral Theory of Quadratic Forms over Global Fields
287
for all a in Lp (V) and all x in V. Similarly for a, r in L p (V) we have Til 5- I.
n 110 lirO
IiolI uni
in general if non-archimedean.
Hence the mapping an of Lp (V) X L p (V) (a, n)
into Lp (V)
is continuous. This makes LF (V) into a topological ring as well as a topological vector space (as the name suggests, a topological ring is a ring with a topology in which addition, the taking of negatives, and multiplication, are all continuous). The mappings (al,
• • • C rr) ->-
± " • • ±
and
(a1 , • • ar) ->- al • • • ar are continuous. So are the mappings (a, x) -›- ax of LAV) x V into
det a of L p (V)
V,
into F.
The continuity of the determinant map shows that the general linear group G L (V) is an open subset of L F (V) . If we restrict ourselves to GL (V) we find that the mapping
a
a-1 of GL (V)
into GL (V)
is continuous. on the algebra of n x n matrices We can introduce a norm if 111-11 (F) by defining II (aii)11 p = max lad y for a typical matrix (a id) over F. We shall write 11 a.0 1 or liciA instead of II (aii)iip. Note that the norm hail of a linear transformation a is equal to the norm of its matrix in the base x 1 , . . xn. 10 1 : 1. Example. What happens to the norm under a change of base ? Take a new base xi, . . .4 for V with
E ail xi and X5 = f bo xi Let
denote the norm with respect to the base xi, . .
general we have
11z11 nIlb5I1 11x11
and
nhlai
jhl
5 -
II< < n2 au Il °bo11
11all
Then in
288
Part Four. Arithmetic Theory of Quadratic Forms over Rings
for any x in V and any Cl in LF (V). In the non-archimedean case we have
ilzil
11a„11
III
Ilxii
and
loi'
1Ia1I Ii
11a11
To conclude we suppose that V has been made into a regular quadratic space by providing it with the symmetric bilinear form B and the quadratic form Q. Then there is a positive constant 2 such that
Allx11 113111 V x, y E
1B(x,
V,
so
(x)i
214 2 VxE V.
The mappings (x, y)
B (x, y) of
x —>— Q (x)
V X V into F,
of V
into F ,
are continuous. 1 0 1 :2. Example. The continuity of the map x —>— Q (x) shows that the set of anisotropic vectors of V is an open subset of V. Let u denote an anisotropic vector of V. Then the mapping u
of the set of anisotropic vectors of V into On (V) is continuous (here t„ denotes the symmetry of V with respect to u). To prove this one considers the defining equation rts x = x
2 B (u,
(u)
of a symmetry. First one shows, using the continuity of the maps u 2 B (u, x), u Q (u), etc., that the mapping u 7.x is continuous for each fixed x in V. One then deduces the continuity of u -L.'. at u0 from the equation — tu0 x511 !iris — risoil = max Hence u -4— T. is continuous. 10 1:3. Example. The continuity of the determinant map tells us that 0+ (V) and OE (V) are closed subsets of 0(V). Hence 0+ (V) is an open and closed subgroup of 0(V). 101:4. Example. (i) Suppose that the field F under discussion is actually a local field. Then for any ci in 0(V) we have deter = ± 1, hence det a is a unit, hence 114711 >_ 1. Now let M be the lattice M ox1 + • • • +1) xn , where x11 .. , xn is the base used in defining 11 11. Then pall = 1 if and only if Ili 1, this is equivalent to 47/1/ S M, and hence to aM = M. So the elements of 0 (M) are precisely the isometries of V with 'loll -= 1. In particular the set of isometries with 11cV = 1 is a group.
Chapter X. Integral Theory of Quadratic Forms over Global Fields
289
(ii) Consider a second lattice L on V. We claim that a L L holds for all a in 0(V) which are sufficiently close to l v. Take a base 4, for L, and let 1 -11' denote norms with respect to this new base. Then by Example 101:1 we see that all a which are sufficiently close to i v satisfy 11o.— 1 v 11' < 1. Each such a satisfies 'loll' = 1, hence a L = L. (iii) Consider a third lattice K on V, and suppose that K = AL for some A in 0(V). We claim that a L = K holds for all a in 0(V) which are sufficiently close to A. By choosing a sufficiently close to A we can make 11 2-l a vii i ii 2-1 Ail arbitrarily small. But all A-' ci which are sufficiently close to i, make A-iaL= L. Hence all a which are sufficiently close to A make aL= AL—K. § 101B. The orthogonal group over global fields We return to the situation described at the beginning of the chapter: quadratic forms over global fields. F is again a global field and V is a quadratic space over F. Since each Vp is a vector space over the valuated field Fp we can introduce norms 11 11 p on V and LFp(Vp) with respect to any given base of Vp, in particular with respect to any given base for V over F. We shall always assume that all norms under discussion are with respect to a common base for V. We let xl , , xn denote the vectors of this base. If a new base xi, .. x t, is taken for V, and if we consider the corresponding norms 4, at each p on F, then it follows from Example 101:1 and the Product Formula of § 33B that we have for almost all p.
11 11p = III]
Consider a typical linear transformation a in LE (V). By considering the effect of a on a base for V we see that there is a unique linear transformation op on Vp which induces a on V. We call ap the p-ification or localization at p of a. It is easily verified, again by considering a base for V, that we have the rules (o.
.r) r
ap
(Occf)p = xa,,
)
(a
p
ap Tp
detap = deter
for all a,1** in LF(V) and all a in F. In particular, the mapping a gives us an injective ring homomorphism LF(V)>-- LFp (Vp) . If a is an isometry, then so is 0.p. If a is a rotation, then so is 0.p. If tu is the synunetry of V with respect to the anisotropic line F u, then a geometric argument shows that (r.)4,, is the symmetry of Vp with respect to the line F4,14; we express this symbolically by the equation (Tu)p = Tu • O'Meara, Introduction to quadratic forms
19
290
Part Four. Arithmetic Theory of Quadratic Forms over Rings
In keeping with functional notation we let A 9 denote the image of a subset A of Lii, (V) in LFp (Vp) under localization at p. We have Ç On (V) p S On (Vp) , 0„+ (V) p S On+ (Vp) , 1.07,(V) p g_ 1-4(Vp) , 0, (V) _ç_. 0 (V,,). 101:5. Conventions. In some situations things become clearer if the notation is relaxed and a is used for the localization ap of a, in other situations the strict notation ap is preferable. We shall use both. We shall also use 99p (or ci,,) to denote a typical element of LFp (Vp); of course this does not necessarily mean that 999 is the localization of a linear transformation 99 of V. 101:6. Example. Let S be a Dedekind set of spots on F, let L be a lattice in the vector space V over F, and let a be an element of LF (V). We claim that ap L p = (a L) 0 Vp ES. To prove this we express L in the form L = ai yi + - • • + a r yr where the ai are fractional ideals and the yi are elements of V. Then ap L p = ap (a/p yi ± • • • + arpyr) —alp (aYi) + • • • + an, (TYr) — (ai (aYi) + ' • • + ar (aYr))p
= (aL) p .
This proves our claim. As an immediate consequence we have S 0„(4) and 0(L) p for all p in S. 101:7. Weak Approximation Theorem for Rotations. Let V be a regular quadratic space over a global field F and let T be a finite set of spots on F. Suppose çop is given in 0+ (V,,) at each p in T. Then for each e> 0 there is a a in 0+ (V) such that iic — 99pilp< 8 VP ETProof. Express each (Pp as a product of symmetries, 99p = -cti,;;
-
-
where the ur are anisotropic vectors in Vp. Here the number r is even, and we can suppose that the same r applies for all p by adjoining squares of symmetries wherever necessary. Using the Weak Approximation Theorem of § 11E on the coordinates of the vectors /41 in the underlying base x1 ,. . ., x„ we can obtain a vector u, in V such that ilui — Will y is arbitrarily small at all p in T. If the approximation is good enough we
Chapter X. Integral Theory of Quadratic Forms over Global Fields
obtain an anisotropic vector u1 such that
Tui
is arbitrarily close to -c•
291 p
at
each p in T, in virtue of the continuity of the map u -4— ru. Now do all this for i 1, 2, . , r and in this way obtain anisotropie vectors U1 , . . , it of V with Tui ruP p
arbitrarily small for all p in 7' and for 1 i r. If good enough approximations are taken all around we can arrange to have
ru, • • VuTuP
.. • r
•
er P <
at all p in T, in virtue of the continuity of multiplication in LFp (Vp). Put a = rui . Tur E 0 ± (V). Then !lap — Ta p < for all p in T or, in the relaxed notation, o— cppil p < e . q. e. d. 101:8. Let V be a regular quadratic space over a global field F with dim V 3. Then 0(0+ (V)) consists of the set of elements of F which are positive at all real spots p at which V I, is anisotropic. Proof. Let R denote the set of real spots p of F at which Vp, is anisotropic. (Needless to say, R may be empty.) Every element of 0(0+(V)) is a product 6f elements of the form Q 1) • - Q
&)
where the yi are anisotropic vectors of V. Each Q (y i) is negative at those p in R at which Vp is negative definite, hence the above product is positive at these p; the same applies at the spots p of R at which VI, is positive definite; hence the above product is positive at all p in R. Conversely, let cc be any element of F which is positive at all p in R. First let n 4. Take any fixed non-zero element 13 in Q (V). Then a 13 is represented by V p at all p in R; and Vp is universal at all remaining spots on F; hence a 13 is represented by Vp at all p on F; hence a /3 is represented by V, by Theorem 66:3. Hence a ,82 = (a 13) fl is in 0(0 1- (V)), hence a is. Now let n , 3. Scaling V by its discriminant shows that we can assume that the discriminant dV is actually equal to 1. Then VI, is positive definite at all p in R. Let T denote the set of discrete spots p on F at which V p is anisotropic. So R i T consists of all the spots p on F at which VI, is anisotropic. Use the Weak Approximation Theorem of § 11E to find an element ,8 of F which is positive at all p in R and which has the following property: both and —a ,8 are non-squares at each p in T. Then by Proposition 63: 17 we see that <—,8> ± V is isotropic at all p in T, it is also isotropic at all p in R, hence it is isotropic at all p, hence it is isotropic, hence V represents #. Similarly V represents a 13. Then a#2 — (a #) 13 is in 0(0+ (V)), hence a is. q. e. d. 19*
292
Part Four. Arithmetic Theory of Quadratic Forms over Rings
101 : 8 a. If F is a function field, then
101:9. Let V be a regular quaternary quadratic space over a global field F. 'Consider a set of spots T on F such that D4 (Vp) c 0;(V p) for all p in T. Then there is a a in 0;(V) such that ar (24 (V p) for all p in T and ap E .Q4 (V p) at all remaining p on F. Proof. By scaling V we can assume that 1 E Q(V). The set T is a subset of the set of non-dyadic spots p at which Vp is anisotropic. Hence T is a finite set. Let W denote the remaining set of spots p (archimedean or discrete) at which VI, is anisotropic. Again W is a finite set, possibly
empty. Using the Weak Approximation Theorem and the fact that IT, is open in Fp we can find an element 4 in F which is a non-square unit at all p in T and a square at all p in W. Similarly we can find an element in F which is a prime element at all p in T and a square at all p in W. Then z1 is represented by Vp at each p in T since a regular quaternary space over a local field is universal; and z1 is represented by Vp at each p in W since A is a square there and 1 is in Q (V); and Z1 is represented by Vp at all remaining spots p on F since Vp is then isotropic by definition of T and W. Hence Z1 E Q (V) by Theorem 66:3. Similarly E Q (V). Similarly A E Q (V). Hence 1, Li, a, nZI E (V) • There must therefore exists a a in 04 (V) of the form =- <1) (A> (x> r Os A> •
(Here, as in § 95, t(Œ > denotes a symmetry with respect to a vector x with Q (x) a.) Then a is in Qi (V) by definition of 0'4 (V). And ap Q4 (V) at each p in T by Corollary 95: la. And ap E X2 (Vp) at each p in W since we then have =- To> T(i) to> r E 124 (Vp) • And Vp is isotropic at each of the remaining spots p on F, hence .Q4 (Vp) q. e. d. = 01 (Vp), hence ap E ‘24 (V,,) . 101:10. Example. V is a regular n-ary space over the global field F and S is a Dedekind set consisting of almost all spots on F. Consider lattices L,K,... on V with respect to S. We always have cls+L cIsL when n is odd (see Example 82:4). This is also true for even n over local fields (see Corollary 91:4a). It is not true in general for global fields. In fact we claim that there is a lattice K on V with cls+K C clsK whenever n is even. Let us construct such a K. We start with an arbitrary lattice L on V. Write dV s with s in F. We know from Example 66:6 that the Hasse symbol S p V is 1 at almost all p on F. Let W denote a set of
Chapter X. Integral Theory of Quadratic Forms over Global Fields
293
non-dyadic spots on F which consists of almost all spots in S, such that s is a W-unit and Sp V = 1 at all p in W, such that JE := PEJY, and finally such that L p is unimodular at all p in W. Pick spots q and q' in W such that s is a square at q and q', and such that a W-unit is a square at q if and only if it is a square at cr. The existence of such a pair of spots was established in Proposition 65:20. By Proposition 81:14 and Example p E S— (q 'u q') and 92:8 there is a lattice K on V with Kr = L
0 (0- (IC q))
ti q
,
0 (0- (K q,))
where 4q is a non-square unit in Fq. We say that this K has the desired properties, L e. cls÷K C clsK, i. e. 0+(K) a= 0(K). For suppose not. Then there is a reflexion a of V such that a K = K. Write 0 (a) = oc with oc in P. If p E W (q q'), then a E 61(0 (4)) = uF. Similarly, a Ezi gn and cc E F. so ac is a non-square at q, it is a square at q', and ordp a is even at all p in W. Since IF = PE JI-v, there is a fl in F with 2 ordp/3 = ordp a at all p in W. Then oc//32 is a W-unit which is a square at q' but not at q. This is impossible by choice of q and cf. So 0+ (K) = 0(K). Hence cls+K ( clsK.
§ 101C. Special subgroups of On (V) Once again we return to the questions raised in § 56. This time we must describe the groups 12„ n Z,, 0:,11-2„ , 010.;, over the global field F which is now under discussion'. Of course if F is a function field and n 5, then V is isotropic and all is already known. However, some extra effort will be needed before we obtain the general case. 101:11. Lemma. Let X be a finite subset of fr and suppose that at each spot p on F there is a regular ternary subspace of V I, which represents all of X. Then there is a regular ternary subspace of V which represents all of X. Proof. 1) For n = 3 the result is an immediate consequence of Theorem 66:3. Suppose n = 4. Let T be a finite set of spots on F which contains all archimedean and dyadic spots and is such that, at each spot p outside T, the discriminant dV p is a unit, the Hasse symbol S, V,is 1, and every element of X is a unit. By hypothesis there is a scalar yi, E Fp and a regular ternary space Up at each p in T such that
Vp
For the structure of .(2.112n Z„ (n 5) over an algebraic number field see M. KNESER, Crelle's J., 196 (1956), pp. 213-220.
294
Part Four. Arithmetic Theory of Quadratic Forms over Rings
By the Weak Approximation Theorem and fact that F is open in Fr there is a y in fr such that y E y r F for all p in T. Then
- • Q (y ,-)
E F, hence by the
E
E On (V). In particular this proves the proposition when 1 n _‹ 3.
Let us now assume that n >, 4. Since a p E (V r) we can conclude from Proposition 95:2 at the discrete spots and from the archimedean theory of quadratic forms at the archimedean spots that there is a regular
Chapter X. Integral Theory of Quadratic Forms over Global Fields
295
ternary subspace of Vp at each p on F which represents all the scalars Q(311), . . Q(y,.). By Lemma 101: 11 there is a regular ternary subspace W of V which represents all these scalars. Pick xi E W with Q (zi) = Q (y1) for 1 i r. By Proposition 55:3 it is enough to prove that T„ . . • Tzr is in Q„ (V). And this follows easily from the fact that rz, " • Tz, induces an element of Q 3 (W) = 03 (W) on W. q. e. d. Now we can describe the groups (2„ n Zn, etc. We have Q7, n Zt, = {±1 } if —lvp is in Q. (Vp) at each p on F, otherwise 12„ n = 1 v. The group 0,112„ is isomorphic to the group
HO"„ (Vp)1(2„ (Vp) (this follows easily from Propositions 101: 9 and 101:12). In particular Q„ when n + 4, while if n = 4 the group 0 41 /Q4 is a direct product of a finite number of groups of order 2. Finally the group OI/0;, can be destribed by Proposition 101:8 (for n 3): it is the group of elements which are positive at all real spots p at which V p is anisotropic, modulo the group F 2. § 101D. The group of split rotations jv We have to work with idèle groups again, particularly with the groups T p
j F3 11 1
TS pS E,
TS, 2
j
defined in §§ 33D, 331 and 65A. Here we are assuming that S is a Dedekind set consisting of almost all spots on the global field F. The idèle concept can be extended to the orthogonal group in the following way. Start with a regular quadratic space V over the global field F. Take a base x11 . . x„ for V and let all the norms IIIlp on Vp and LFp (Vp) be with respect to this fixed underlying base. Consider the multiplicative group H O+ (V) p ED
consisting of the direct product of all the groups 0- F (V,). A typical element of this group is defined by its coordinates, say
(Ep)pED (Ep (°+ TO) , and multiplication in the direct product is, by definition, coordinatewise. If we are just told that E is a typical element of the direct product, then Ep will denote its p-coordinate. For two such elements E, A we have (E A) p EpA p (E-1) p = EF,1 , for all p in Q. We shall call E a split rotation of the quadratic space V if E is an element of the above direct product with. the property llEPlp = 1 for almost all p.
296
Part Four. Arithmetic Theory of Quadratic Forms over Rings
This definition is independent of the underlying base since any two systems of norms agree at almost all p by Example 101:1. The set of all split rotations is a subgroup of the above direct product by Example 101:4, it is called the group of split rotations, and it will be written / v. 1 The set of all split rotations Z with the property Er 0' (Vp) YpED is clearly a subgroup of hi; we shall denote this subgroup with the letter Jr. It is evident that liv contains the commutator subgroup of jv ; in particular, .rv is a normal subgroup of Jv and the quotient group JviTv is abelian_ Consider a typical element a of 0+ (V). Then a has a localization up p in D. And Ijap il p = 1 for almost all p, by the Product Formula. atech Hence a determines a split rotation (a) = (ap)p ED • The rule a —>-- (a) therefore provides a natural multiplicative isomorphism of 0+ (V) into Iv. We shall call the split rotation Z principal if there is a rotation a such that Z = (a). The principal split rotations form a subgroup Pv of /v . And we have 0+ (V) >---> P. We shall let D stand for the subgroup 0(04- (V)) of F, and we let PD be the group of principal idèles of the form (a) p ED with a in D. In other Pp . words, P D is the image of D under the natural isomorphism F 101:13. Example. Suppose n 3. Proposition 101:8 tells us that D is then the set of all elements of È which are positive at those real spots p at which Vp is anisotropic. The Weak Approximation Theorem of § 11E shows that (F D) is finite. Hence (Pp : PD) < co. We know from Corollary 33:14a that there is a Dedekind set of spots So which consists of almost all spots on F and is such that IF= PFI: whenever S S So. By considering a finite set of representatives of pp, mod PD we see that there is a set So of the above type such that = PD J;,,g whenever S Ç So . Now consider lattices L, K, . . . on V with respect to the set of spots S under discussion. We define the subgroup h of jv by the equation IL = {E E Tv Ep E 0+ (LO V p E S) . If a is an element of 0+(L), then ap E O nf (L p) holds for all p in S by Example 101:6, hence (a) is an element of Pv n h. On the other hand, if (a) denotes a typical element of Pv n IL , then (a L) a p L p = L p holds for all p in S, hence aL — L by § 81E, hence a E 04- (L). Hence the natural isomorphism of 04- (V) onto Pr carries 04 (L) onto P v n IL . So For a discussion and application of the, topology on J v see M. KNESER, Math. Z. (1961), pp. 188-194.
Chapter X. Integral Theory of Quadratic Forms over Global Fields
297
we have the diagram
0+ (V) 1
0+ (L)
Pv
Pv r\ h.
We define the subgroup n of IF by the equation JP, = {i E
IF I ir E 0(0+ (Lp)) V p E .5} .
Take a typical split rotation E and a typical lattice L on V. We know that EL p is a lattice on V r at each p in S, we claim that there is exactly one lattice K on V with Kr ELp for all p in S, and we then define EL to be this lattice K. In order to prove the existence of K it is enough, in view of Proposition 81:14, to show that ELr = Lp for ox,. Then the condition ilEplip— 1 almost all p in S. Put M = o ± • • is equivalent to Er Mi, = M. hence EMr = Mr for almost all p in S. But Mp = Lr for almost all p in S by § 81E. Hence Ev il, = L I, for almost all p in S. Hence K exists. It is unique by § 81E. So the lattice EL is defined; its defining equation is (EL) =ELr Vp ES. Incidentally, note that EL p = Lp for almost all p E S. We have (E A) L = E(AL) V E, A E jv
If a is a rotation of V, then a L = (a) L . The group h can be described as
= {E E Jv I
L = L} .
102. The genus and the spinor genus § 102A. Definition of gen L and spn L We define the genus genL of the lattice L on V to be the set of all lattices K on V with the following property: for each p in S there exists an isometry Ep E O (Vp) such that Kr = Ep L p. The set of all lattices on V is thereby partitioned into genera. We immediately have genK = genL •:* clsKp = clsLr Vp ES. The proper genus can be defined in the same way: we say that K is in the same proper genus as L if for each p in S there is a rotation 4, c O+(V) such that Kr = ErL r. This leads to a partition of the set of all lattices on V into proper genera. The proper genus of L will be written gen+L. We immediately have gen-FK = gen+L cls+Kr = cls+Lr Vp ES.
298
Part Four. Arithmetic Theory of Quadratic Forms over Rings
But we already know that the class and the proper class coincide over local fields. Hence we always have genL = gen-FL . The genus can be described in terms of split rotations:
K E genL
K = EL
for some
in I v
almost all p in S). (to prove this use the fact that K = L We say that the lattice K on V is in the same spinor genus as L if there is an isometry a in 0 (V) and a rotation E in 0' (V p) at each p in S such that = Vp(S. This condition can be expressed in the language of split rotations: there is a a in 0(V) and a X in j'v such that K = L' L. We shall use spnL to denote the set of lattices in the same spinor genus as L. It can be verified without difficulty that the set of all lattices on V is partitioned into spinor genera. It is an immediate consequence of the definitions that lattices in the same class are in the same spinor genus, and lattices in the same spinor genus are in the same genus. So the partition into classes is finer than the partition into spinor genera, and the partition into spinor genera is finer than the partition into genera. We have clsL spnL çT genL . We let h(L) be the number of classes in genL, and g(L) the number of spinor genera in genL. We shall see later that h(L) and g(L) are always finite. We say that K is in the same proper spinor genus as L if there is a 0' (V p) at each p in S such that rotation a in 0+ (V) and a.rotation X K1,
Lp
VpES.
This condition can be expressed in the language of split rotations by saying that there is a a in 0+ (V) and a E in Ty such that K = ci X L, or
gen L spn L
cis L gen#L spi) c/8"L
equivalently by saying that there is a A in Pv and a X in Tv such that K A EL. We shall use spn+ L to denote the set of lattices in the same proper spinor genus as L. It is easily seen that the set of all lattices on V
Chapter X. Integral Theory of Quadratic Forms over Global Fields
299
is partitioned into proper spinor genera. Again we find that the partition into proper classes, is finer than the partition into proper spinor genera, and the partition into proper spinor genera is finer than the partition into genera. We have cls+L C spn -FL S gen+L .
We let h+ (L) be the number of proper classes in gen L, and g +(L) the number of proper spinor genera in genL. We shall see that h+ (L) and g+ (L) are always finite. All lattices in the same genus have the same volume. For consider K E genL. Then Kr -.".# L t, for all p in S, hence (t3K) 1, = »Kr = 13 L i, = (bL) 1, Vp ES, hence D K = b L. We define the volume of a genus to be the common volume of all lattices in the genus. In the same way we can define the volume of a proper class, of a class, of a proper spinor genus, or of a spinor genus, since each of these sets is contained in a single genus. Similarly we can define the scale and norm of a genus, proper class, etc. If a genus contains an a-maximal (resp. a-modular) lattice, then every lattice in that genus is a-maximal (resp. a-modular). 102:1. Example. If E is any element of jv, then E spnL = spnEL , 2' spn-FL = spn+EL . 102:2. Example. Each class contains either one or two proper classes, hence h (L) h+ (L) _. 2h (L) . It is easily seen that each spinor genus contains either one or two proper spinor genera, so g (L) ._. g+ (L) 2g (L) . But we can actually say more, namely that g+ (L) is either g (L) or 2g(L). For suppose that spnL contains two proper spinor genera. Then spn+L C spnL, hence by Example 102:1 we have spn+EL C spn EL for every E in Tv, hence every spinor genus in genL contains two proper spinor genera, hence g-F (L) = 2g (L). In the same way we find g+ (L)= g(L) when spn+L = spn L. 102:3 Example. Consider the genus genL of an a-maximal lattice L on V. We have already mentioned that every lattice in genL is also a-maximal. Now consider any a-maximal lattice K on V. Then Kr and are ap-maximal at all p in S, hence Kr -2! LI, by Theorem 91:2, hence K ( genL. So the genus of an a-maximal lattice on V consists precisely of all a-maximal lattices on V. In particular, all a-maximal lattices on the same quadratic space have the same scale, norm and volume. 102:4. Example. Suppose K ( genL. Consider a finite subset T of the underlying set of spots S. We claim that there is a lattice K' in
300
Part Four. Arithmetic Theory of Quadratic Forms over Rings
cls+K such that IC; = L all p in T. By definition of the genus there is a rotation ipt, E O+(V) at each p in T such that (pr Kt,= L. By the Weak Approximation Theorem for Rotations there is a rotation a in 0+ (V) such that lic — Ta p is arbitrarily small at each p in T. If the approximations are good enough we will have, in virtue of Examples 101:4 and 101:6, (47K) 1, ap K y 991,Kt, = L t,
for all p in T. Then aK is in cls+K, hence it is the desired lattice K'. 1 0 2:5. Example. Let L be a lattice on the quadratic space V, let K be a regular lattice in V. Suppose there is a representation K r -4— Li, at each p in S. We claim that there is a representation K L' of K into a lattice L' in genL. In fact we shall find a lattice L' in gen L such that K L'. To do this we take a finite subset T of S such that Kt, S L all p in S T. Since there is a representation Kr L t, at each p in T, there is an isometry (p p E O (Vr) such that T rL p D Kr. Define L' to be that lattice on V for which VpEST — L' — tinp Ly VpET. —
Then 4D Kt, for all p in S, hence L' D K. Clearly L' E genL. Hence we have proved our assertion. We have the following special case: suppose that the scalar a EF is represented by V and also by L each p in S; then a is represented by some lattice L' in the genus of L. 1 0 2 :6. Example. It is possible for a scalar in Q (V) to be represented by L all p in S without its being represented by L. For instance, consider the set S of all discrete spots on the field of rational numbers Q. Let L be a lattice with respect to S on a quadratic space V over Q with L <1> ± <11>. Then the equation 3= (8/5) 2 + 11(1/5)2 shows that V represents 3, and also that L, represents 3 for all p + 5; but 1.5 also represents 3 by Corollary 92: lb. There is clearly no rational integral solution to the equation E2 + 11772 3. Hence V represents 3, L, represents 3 at all spots p, yet L does not represent 3.
§ 102B. Counting the spinor genera in a genus 1 02: 7.
Proof. 1) First a remark about abstract groups. Let G be any group and let H be any subgroup of G which contains the commutator subgroup of G. If x, y are typical elements of G, then the normality of H in G implies that xH Hx, and the fact that H contains the commutator subgroup of G implies that xyH= yxH, hence the set Hxy is independent of the order of H, x, y. From this it follows for any subgroups X, Y
Chapter X. Integral Theory of Quadratic Forms over Global Fields
301
of G that the set HXY is independent of the order of H, X, Y, and that this set is actually the group generated by H, X, Y. In particular, this applies to the subgroups Pv, h of J. So the group generated by these three groups is equal to Ti7Pv.11, this is a normal subgroup of j v, we can form the quotient j virv Pv h, and we can write down the index (Iv: ji,P v j b). Our next claim is that this index is equal to g+ 2) Consider two typical proper spinor genera in genL. They can be written spn+El L and spri.-1-E2 L with E, E E Iv since EL runs through gen L as E runs through J. We then have spn+Ei L = spn+E2 L if and only if L E spn+EVE2 L, and this is equivalent to saying that L = AT Er' Ez L for some A E Pv , T E Tv. Hence spni- El L = spn+E2 L
E2 E EirvPv.h -
Therefore, if we let E run through a complete set of representatives of distinct cosets of jv modulo ji,P v IL we obtain each proper spinor genus spn+EL in genL exactly once. So we have the formula
g+ (L) -- (Iv: ii7Pvh) for the number of proper spinor genera in a genus. 3) We are going to construct a group homomorphism of Iv into IF/PD jt, in a certain natural way. Take a typical element E in J. Then EL p = L almost all p, hence 0(4) S up F21, for almost all p by Proposition 92:5. We can therefore choose an idèle I in JE with ip
E (4) YPED•
If j is any other idèle that is associated with E in this way, then j E in by definition of the spinor norm. But .7 .1,c n c IDA . Hence the natural images of i and j in J.FIPD J.-k, are equal. We therefore have a welldefined map
0 : Tv obtained by sending E to the natural image of I in jp/PA. It is immediately verified that dis is a homomorphism. 4) Let us show that 0 is surjective. We must consider a typical idéle and we must find E E Iv such that it, E 0(4) for all p in D. By definition of we can assume that it, = 1 for all p in D — S; we define 4, to be the identity map on Vp for all p in D— S. What about the remaining p? Since n 3 we have 0(0+ (V)) = F all p in S by Proposition 91:6 and 0(0+ (4)) urFti for almost all p in S by Proposition 92:5; we can therefore choose E E 0+(V1,) at each p in S, with almost all E in Q 4- (L), such that 11, E 0(Er) for all p in S. Then IiEJI = 1 for almost all p in S. Hence E = (E0 1, 02 is a split rotation. And 0E is the natural image of t in jp/P D.3. So 0 is a surjective homomorphism. .3
302
Part Four. Arithmetic Theory of Quadratic Forms over Rings
5) It is clear that Pv, Pv, j.L are all part of the kernel of ao, hence Pv Pv jj; is. We leave it to the reader to verify that Tv Pv IL is the entire kernel. Hence
g+(L) =
rvPv.h) = Up: PD.* •
q. e. d.
102:7a. g+ (L) = (Jv : ji,Pv h), and g+ (L) divides (JF : 1. for any n Proof. The assumption n 3 in Proposition 102:7 is used only in q. e. d. showing that the map in the proof is surjective. 102 : 8. Theorem. V is a regular quadratic space over a global field with dim V 3, and L is a lattice on V. Then the number of proper spinor genera in genL is of the form 2r with r 0. Any value of r can be obtained by taking a suitable L on V. Proof. 1) Write the discriminant dV in the form dV e with e in P. We know from Example 66:6 that the Hasse symbol Sp V is 1 at almost all spots p on F. We fix a non-dyadic set of spots W on F which consists of almost all spots in S, such that e is a W-unit and Sp V = 1 at all p in W, such that hp = Ppir, and finally such that L I, is unimodular at all p in W. Proposition 92:5 tells us that. Jr 2) Let us prove two formulas that will be needed in the course of the proof. Consider a lattice K on V with respect to S, and let T be any set of spots on l; with T S and hp = Pill:. Then
g+(K) =
Pilf) =. (114: Pilf.) PD.* = (Png (.11: g n PD.* •
This is our first formula. Now let M be some other lattice on V with respect to S, and suppose that J, Then Then g+ (K ) =
CIF:PDTP
= (IF: PD.11)(PDPI P.Dip = g+ (m) (PD1f, : PD.T1P • Hence
g+ (K) = g+ (M) (Jr : j.y r if
PD.*
)1 c
J. This is our second formula. 3) The proof that g+ (L) is a power of 2 is at hand. We have g+ (L) = (Jr :
PA)
Chapter X. Integral Theory of Quadratic Forms over Global Fields
303
by the first formula in step 2). But rir. 2 S it,. Hence
Tr) '2 n g+ (L) (17 .17 12) This is a power of 2 by Proposition 65:7. 4) Next we construct a lattice K on V with g+ (K) = 1. Start with the given lattice L and Use Propositions 81:14 and 91:7 to obtain a lattice K on V with K p = Li, for all p in W and 0 (0+ (KO) =
Then
VpES—W.
frir jf, since
0(0+ (K))=u ,, 1 VI) EW by Proposition 92:5. Hence g n PDJ =Jf. Hence g+ (K) = 1 by the first formula of step 2). 5) We must digress for a moment in order to prove the following: let M and K be lattices on V, let 44 be a non-dyadic spot in S at which
0(0+ (M4)) = u4P, (0+ (Kg)) =.g& , and suppose that Mp = K p for all p in S — cr, then g+ (K) is either equal to g+ (M: or to 2g+ (M). By definition of M and K we have su c nt, hence by the second formula of step 2) it is enough to prove that 2, p n ni hence that uy, : 2. But this last inequality follows easily from the fact that (u4 : = 2. Hence our contention is proved. 6) Now we can complete the proof. It is enough to show how to obtain a lattice K from L with g+ (K) = 2g+ (L). For L is arbitrary, so we can start with an L with g+ (L) = 1, then obtain a new L with g+(L) then an L with g+ (L) = 4, and so on. So consider the given lattice L. Pick q and cr in W in such a way that e is a square at q and g', and such that a W-unit is a square at q if and only if it is a square at g'. The existence of such a pair of spots was established in Proposition 65:20. By Proposition 81:14 and Example 92:7 there is a lattice K on V with Kp = Li, for all p in S (q q') and with
Pal.)
:
0(0+ (K4)) = ft, 0 (0+ (K c()) =.
Clearly 1.15 c J. We claim that TL
nPD.!: C Tfi •
Suppose not. Then .3 S Pal,. Take i E n with ip = 1 for all p ED— q and 14 a non-square unit in Fq. Then i E Pill, hence there is an a in D and a j in If with i = (a) j. So we have a field element a which is a square at cr, a non-square at q, and of even order at all p in W. Now h = 1371y, hence there is an element /3 in F with 2 ordp fi = ordp a for all p in W.
304
Part Four. Arithmetic Theory
of Quadratic Forms over Rings
Then och92 is a W-unit, it is a square at cr, it is not a square at q. This contradicts the choice of q and q'. Hence we do indeed have nrA PD J-1,c c J. So by the second formula in step 2), g+ (K) 7->, 2g+ (L). all p in S — cr, and Let K' be the lattice on V with K'p = L K cr . Then by step 5), g (K) =
(K 1)
Hence either g (K')
or
2g±(K') ;
g+ (K1) = gl (L)
or
2 g+ (L) , or else g+ (K) = 2 g+ (L).
2g+ (L) . q. e. d.
1 0 2:8a. Corollary. The number of proper spinor genera in genL 1. is a power of 2 in general, i. e. when dim V Proof. The case n 3 is covered by the theorem. The case n = 1 is trivial since then genL = L. There remains n = 2. By Corollary 102:7a it is enough to show that (JE : D is a power of 2, and this follows as in steps 1)--3) of the proof of this theorem, q. e. d. 102:9. Theorem. L is a lattice on the regular quadratic space V over the global field F. Suppose that the underlying set of spots S satisfies the I equation J1 PD 0(0+ (Lp)) 2 up Vp ES, then spn 1 L = genL.
Proof. We must show that g+ (L) = 1. We use the fact that g+ (L) divides (IF: Pil). Then J Ç J since 0 (0+ (4)) D u p for all p in S. Hence IF , PD J_Z Ç PD /t. Hence g+ (L) = 1. q. e. d. 1 0 2:1 0. Example. The last theorem can be used to give sufficient conditions under which spn+L genL. Consider the lattice L on the regular quadratic space V over the global field F with dim V 3. (i) First let us examine the condition IF = ppjp:. This is satisfied in the function theoretic case whenever the class number liF (S) is 1 since then IF = PFIlis, with F.D, hence IF = PD .a. It is also satisfied when F is the field of rational numbers for then IF = ./4/4 with F ( ± 1) D, hence I) F _L 1) PD, but (-1) E ji„ hence JE = PDJ. (ii) The local theory has provided sufficient criteria for testing the condition 0(0 -1- (4)) D U p . For instance this condition is satisfied at all p in S whenever L is either modular or maximal on V (see Propositions 91:8 and 92:5, and Example 93:20). Again, it is satisfied at p if there is a 2 Jordan splitting for L p with a modular component of dimension when p is non-dyadic and of dimension 3 when p is dyadic. Example 92:9 shows how to derive from this a simpler, but weaker, criterion involving volumes (the formula given there is for the non-dyadic case only; but a similar, though not identical, result can be obtained at the dyadic spots).
Chapter X. Integral Theory of Quadratic Forms over Global Fields
305
§ 103. Finiteness of class number This paragraph is devoted to a single purpose: the proof that the number of proper classes of integral scale with given rank and volume is essentially finite'. This will imply that the number of proper classes in a (L) of § 102A genus is finite. Hence all the numbers h (L), h+ (L), g(L), are finite. We shall need the counting number Na introduced for fractional ideals in § 33C. For any non-zero scalar a we define N a = N (ao). Thus N (a fl) = (Na) (N )3). By Proposition 33:3 and the Product Formula we have
H,
Noe
pED—S
PCS
104.
103: 1. Let L be a lattice on the abstract vector space V over the global field F, and let 99 be a non-singular linear transformation of V into V such that 99L L. Then (L: 99 L) = N (det 99) .
Proof. 971, is a lattice on V since 99 is non-singular. By applying the Invariant Factor Theorem to L and 921, we can find a base x1 . . for V and fractional ideals a 1 ,. . an and 1)1. , bn, such that L
-1- • • • + bn ien
99L = al x/ ± • • • + a„,cn .
Now write Yi
99X J =
aiixi
E F) .
Then + • • • + bnYn • 97 /By comparing the above expressions for 99L we obtain, from Proposition 81:8,
. . . an = b1 . . . bn det (a ii) . Then by Proposition 33:2,
(L: 6, L) =
al) • • (b.: an)
— (Nal . . Nan)/(Nbl . Nb 7 )
= N(det 92) . q. e. d. 103:2. There is a positive constant y which depends only on F and S, and which has the following property: given any n x n matrix (aii) with 1 Using "reduction theory" it is possible to give a short and elementary proof of the finiteness of class number in the classical situation where the ring of integers in question is the ring of rational integers Z. See G. L. WATSON, Integral quadratic forms (Cambridge, 1960). O'Meara, Introduction to quadratic forms 20
306
Part Four. Arithmetic Theory of Quadratic Forms over Rings
entries in F and det(a ii) Eu, there are elements $1, . . ., them 0, such that faiii + • - - ± ain flip ..- 7
en in o, not all of
i n and for all p in D — S. Proof. 1) Fix a spot go in D — S. By Theorem 33:5 there is a constant C (0 < C < 1) such that the density M (i)1 II ill of the set of field elements bounded by any idèle i satisfies for I
M(i)lliiii > C •
We can assume, by taking a smaller C if necessary, that C is in the value group IFool go . Put y = 4C-1 . This will be the constant of the proposition. 2) Consider the given matrix (a11). Take a non-zero scalar A in o such that all the elements bii = Aa ii are in o. We have to find elements E, in o, not all of them 0, such that .
.
.
,
ibiii. ± ' • ' ± binnip for 1SiSn and for all p in D— S. Form the cartesian n-space V=Fx•-• x F over F and let L be the lattice L . {(a1, .. . , an)lai E o
for
1s 1
n}
on V. Define a linear transformation 9, E L.F. (V) by the equation
where
97 (a1, - • • , an) — (111'. — , i3n)
fi i= f bu m,
Then
(1 .. i
n) .
i
det 92 --- det (b i 1) — An det (a11)
and 92L S L, hence by Proposition 103:1 (L: 9 2 L) . N (det9,) . (N A)n .
3) Construct an idèle i with if p E S Nip= { 1C-1 lAlq. if P = cio if p E D — (S Li q 0) . Thus 11,4 5 C-1 IA fp for all p in D — S since 0 < C < 1. The volume of i is given by ilill = C-1 H Ai, = C-1 (NA) . IAlp
Then by Theorem 33:5 the idèle I bounds strictly more that NA field elements. Hence there are strictly more than (N A)n vectors (oci, ... , oc,) in L which satisfy kilt>
C-1 !Alp
Chapter X. Integral Theory of Quadratic Forms over Global Fields
307
for 1 i S n and for all p in D — S. But (L: 924 . (N A)". Hence at least two of these vectors, say (oh, . . . , an) and (oci, . . . , oc), are congruent modulo 92L. Put ni = ai — cci for 1 5= 1 n. Then (271, • • • , ?In)
E (FL ,
hence we can find El, . . .,
in o such that ni . Li bij el (1 .. i
n) .
i
On the other hand, — fai — ccilp
2 (icxiip ± iceiip) .. 4C -1 !Al p
for! i n and for all p in D— S. In other words, jba + • • • + bi n &lp ,.
71A1p •
as required. q. e. d. 103:3. Lemma. Let V be a regular quadratic space over the global field F, let c be a given fractional ideal. Then there is a finite subset 0 of P such that Q (L) n (1 + 0 for every lattice L on V which satisfies sL C o, DL D c. Proof. 1) We have c S DL S o since sL So. Now the number of fractional ideals bet*een c and o is finite. It therefore suffices to prove the lemma for all lattices L on V which satisfy the condition DL = c (instead of the condition OL D c). 2) First suppose that V is isotropic. All o-maximal lattices on V have
the same volume, let it be b. Consider any lattice L on V of the type under discussion in this lemma. Then L is contained in an 0-maximal lattice M since nL S sL S o. By Proposition 82:11 there is an integral ideal a such that c = a2 b and a M ç L. The ideal a obtained in this way will be the same for all lattices L under discussion since a2 = cb-1. Take a non-zero scalar a in a. Then M S a-1 /, S a-lL . It therefore suffices to prove the following: there is a non-zero field element which is represented by all 0-maximal lattices on V. What is this field element to be? Take a complete set of representatives a1 ,. . ., ak of the group of fractional ideals modulo the subgroup of principal ideals, i. e. of the ideal class group of F at S. Let fl be a non-zero scalar which belongs to all the ideals a1 ,. . . , ak and or', . . . , ail. We claim that every o-maximal lattice M on V represents 132. By Proposition 82:20 there is a splitting M =--- K ± - - • in whichFK is a hyperbolic plane. By Proposition 82:21 and its corollary there is an ideal ai for some i (1 i k), and there is a base x, y for F K with 12 (x) = Q (y) = O, B (x, y) = 1, 20*
308
Part Four. Arithmetic Theory of Quadratic Forms over Rings
such that K a i x •y (Er y /32 Then fix -F. •y y is in K and hence in M. But Q x + y) So M represents )62 as asserted. 3) Now let V be anisotropic. By Proposition 81:5 every lattice L on V can be written in the form o x2 + • • • + ox,, L= x„ is a base for V and ai is one of a finite number of where fractional ideals al, ak. We may therefore restrict ourselves to the following situation: given a fractional ideal a prove that there is a finite subset 0 of F such that Q(L) n 0 + 0 holds for every lattice L on V which satisfies L o, » L = c and which has the form L ax, + o x 2 + • • • + oxn in some base , xn for V. We can assume that a D o. Take a lattice K of the above type, write it in the form K= azi + o z2 + • • • + ozn , then fix it and fix the base z1 .. z„ for the rest of the proof. We have (ID°, sifSo, K=c. Let t be an idèle with iv = 1 for all p in S and 2ti' ft 2 y2 max I B (zi, liv I r ,
for all p in D— S, where y denotes the constant of Proposition 103:2. The idèle i bounds just a finite set of field elements by Theorem 33:4. It therefore suffices to show that every lattice L of the type under consideration represents at least one non-zero field element that is bounded by t. So consider the lattice L expressed in the form L ax l -F o x2 + • - - + oxn
with a o , sL co , oL= c Let 97: V V be the linear transformation defined by the equations 92z5 = ; for 1 S j ft. Thus 99K = L. Put 921 = f aii zi
E F) .
Now bK.c.b L, hence det(B(zi, zi)) is a unit times det (B (x1, xj)), iiuncc det (aii) is a unit. By Proposition 103:2 we can find elements in o, not all of them 0, such that E1 .. . 1 4111 ei
'•'
ain enip
y
Chapter X. Integral Theory of Quadratic Forms over Global Fields
for 1
i
309
n and for all p in Q — S. Put
z = $1 zi. + • • • + $„z„ . Then Q (n) + 0 since g) is non-singular and V is anisotropic. And z is in K, hence fin is in L, hence Q(g)z) 5L Ç o, hence IQ(cpz)l p
1 V p ES .
Now we also have
pz = ni zi where Here we have Ini l t, i y for 1 calculation then gives
1•2( 9)2.)1,,
i
n, =
n and for all p in Q — S. A direct
2 ' n y2 max IB (zi,
for all p in D — S. Therefore Q ((pz) is a non-zero scalar which is bounded by I and represented by L. q. e. d. 103 : 4. Theorem. Let V be a regular quadratic space over a global field. Then the number of proper classes of lattices on V with integral scale and given volume is finite. In particular, the number of proper classes in a genus is finite. Proof. 1) Let c be an integral ideal. We shall actually prove the following: the number of classes of lattices on V with integral scale and with volume containing c is finite. This of course gives the theorem since each class consists of either one or two proper classes. The proof is by induction on n dim V. For n — 1 the result is trivial. Assume it for n — 1 and deduce it for the given n-ary space V. In virtue of Lemma 103:3 it is enough to prove that the lattices L on V which represent a fixed non-zero scalar a and which satisfy sL Co, »L c fall into a finite number of classes. 2) Fix a vector y in V with Q (y) — a and take the splitting V = Fy ± U. By the inductive hypothesis we can find lattices K1, . . K r on U such that every lattice K on U with 5K c 0 1) K 11 2n c
is isometric to one of them. Define lattices L1 ,. , L. on V by the equations oy K i (1 r) . We claim that for each of the lattices L under consideration there is a lattice L i (1 5 i y) and a a in 0 (V) such that crL D L i . Once this has been demonstrated we shall be through for the following reason: we will have B (crL , L i) ci B (c , L) B (L, L) o, hence L Ç aL
; but the number of lattices between L i and L@ is
310
Part Four. Arithmetic Theory of Quadratic Forms over Rings
finite; hence L will be isometric to one of a finite number of lattices; hence the lattices L under consideration will fall into a finite number of classes. 3) So we must find Li and a. By Witt's Theorem we can assume that y E L. Define the sublattice K' = {cix B(x, y) ylx L) of L. Clearly K' is a lattice on U. Put L' == a„y
IC'
where ay is the coefficient of y in L. Thus ay D e since y E L. For each x in L we have ocx B (x, y) y
(ocx B (x, y)
E Li ,
hence ceL L' cL.
Therefore Cit2n C = a:CC (bK 1). But a:a C 6L S o. So bK 1 D 12211 C. Now K' has integral scale since L does. There is therefore an isometry r). Hence there of U onto U which carries K' to Ki for some i (1 j is a a. in 0 (V) such that a L' = avy Iff D Li Then ci L D
D L.
We have therefore found the desired L i and a.
q. e. d. 103:5. Remark. Suppose the global field F and the underlying set of spots S are kept fixed. Let c be a given integral ideal, let n be a given natural number. Then the number of quadratic spaces V of dimension n which can support a lattice L with integral scale and with volume c is finite (at least up to isometry). For let us take a set of non-dyadic spots T which consists of almost all spots in S, such that cp = op for all p in T, and such that hp = PE TT. Consider an n-ary quadratic space V over F and suppose there is a lattice L on V with integral scale and with = c. Put dV = a with cc in F. Then Lp is a unimodulax lattice on Vp at each p in T. Hence the Hasse symbol Sp V is 1 and the order ordp cc is even at all p in T. The information JEzz-- Pill; gives us a fl in F with 2 ordp fi = ordp a at all p in T. Hence we can write dV = s for some s in u ( T), L e. for some T-unit s. But u (T) modulo u (T) 2 is finite by Proposition 65: 6.Therefore there are just a finite number of possibilities for the discriminant dV of a quadratic space V with the given properties. Consider those quadratic spaces V which have the given properties and have fixed discriminant dV = s with s in u (T). Then Sp Vp = 1 at each p in T, hence Vp is unique up to isometry at each p in T by Theorem 63:20. Now the number of quadratic spaces of given dimension and given discriminant over a local field or over a complete archimedean field is clearly finite
Chapter X. Integral Theory of Quadratic Forms over Global Fields
311
(up to isometry). In particular this is true over the fields Fr at each p in D — T. Hence by the Hasse-Minkowski Theorem there are only a finite number of possibilities for V. § 104. The class and the spinor genus in the indefinite case 1 0 4 : 1. Lemma. L is a lattice on the quadratic space V under discussion. Suppose dimV 3. Let T be a finite subset of the underlying set of spots S. Then there is a scalar ts in o which is a unit at every spot in T and has the following property: every element of to n Q (V) which is represented by L 9 at each p in S is represented by L. Proof. By enlarging T if necessary we can suppose that S — T contains only non-dyadic spots and that Lp is unimodular at each p in S— T. Hence Q (L p) . op at each p in S— T by Corollary 92: lb. Take lattices LI, . . . , LA on V, one from each of the classes contained in genL, and let these lattices be chosen in such a way that L ip . Lp for all p in T and for 1 S i 5._ h (this is possible by Example 102:4). Using Corollary 21:2a we can find a A in o which is a unit at each spot in T and such that AL i C L for I i h. Put its -,-, A2. This will be our i.z. To prove this we consider an element a of tto which is represented by V and also by Lp at each p in S, and we must prove that a is represented by L. Now at each p in S — T we have
mitt E p .0 op — Q (Lp) And at each p in T we have 422 EQ (L p) since a is represented by Lp and A.2 is a unit at p. So cep, is represented by V and also by Lp at each p in S. Hence 422 is represented by some lattice in genL by Example 102:5, hence a/.1,2 EQ (Li) for some i (1 i 5_, h). Then a EQ(AL i) .s Q (L) . q. e. d. 1 0 4:2. Definition. Let S be a Dedekind set consisting of almost all spots on the global field F. Let V be a regular quadratic space over F. We say that S is indefinite for V if there is at least one spot p (archimedean or discrete) in D — S at which Vp is isotropic. If Vp is anisotropic at each p in D — S, then we say that S is a definite set of spots for V. 1 0 4 :3. Theorem. V is a regular quadratic space over the global field F with dimV 4, S is an indefinite set of spots for V, and T is a finite subset of S. Let a be a non-zero element of Q (V) and suppose that at each p in S there is a xi, in Vp with Q (zp) -, a such that
iizpiip _.1. YpES—T.
312
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Then for each s > 0 there is a z in V with Q (z) = a such that
z
I
Vp ES—T
and
liz
zpiip<E VPE T.
Proof. 1) By scaling we can assume that a = I. We may take 0 < €<1. Since S is indefinite for V there is a spot go in D S at which Vq , is isotropic. We fix this spot go. By adjoining all discrete spots in (S u go) to T (and hence to S) we see that we can make the following assumption: S consists of all discrete spots when go is archimedean, S u go consists of all discrete spots when go is discrete. Let x11 . . xn, be a base for V which determines all the given norms It It p. Suppose this base is replaced by the base 6 z1 ,. , ô xn where 6 denote norms with respect to this new is any (S — T)-unit, and let base. Clearly liii p = liii for all p in S — T. Now Proposition 33:8 provides us with an (S T)-unit ô which is arbitrarily large at all p in T. In fact 6 can be chosen in such a way that lizpli p 1 for all p in T, and hence for all p in S. In effect this allows us to assume that the given norms satisfy lizpli p 1 for all p in S. We define the lattice L with respect to S on V by the equation L = o + • • + o x,, .
If p is in S and x is in Vp, then x E L p if and only if 11.x1I p 1. So each of the given vectors xi, is assumed to be in L. The required vector z will actually be found in L, i. c. it will satisfy lIzil p 1 for all p in S. We adjoin to T all those spots p in S which are either dyadic or such that L not unimodular. The enlarged set T is still finite. It suffices to prove the theorem for the new T (since s < 1). We therefore assume for the rest of the proof that qo is either discrete or archimedea.n, that S u go contains all discrete spots on F, that every spot p in S T is non-dyadic with L p unimodular, and that xi, E Lp for all p in S. We illustrate these facts in the spot diagram
apoikiedeen
discrete
2) Pick x E V with Q (x) = 1. By Witt's theorem there is a rotation (pp in 19+ (Vp) at each p in T such that zp =(pp•x. By the Weak Approximation Theorem for Rotations there is a rotation 41 in 0-F (V) with (ppil p arbitrarily small at each p in T. Hence [lax — zpil p
!la — (pplir
Chapter X. Integral Theory of Quadratic Forms over Global Fields
313
is arbitrarily small at all p in T. So there exists a vector y in V with Q (y) = 1 and lb/ — zpl! p <s Vp E T. This implies that IlYllp 1 for all p in T. By the Product Formula, = 1 holds at almost all p; hence by the Strong Approximation Theorem there is a A in 0 such that IIYII:+ I
{
V PE T V p ES—T.
1 Put y = Ay. Then Ily — zp li p < e holds for all p in T. The rest of the proof now consists in using this A and this y to find a vector z in L with Q(z) = 1 and such that liz — O p < s for all p in T. Once this is done we shall have our vector z and we shall be through. It is easily seen that A and y have the following properties: we have a scalar A and a vector y such that A E 0, v E L, Q (v) = A2 with 1 2—lj,<e VpET. Only these properties of A and y will be used in the construction of z. 3) Obviously A is a unit at all spots in T. Let TA denote the set of spots p in S at which A is not a unit, i. e. at which I Al p < 1. Then TA is a finite, possibly empty, subset of S — T. Take the splitting V = Fv _L U. Then L r U is a lattice on U with respect to S, and L p n Up is a lattice on Up with respect to p. (All localizations are taken in Vp .) It is easily seen, say by Theorem 81:3, that (L n U) p = L p n Up VrES. So by Proposition 81:14 there is a lattice K on U such that Kp = L p n Up Vp ES— T and K p Lp n Up with liKpil p ‹s Vp ET. Thus K Ç L. This lattice K has the following properties at the different spots in S: at each p in S — (T u T A) the lattice K p is unimodular since the unimodular lattice L p is split by the unimodular sublattice o p v, hence by Corollary 92: lb Q (K p) =op V p ES— (T T a) ; at each p in TA the lattice L p is unimodular, hence one can use Proposition 82:17 to find a binary unimodular sublattice of Lp that contains y, this sublattice splits L p by Proposition 82:15, hence Kip contains a binary unimodular sublattice, hence by Corollary 92: lb Q (Kp ) up V p TA; illYiip
Part Four. Arithmetic Theory of Quadratic Forms over Rings
314
at each p in T the lattice Kp will contain a maximal lattice of some norm, hence by the last part of Example 91:3 there is an element ap in Fp with < lapip < 14Ip such that Q (K ) 2: ap up V p T • By Lemma 104:1 there is a scalar ts in o which is a unit at all spots in TAU T and which has the following property: every element of tto which is represented by U and also by Kp at each p in S is represented by K. We let To denote the set of spots p in S at which IA is not a unit,
e. at which Lui p < 1. Clearly To is a finite, possibly empty, subset of S — (Ta v T). We now have the spot diagram j.
Mr=
NV.k■
N'kk. dihrete
I
archimedeall
4) We claim that there exists a scalar /3 in F such that 11 — filp 5- liuip 5- 1 , iflip 1 , 1 — 2,2 /32 E Q(K p) Vp ES and flip < 8 V P ET and 1_12 132 E Q(up )
v p 0-2—s
In order to prove this we use the Very Strong Approximation Theorem and its corollary (§ 33G) to find a 13 in F in which approximations are made in the following way at the various spots on F: 1, 1. if p ES — (T TAU T) make in, /32I p !duly, 2. if p E To make 1 )1— 2.-4 1p so small that 11 3. if p E TA make Ply 1, 4. if p ET make /3— 2-4 (1 + exp)lp so small that 11 — 131p < e and
1 1 — #21r 5. if
(To make AN so large that
-(1 ___ ) 2 (l2) = 2. 2 /32 (1
1/ 2.2 (l 2) E
6. if p is a real spot in D— (S v q0) at which Up represents 1 make fi2 E 1/31p so small that 1 7. if p is one of the remaining real spots in D— (S v q0) make I p1p so large that —(1 — 2.2 (l 2) E 8. if p is a complex spot in D— (S v cro) nothing special is needed. Using the results of step 3) it follows easily by direct calculation at each stage of the above approximation that the element /3 chosen in the above way satisfies the conditions stated at the beginning of step 4).
Chapter X. Integral Theory of Quadratic Forms over Global Fields
315
5) Consider the element fi with the properties stated at the beginning of step 4). Then 1— 2.2 P2 E Q (U at all p in D, hence 1 _22 9 2 E Q (u) by Theorem 66:3. Now 1— 2,2 P2 E ,u 0. And 1 _ a2 P2 is represented by Kv at all p in S. Hence 1 --- 2,2 EQ(K) by the choice of ,u in step 3). Take y in K with Q (y) = 1— 2 p 2 and put z = 13 y ± y. Then z is an element of L with Q (z) = 1. Also llYllp < e for all p in T by the choice of K in step 3), hence we have )
vilp max — v1lp , 11Y11p) = max — #1p fivilpi 1IY!!0 < for all p in T. Therefore z satisfies the conditions mentioned in step 2). So the theorem is proved, q. e. d. 104:4. Strong Approximation Theorem for Rotations. V is a regular quadratic space over the global field F with dim V 3, S is an indefinite set of spots for V, and T is a finite subset of S. A rotation Tv is given in 0' (Vp) at each p in T. Then for each s > p there is a rotation o. in 0' (V) ilz
such that 92 pli p< 8 VPET
and
11016 1 VpES—T. Proof. The proof consists of two major steps, the first for n 4 and the second for n = 3. We let . , xn denote a base for V which determines all the norms pp. We can assume that 0 <e < 1. la) We start with n. 4. First suppose that each cpv is a "short commutator", i. e. that each rpv has the form = TUp TVp rliprVp
where ruv and reyv are symmetries of Vv with respect to the anisotropic vectors up and yv of V. By using the Weak Approximation Theorem on the coordinates of the up in the base x11 . , x„ we can find a vector u in V which is arbitrarily close to up at each p in T. So by the continuity of the map x Tx (see Example 101:2) we can assume that Tu is arbitrarily close to ruv at each p in T. Now do the same thing with the vv to obtain a vector y in V such that the symmetry Tv is arbitrarily close to Tvp at each p in T. Then by the continuity of multiplication in L,Fr (Vv) we can assume that we have found vectors u and y in V such that Tu Tv ru — TUp TVp TUp TVP < 8 p E T. Put w = ;u. Then Tw = Tv Tu Tv. Hence we have a pair of vectors u and w in V with Q (u) = Q (w) 0 such that llrurto — (Pp I P < 8 VPET• Therefore it is enough to find o E (V) with llolip = IL for all p in S — T
such that
— ru'rwiip ‹E
Vp E T.
316
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Let L be the lattice L = o + • • ox, on V. We can assume that u and w are in L (replace them by Au and Aw with a suitable A in o if necessary). Let T1 be a finite subset of S T such that L is unimodular with Q (u) = Q (w) a unit at each p in S (T1 u T). By Theorem 104:3 we can find a vector z in L with Q (z) Q (u) = Q (w) such that z is arbitrarily close to u at all p in T1 and arbitrarily close to w at all p in T. So by the continuity of the map x rz we can assume that the symmetry rz with respect to this vector z is arbitrarily close to ru at all p in T1 and arbitrarily close to -r at all p in T. Put a ru t... Then a is in 0' (V). By the continuity of multiplication we can suppose that we have found a a which is arbitrarily close to 1vp at all p in T1 and arbitrarily close to ru Tto at all p in T, in other words such that
1110.11p = 1 11 Icl — wlIp<S
vP E VPET•
It remains for us to prove that --- 1 for all p in S — (T1 u T). Now at each such p both u and z are elements of Lp with Q (u) =Q (z) a unit at p, hence ru and; are units of Lp by § 91B, hence 0.4, is a unit of L p , hence Doll y = 1. 1b) We continue with the case n 4, but now we let the gap be arbitrary elements of the commutator subgroups Du (Vp) at the spots p in T. We can express each Tr in the form 99p =
- - • V;
where /pi'? is a short commutator in On (Vp) (see Proposition 43:6). We can assume that the saine r applies at all p in T, by adjoining trivial short commutators wherever necessary. By step 1 a) we can find vi in 0' (V) with Vi arbitrarily close to 4, at each p in T and My = 1 at each p in S T. Do this for i = 1, ..., r. Then 1 = tp* is the required element. 1c)We conclude the case n 4 by considering arbitrary elements Tr in O' (Y u) at each p in T. We can assume that n — 4, for otherwise 0;,(Vp) = Q (Vp) for all p in T and this is covered by step 1b). We enlarge T by adjoining to it all spots p in S T at which Vp is anisetropic, and we define tpp to be 1 -vp at each of the new spots in T. In effect this allows us to assume that VI, is isotropic, hence that (4 (Vp) =f24(Vp), at all p in S — T. By Propositions 95:1 and 101:9 we can find e in Q't (V) such that gp epp is in D4 (Vp) at all p in T. Let T1 be the set of spots in S - T at which > 1. By step lb) we can find a in (4 (V) such that IkiJi = at all p in S (T1 u T), with a arbitrarily close to e „ at each p in T1 , and arbitrarily close to e p 99p at each p in T. Then at each p in S — (T1 u T) ,,
,,
Chapter X. Integral Theory of Quadratic Forms over Global Fields
317
we have 11'o 5- 11 Q-111p 114 p 1) hence 11 e1 ; at each p in T1 we have e-1 1 Vp11 p Ile' (0. — lip I e'ff p i — elf p and this is arbitrarily small, hence li e-1 o p 1 ; similarly we can obtain Ile-1 — iPp p < e at each p in T. Hence a is the required element. 2 a) Now the case n = 3. By scaling V we can assume that the discriminant dV is 1. All norms are determined by the base xi., x2, x3 for V. Let 4 4, x; be an orthogonal base for V, and let 11 11 pi denote norms with respect to this new base. Take a finite set of spots T' with T T'ES such that 1 lip II lip' and Q (xl) E up at each p in S —T' for 1 5_ I 3. Define rpp = 1 -pp for all p in T' —T. If we can prove the theorem for the new set T' and the new system of norms, then we can find a a in 0' (V) with 114 piioIi1 pf for all p in S— T' and 994; arbitrarily small for all p in r n particular we can make — 1v pil p < 1 for all p in T' —T and lla (pa p < e for all p in T; this means that = 1 for all p in T'— T, and hence for all p in S — T. In other words, it is enough to prove the theorem for the enlarged set of spots T' and the new system of norms. In effect this allows us to make the following assumption: the base xi., x2, x3 used in determining the norms 11 11 p is an orthogonal base for V in which V
with al, a2 in up at each p in S — T. 2b) Suppose the localization Vp of V at p is replaced by some other localization V10, of V at p. Then there is a unique isometry Vp >--> which induces the identity map on V. This isometry determines an algebra isomorphism LFp (Vp)›.-0- LFp (Vg) in a natural way. The algebra isomorphism so obtained preserves isometries, rotations, symmetries, spinor norms, norms 1 li p, and localizations of elements of L7 (V). It sends 0 ( Vp) 0(n) and 0' (Vp) to 0' (V). From this it follows that the theorem holds for the given localizations Vp (p ED) if and only if it holds for some other system of localizations v,o, (p E D). Hence we can assume that each Vp is taken in the localization Cp of a quaternary quadratic space C which is defined in the following way: fix a 4-dimensional F-space C containing V, fix a vector 1a in C — V, and make C into a quadratic space over F in such a way that C
Fla V with QM. 1 .
The localization Cp of C is taken and fixed at each p in D. The space Cp is regarded as a quadratic space over F. and Vp is the subspace of Cp spanned by V. A norm 11 ll p is put on Cp with respect to the base 1a, x1, x2, x3. This induces the given norm on vp.
318
Part Four. Arithmetic Theory of Quadratic Forms over Rings
Recall that in the theory of quaternion algebras (§ 57) we started with a 4-dimensional vector space and a base, we fixed them, we put a multiplication on the vector space by means of a multiplication table, and we called the resulting object a quaternion algebra. The initial choice of vector space and base was quite arbitrary. Now do all this starting with C = Fla ± ± F x2 F and make C into the quaternion algebra ( —alF1—a2 ) in the defming base l a, x1, x2, 4Similarly regard Cr as the quaternion algebra ( —31Fp ' —a2 ) in the defining base la, xi, x2, h. Clearly the quaternion algebra Cr is the Fr-ification of the quaternion algebra C. Let bar stand for conjugation in C and in each Cr. The space C can be regarded as a quadratic space in two different ways. First we have the quadratic structure used in defining C, namely C = F1a 1 V with Q (l c) = 1. Secondly, we have the quadratic structure associated with the quaternion algebra C in the manner of § 57B. It is easily seen that these two quadratic structures are the same, namely B (x, y) lc = -y (xj", + yi)
V x, y E C
The same holds for each Cr. 2c) So much for the logical niceties. We see from the multiplication table used in defining Cr that iixYlip 5- iixiip Nip holds for any x, y in Cp at any p in S — T. And there is a positive constant r such that frYlip 5 r IlYlip for any x, y in Cr at any p in T. By Proposition 57:13 there is a vector zy in Cr at each p in T with Q (zr) = l and Tr x = x.:17,1 Apply Theorem 104:3 to the quadratic space C. We obtain a vector z in C with Q (z) = 1 such that lizii r 5_ 1 at all p in S — T and Ilz—zrli r < s' at all p in T, where e' is a positive number with < ilZplip ., el < 81112 for all p in T. By Proposition 57:13 there is an element a in 0' (V) with crx=zxz -1 VxEV.
Then for any x in V and any p in S — T we have
(1 0'4 p = Ilz x
iilp
Piip 5-
Chapter X. Integral Theory of Quadratic Forms over Global Fields
319
hence Ilaxill p 5_ 1 for 1 i 3, hence kril l, 5_ 1, hence Ilall p = 1. A similar argument (essentially an argument of continuity of multiplication) gives fc — < e at all p in T. q. e. d. 1 0 4:5. Theorem. V is a regular quadratic space over the global field F with dimV 3, S is an indefinite set of spots for V, and L is a lattice on V with respect to S. Then
= spn-FL and clsL = spnL
Proof. We shall prove clsL = spnL. The equation cls+L = spn+L is done in the same way. So we consider a lattice K in spnL and we must prove that K E cIsL. By the definition of the spinor genus we have a in 0(V) and Er in 0' (V r) at each p in S such that K=iEL Vp ES.
Then (0.-1 K) t, = L'p L I, for all p in S. But 0.-1 K E cisK. Hence we may assume that a = l y. So we now have Kt, = Ep L p Vp ES. Fix a base x„ for V, let norms lip be defined with respect to this base on VI, at each p in S, and let M be the lattice M =0 x1+ . . . Take a finite subset T of S such that K r = L p = Mr V p ES —T .
Then by the Strong Approximation Theorem for Rotations there is a rotation e in 0 (V) such that !la p = 1 for all p in S — T, with Ile — Evil p arbitrarily small for all p in T. The first condition informs us that (eL)=eM=K
p
VP EST.
The second condition informs us, in virtue of Example 101:4, that we can obtain er Li, = K t, at all p in T by taking good enough approximations in the selection of e Then L) r . K r for all p in S. Hence e L = K. Hence K E clsL. q. e. d. 1 0 4:6. Example. The purpose of this example is to show that the condition dim V 3 in Theorem 104:5 is essential. Let F be the field of rational numbers, let S be the set of all discrete spots on F, let V be a hyperbolic plane. So S is certainly indefinite for V. And o is the ring of rational integers. Take a lattice L = ox oy on V and suppose that (e
.
L
Put K
o (4 x)
0 (4 y).
10 9\ inx,y.
So
K10 k9
9\
m 321
1
— 4
x' 4y.
320
Part Four. Arithmetic Theory of Quadratic Forms over Rings
We claim that but K E spn+L . It is easily seen that L does and that K does not represent 2, hence K cls+L. Let us show that K E spn+L. In fact we shall find rpr E 0' (Vr) at each p in S sucli that K r = rppi.r. If p is non-dyadic this is trivial since then K r — LI, by definition of K and L. So consider the 2-adic spot p. Then Kr and Lr are unimodular with nKr = nLp = 2or, hence by Corollary 82:21a there is a vector w in Vr with Q (w) = 0 and K
cls+L
L r = o r x or w ,
K r = or x)
or (4w) .
By Example 55:1 we have a rotation Tr of V (prx=
x and
itself such that
Irpr w, = 4w , •
and the spinor norm of this rotation is equal to T N. Hence Tr E0' (V r). Moreover epr L r = K. Hence K E spn+L. 1 0 4 :7. Remark. The reader will be able to use Theorem 106:13 to show that the assumption of indefiniteness in Theorem 104 :5is essential. 1 0 4 :8. Remark. Theorem 104:5 tells us that we have cls+L = spn+L in the indefinite case in 3 or more variables. On the other hand, Theorem 102:9 and Example 102:10 give sufficient conditions for spn+L and genL to be equal. Hence we have sufficient conditions for determining when cls+L and genL are equal. Now the genus is completely described by the local theory. Hence we have certain sufficient conditions that can be used to describe the proper class (in general these conditions are not necessary). 104:9. Theorem. V is a regular quadratic space over the global field 3, and S is an indefinite set of spots for V which satisfies F with dim V the condition jp .1)D 11, where D = 0(04-(V)). Suppose that L is either a modular or a maximal lattice on V with respect to S. Then
cls+L = genL
Proof. We have cls+L spn+L by Theorem 104 : 5. But spn+L = genL by Theorem 102:9 and Example 102:10. Hence cls+L = genL. q. e. d. 104:10. Theorem. V is a regular quadratic space over the field of rational numbers Q with dim V 3, and S is an indefinite set of spots for V. L and K are lattices of the same norm and scale on V with respect to S. Suppose that L and K are either both modular or both maximal. Then
cls+L = ds+K
Proof. In the modular case we have genL = genK by Theorems 92:2 and 93:29. In the maximal case we note that both L and K must be
321
Chapter X. Integral Theory of Quadratic Forms over Global Fields
a-maximal where a is the common norm a = nL = nK, hence genL = genK by Theorem 91:2. So in either case we have genL = genK. And jr, =PDJ by Example 102: 10. hence by Theorem 104:9,
cls+L genL = genK = cls+K
q. e. d. 1 0 4: 11. Example. Let L be a unimodular lattice on a binary space V over the field of rational numbers, with respect to the set of all discrete spots S on Q. Suppose S is indefinite for V. (Thus L is a Z-lattice and V cr, is isotropic.) Prove from first principles that
L
<1> I <-1> or L
1\ 0) '
and hence cls+L gen L. 104: 12 Example. Let L be a unimodular lattice with respect to Z on a quadratic space V over the field of rational numbers Q. Suppose that L has norm Z and that V co is isotropic. Use the Hilbert Reciprocity Law, Theorem 63:20, and the Hasse-Minkowski Theorem to show that
V a'
< 1>
• <1> <
-
1>
<
-
1> .
Deduce that
L <1>
<1 > <- 1 > ± • ±
• •
(- 1
>•
§ 105. The indecomposable splitting of a definite lattice Consider a lattice L in a quadratic space V. We say that L is decomposable if there exist non-zero lattices K1 and K2 contained in L such that
L = KIL
K2
.
If L is not decomposable we call it indecomposable. It is clear that every lattice L is the orthogonal sum of at most n indecomposable components, where n is the rank of L. A splitting of this sort is called an indecomposable splitting of L. I 0 5 : 1. Theorem. L is a lattice on the regular quadratic space V over an algebraic number field F. Suppose that the underlying sets of spots S is definite for V and also that it contains all dyadic spots on F. Then the components L 1 , . . L. of an indecomposable splitting L = L 1 _1_ • • • I L. are unique (but for their order). Proof. 1) By scaling V we can assume that s L Ç o. We shall again need the counting number Na of § 33C. As in § 103 we put N a = N (ao) for any a in F. We have
Na= 11-T , = H PCS
O'Meara,
Introduction to quadratic forms
I OE!i,
pED—S
21
322
Part Four. Arithmetic Theory of Quadratic Forms over Rings
The assumption that sL Co implies that N (Q x) is a natural number for all non-zero x in L. . 2) We shall call a vector x in L reducible if there are non-zero vectors y and z in L with B (y , z)= 0 such that x = y + z. We call x irreducible if it is not reducible. Our purpose here is to show that every vector in L is a sum of irreducible vectors of L. First consider the sum y + z of non-zero vectors y and z of V with B (y, z) = O. Then Q (y z) = Q (y) + Q (x). Here dim V 2, so definiteness implies that the spots in D — S are either real or discrete. If p is any real spot on F, then VI, is anisotropic and so Q (y) and Q (x) are either both positive or both negative at p, hence IQ (y + z)l p > IQ ()lp. Now consider a discrete spot p in D — S. We say that [Q (y + z)l p IQ (y)p. Suppose not. Then IQ (y + z)I p < IQ (y)k, and so 0 mod Q (y) m p .
Q (y) + Q (z) Q (y
Hence by the Local Square Theorem Q (y) is a square times —Q (z). This is of course absurd since Vp is not isotropic. We have therefore proved that IQ (y + z)l p (y)l p holds at all p in D — S with strict inequality at least once. Hence N (Q (Y + 4) > N(Q(Y))
Similarly N (Q (y x)) > N (Q (x)). The proof that every vector x in L is a sum of irreducible vectors of L is now done by induction on the natural number in = N (Qx): if x is reducible write x = y + z with y and z in L and 1 5 N (Q y) in — 1, 1 N (Q m — 1.
3) We put an equivalence relation on the set of non-zero irreducible vectors of L as follows: write x y if there is a chain of irreducible vectors x
z2,
=y
(q
1)
in which B zi+i) =1= 0 for 1 i E q— 1. Let C1, C2, . . . denote the equivalence classes associated with this equivalence relation. Let K1, K2, . . . denote the sublattice of L that is generated by the vectors in 0 since B (C 1, = 0 for I j. Hence the C1, C2, . .. Then B(K i, number of equivalence classes is finite, say C1 , , C. And the sum of the lattices K1 , . . , K t is actually an orthogonal sum: IC1 J - • • ± K . Now we proved in step 2) that every vector in L is a sum of irreducible vectors of L. Hence L=
j_ • • j_ K
4) Consider x in Then x is in L , e. x is in Li j_ • • j_ L,.. But x is an irreducible element of L. Hence x falls in exactly one of the above components of L, say x E LI. It follows from the definition of that G1 c 4. Hence K1 .0 L1 . Hence each K i is contained in some Li. Since L
Chapter X. Integral Theory of Quadratic Forms over Global Fields
323
is also equal to Ki j. • • ± K i we therefore see that each L i is the orthogonal sum of all the Ki contained in it. But L i is indecomposable. q. e. d. Hence each L i is a Kj . 105:2. Example. Let V be a quadratic space over the field of rational numbers Q and suppose that V has a base x1,. . . , x4 in which V (1> ± (1> ± (1> ± (1> . Consider an underlying set of spots S consisting of all non-dyadic spots on Q and let L be the lattice L = ox1 j. • • • ± øx4 . It follows easily from the local theory that S is a definite set of spots for V. And the four vectors 1 1 -2- (x ± x2 + x3 ± x4) , -2- (x1 -± x2 — x3 + x4)
form a base for L in which L (1> ± (1> 1 (1> 1 (1> . So the assumption made in Theorem 105:1 that S contain all dyadic spots cannot be relaxed. The reader may easily verify that L also has a splitting in which L <2> ± (2> j_ <1> ± (1> . It is also easily verified that the assumption of definiteness in Theorem 105: 1 is essential.
§ 106. Definite unimodular lattices over the rational integers We conclude with some very special results on the class of a unimodular lattice of small dimension over the ring of rational integers Z. If the -underlying set of spots is indefinite for the quadratic space in question, then the class is equal to the genus and all is known. This is no longer true in the definite case (for instance we shall see that there is a unimodular lattice of dimension 9 whose genus contains two distinct classes). We shall confine ourselves to the definite case. The situation then is this. F is now the field of rational numbers Q, S is the set of all discrete spots on Q, and Z is the ring of integers o (S) of F at S. Lattice theory is with respect to S. As usual we use the same letter p for the prime number p and the prime spot 5 which it determines. V is a regular n-ary quadratic space over Q and it is assumed that S is a definite set of spots for V. This is equivalent to saying that the localization Voc, is either positive definite or negative definite since S consists of all discrete spots on Q. By scaling we can assume that Vœ is actually positive definite. We shall assume that there is at least one O'Meara, Introduction to quadratic forms
21*
324
Part Four. Arithmetic Theory of Quadratic Forms over Rings
unimodular lattice on V. Now the discriminant of any unimodular lattice over the ring Z is either + 1 or —1, so in the situation under discussion it has to be +1. In particular, dV = 1. Furthermore there is a unimodular lattice on the localization V„ at each discrete spot p, hence Sp V = 1 for p 3, 5, 7, . . . ; but S,,,, V = 1 since Vc„ is positive definite; hence S2 V = 1 by the Hilbert Reciprocity Law. Hence by Theorem 63:20 and the Hasse-Minkowski Theorem we have
V
<1> ± • • • ± <1> .
We therefore assume throughout this paragraph that V has the above form. The symbol 4 will denote the n x n identity matrix. Thus we have V ,.==- In . We call a lattice D on V completely decomposable if it splits into an orthogonal sum of lattices of dimension 1. Thus in the situation under discussion the unimodular lattice D on V is completely decomposable if and only if it has the matrix 4. § 106A. Even and odd lattices
Consider a unimodular lattice L with respect to Z on the given quadratic space V over Q. Then 6L = Z and 2 Z Ç nL S Z so that nL is either Z or 2Z. We call the unimodular lattice L odd if nL Z, we call it even if nL = 2Z. Thus L is even if and only if Q(L) c 2Z. An analogous argument leading to an analogous definition can be employed for unimodular lattices over Z2 (but there is no distinction between odd and even over Z„ when 5> 2). It is easily seen that the unimodular lattice L over Z is even if and only if the localization L2 over Z2 is even. 106:1. V is a regular quadratic space with matrix In over Q. Then there is an even unimodular lattice with respect to Z on V if and only if 0 mod8. Proof. 1) In the course of the proof it will be found necessary to use the 2-adic evaluations of the Hilbert symbol, also the fact that 1, 3, 5, 7 are representatives of the four square classes of 2-adic units, and finally the fact that 5 is a 2-adic unit of quadratic defect 4 Z2. All these things were established in the statement and proof of Proposition 73:2. As in n
§ 93B, we let A (a, /3) stand for the 2-adic matrix (II 131) . 2) First suppose there is an even unimodular lattice on V. Then there is an even unimodular lattice on the localization V2, hence by the local theory (Examples 93:11 and 93:18) we must have either V2
(A (0, 0)) j - • • ± (24 (0, 0)>
V2
(A (0, 0)) ± • • • j_ (A (2, 2)) .
or
Chapter X. Integral Theory of Quadratic Forms over Global Fields
325
But d V2 = + 1 and each of the numbers —1, —3, +3 is a non-square in Q2, hence we must actually have V2 2_-'
(A (0, 0)) ± • • - j..
with n 0 mod 4. A computation of Hasse symbols over Q2 shows that n 4 mod 8 is impossible. Hence n 0 mod 8. 3) Conversely let us assume that n 0 mod 8. Then the criterion of Theorem 63:20 informs us that QI (0, 0)) ± • - i Hence by Proposition 81:14 there is a lattice L on V with L ,, In when V2
p = 3, 5, 7, . . .
and L2
I • • • 1
This L is clearly unimodular and even,
q. e. d.
§ 106B. Adjacent lattices
We continue our investigation of unimodular lattices with respect to Z on the quadratic space V with matrix I , over Q. For any such lattice L we define a (L) to be the set of all unimodular lattices K with respect to Z on V such that K,
L a, for
a
p
3, 5,.7,
.
Note that aa (L) = (aL) for any 0- in On (V). 106:2. Suppose n 5 and let K and L be any two unimodular lattices on the space V under discussion. Then there is a lattice J in cls+K such that J E (L). Proof. Let S' denote the set of all non-dyadic spots on Q and put Z' = o (S'). Then Z Z' S Z„ holds for each odd p. The Z'-module L' generated by L in V is clearly a lattice on V with respect to Z', and L S L' Ç L„, holds for each odd p. Hence L' has the property that Li,' =4 for each odd p. Now do the same with K to obtain a lattice K' with respect to Z' with analogous properties. The set of spots S' is indefinite for V since the localization V2, having dimension 5, is isotropic. But L' and K' are unimodular of norm Z'. Hence cls+K' by Theorem 104:10. So we can pick a E (V) such that a K' = L'. Put J = a K. So J E cls+ K . And for each odd p we have
L„ . J = (a K)„ = a„K„ = a.K;„ = (a K'),, q. e. d. Therefore j E (L). If K and L are unimodular lattices on V, then K is in (L) if and only if the invariant factors of K in L are of the form
a
a
Z,
. , 2r* Z 21a
326
Part Four. Arithmetic Theory of Quadratic Forms over Rings
(with r 0 and r„ 0). We say that K is adjacent to L if the invariant factors of K in L are equal to 1 -2- Z, Z, . . ., Z, 2Z (assume n 2 for this definition). Suppose K is adjacent to L. Then it follows immediately from the definitions that L is adjacent to K, that K is in (L), that there is a base xi, . . . , xn, for V in which
a
{ L . Zx l. + • • • + Zxn 1 K . Z (-T xi) + • • • + Z (2 xn) , 1
that 2 L C K -2-L, and finally that aK is adjacent to aL for any a in 0„(V). The number of lattices adjacent to L is finite since the index (--L : 2 L) is finite. 2 106:3. Suppose n-=--- 0 tnod8 and let L be any unimodular lattice on the space V under discussion. Then there are even and odd unimodular lattices on V which are adjacent to L. Proof. By the criterion of Theorem 63:20 we know that V2 is the orthogonal sum of hyperbolic planes, hence by Example 93:18 we have
either L2 2f—
L2-2-'--
± • • • 1 (A (1, 0» .
or
There is clearly a unimodular lattice K2 on V2 which is either odd or even as desired, and whose invariant factors in L2 are equal to 1
ay
6
. .7
2 dr...2, i...2, . . • , Z2,
2 Z2 .
Take a lattice j on V with if p . 2 if p . 3, 5, 7, . . This j is the required lattice on V. q. e. d. 106:4. Suppose n 2 and let K and L be any two unimodular lattices (L) and K+ L. Then there on the space V under discussion with K is a chain of unimodular lattices
Ea
L — L., .12, • • • , Tt - K on V with j i+1 adjacent to L. Proof. It is enough to find a chain of unimodular lattices
P, J2, . . . , »
(over Z2)
Chapter X. Integral Theory of Quadratic Forms over Global Fields 327
on the localization V2 with P = L2 and ft = invariant factors of ja+ 1 in ji are equal to I -
y
K2,
and such that the
Z2; Z 2; • • • 3 Z fir 2Z 2 •
For then we can define lattices fi (1 i t) with respect to Z by the equations T {If if p . 2 .1 fp= L„ if p . 3, 5, 7, . . and L = L., j2 , .. ., j t = K will provide the desired chain from L to K. If n = 2 we have
I
L2 =---- Z2X + Z2y K2 = Z2 (-2X7)
+ Z2(2r y)
and the required local chain is obvious. We may therefore assume that n 3.
By Proposition 106:3 we can assume that both L and K are odd, hence that the localizations L2 and K2 are odd. The local theory (Theorem 93:29) then gives an isometry a in 0„(V2) such that K2 = CIL2. By expressing a as a product of symmetries on V2 we see that it is enough if we assume that a is itself a symmetry such that aL 2 + L2. Thus a = *ru with u a maximal anisotropic vector in L2, say. It follows easily from the local theory that there is either a 1- or a 2-dimensional unimodular sublattice M of L2 which contains u. In the first event we would have ;42,2 = L2. Hence M is actually binary. Take the splitting L2 = M IN. Then N = -44 N C K2 and we therefore have a splitting K2 = M' IN. Write M — Z 2 x + Z2 y M' . Z2 (j-) + Z2(2r y) . {
The required local chain from M to M', hence from L2 to K2, is now q. e. d. obvious, We are beginning to see how the theory of adjacent lattices can be used in determining the unimodular classes on V. We start with a fixed lattice D of the form D
<1> j_ - • • ± <1>
on V. By the finiteness of class number (Theorem 103:4) and by Proposition 106:2 there are lattices K,,. . . , K t in (D) such that
a
cls+Ki, . . ., cls+Kt
are all the distinct proper unimodular classes on V (at least for n 5). So we shall certainly achieve our purpose if we can find the lattices
Part Four. Arithmetic Theory of Quadratic Forms over Rings
328
, K. How is this to be done ? By a step-by-step construction of adjacent lattices starting with the fixed lattice D. First we construct by a certain procedure all the lattices adjacent to D (we have already seen that these are finite in number); then all lattices adjacent to these; and so on. By Proposition 106:4 we can obtain K1,. , K t by steps of the above type. The essential technical features of this procedure are the following: the method of construction, deciding when all Ki have been found, eliminating duplications. The rest of the chapter is devoted to these matters and their application up to n — 9.
§ 106C. Rules of construction Here we give rules for operating with adjacent lattices. Throughout this subparagraph we assume without further reference that K and L are two unimodular lattices with respect to Z on the quadratic space V with matrix I„ over Q. 106:5. The following assertions are equivalent: (1) K is adjacent to L
L) = (3) 13 (K n L) = 4Z. Proof. That (1) implies (2) is immediate from the definition of adjacent lattices and the definition of the volume b. To prove that (2) implies (3) we use the fact that LA = L holds for any unimodular lattice L; then (2)
b (K
(K L) =
(IC4* n .0) = » (K+L)*== 4Z.
That (3) implies (1) follows easily from the Invariant Factor Theorem. q . e. (1. 10 6 : 6. Suppose K is adjacent to L and let y be a vector in L — K. Then (1) L K = Zy K (2) L Zy (L K) (3) L K = (Zy K)#. {w EKIB (w , y) E Z} Proof. (1) We have KcZy+K(L+K. But there are no lattices properly between K and L K by Proposition 106:5. Hence Zy K = L K. (2) We have L r KCZy (L r K) L. But there are no lattices properly between L r. K and L. Hence Zy (L n K) = L. (3) Since L and K are assumed to be unimodular we have .
L nK=L4 r
= (L
(Zy K)#
The second equation follows from the definition of the dual (Zy
K) 4t. q. e. d.
106:7. Let y be any vector in (21- K) K with Q (y) in Z. Then there is exactly one unimodular lattice J . on V that contains y and is adjacent
Chapter X. Integral Theory of Quadratic Forms over Global Fields
329
to K. This lattice can be constructed by forming
Zy + (Zy + K)* .
Proof. The fact that J, if it exists, will have the above form is an immediate consequence of Proposition 106:6. Incidentally, this also proves the uniqueness. We now define a lattice J. on V by the above equation and we prove that the J so defined is first of all unimodular, and secondly that it is adjacent to K. A direct computation shows that B (J, .7) Ç Z, hence 6 J Ç Z. And if we write K =Zw + Zx 2 + + Zxy, with 2y = mw (m E Z), then Zy + K ç_ Z
+ Z; + • • • + Zx n
so that Z b
(Z y
K)
44- Z ,
hence 1) (Zy + K) = Z, hence b (Zy + K) 4* = 4Z. Now y is not in K# = K, hence it is certainly not in (Zy + K)#, hence 4Zc jS Z. Therefore »j = Z. SO J is unimodular. Since (Zy + n K we must have 4 Z b(J nK)CZ, hence K)= 4Z, hence J is adjacent to K. q. e. d. 106:8. Suppose K is adjacent to L, let y be a vector in L K, let z be any vector in V. Then y + z is in L — K if and only if z is in LnK—{wEKIB(w,Y)EZ}•
Proof. If y + z is in L — K, then by the definition of adjacent lattices we have (L + K : K) = ( 2-1Z: Z) --- 2, hence z = (y z) — y is in K, hence z is in L n K. The converse is obvious, q. e. d. 106:8a. Let L and L' be unimodular lattices adjacent to K, let y be an element of L — K, let y' be an element of L' — K. Then L = L' if and only if
y— y' E K
with
B(y, y') EZ Z.
We now have a procedure for constructing all unimodular lattices adjacent to K. Start with a complete set of representatives of T K modulo 2K. Eliminate all those on which the quadratic form Q is not integral and also all those that fall in K. Let y. . . , y,. be the remaining representatives. Form the lattices Zyi + (Zyi + K) 4t for 1 i r. This gives every unimodular lattice adjacent to K at least once. And when do we have duplication, say L = L i ? If and only if yi yi E K with B (yi, yi) E Z.
330
Part Four. Arithmetic Theory of Quadratic Forms over Rings
106:9. Suppose K is adjacent to L and let y be a vector in L — K which has the form y = w ± z with Q (w) .1,
zEK,
B (y, w) E(--i i Z) — z.
Then cla = clsK. Proof. In fact we shall prove that K .--. r L where T. is the symmetry Tw. Here TX = X— 2 B(x, w) w V x EV . Note that 2w = 2y — 2z is in L n K. If we let m denote the odd integer 2 B (y, w) we have Ty
= y — rnw = (y— w) — (m — 1) w = z— (m — 1) w ,
hence Ty is an element of K. And for each x in L n K we have B (x , w) = B (x , y) — B(x, z) E Z
so that Tx is an element of L n K. Hence r L = r (Zy + (L n K)) C K . q. e. d. § 106 1) . 1...< is _‹ 7 1 06: 1 O. Let L be a unimodular lattice on the space V under discussion, and sup ose that 1 n S 7. Then L'-'---- In' Proof. By adjoining a lattice with matrix 1-7 _n to L we can assume, in virtue of Theorem 105:1, that we have n = 7. Let D be a completely decomposable unimodular lattice on V. By Proposition 106:2 we can assume that D E(L). By Proposition 106:4 it will be enough if we prove that clsD = cls.L under the assumption that D is adjacent to L. Take an orthogonal base for D, say D= Zxj . .1. • • - 1 Zx7. Let y be any vector in L — D. By reordering the above base for D if necessary,
we can write 1
y ---- T (a1x1 + • • • ± ax) + ar+1 x,÷1 + • • • + a7x7 with all ai in Z and ai, . . . , a, odd. Here we must have r= 4 since i (xi + x2 + x3 ± x4) and Q (Y) E Z and since y is not in D. Put w — -y then define z by the equation y . w + z. We have Q (w) = 1 ,
1 zED, B(y,w)E(- 2-Z)—Z,
provided the sign ± in the definition of w is correctly chosen. Then cis /. . cis D by Proposition 106:9. q. e. d.
Chapter X. Integral Theory of Quadratic Forms over Global Fields
§ 106E. The matrix On(/'
331
8, 12, 16, . . .)
Throughout this subparagraph we assume that the dimension n of the space V under discussion is one of the numbers 8, 12, 16, . . . We take a completely decomposable unimodular lattice D on V and we fix it. We take a base for D in which D = Zxl ± • • ± Z x7, and we fix it. Our first purpose is to show that there is always an indecomposable unimodular lattice adjacent to D, that all such lattices are in the same proper class, that all have the matrix
where P,,, is the matrix
in particular that the matrix On just defined is unimodular for the specified values of n. Needless to say, every lattice on V with matrix On can be obtained from a suitable completely decomposable lattice in the above way. Put y = (x1 + • • + x,). Then y is a vector in G1 D) — D, Q (y) E Z since n O mod4, hence by Proposition 106:7 there is exactly one unimodular lattice E on V which contains y and is adjacent to D. (This E will be a lattice with the desired properties.) Now D = Z (2y) Z x 2 Z (x3 — x2) + • • • + Z (xn xn--1) • And Zy + Z (24 Z (x3 — x2) + • • • + Z (xn xn_1) is a unimodular lattice adjacent to D which contains y, hence E = Zy Z (2x2) + Z (x3 — x2) + • • • + Z (x„— . By inspection we see that the matrix of E in the base y, 2x2, — (x3 — x2), (x4 — x3), . . . , (xn — is equal to On . Before we prove that E is indecomposable we must give a coordinate description of the vectors in E.
332
Part Four. Arithmetic Theory of Qua.dratie Forms over
106: 11.
E A,x, is in E if and only if A i — Aj EZ, E A i E2Z
Ai E hold for 1
Rings
I
n and I j
n.
Proof. Put x= E A.;. First suppose that x is in E. Then x is in -2- D and so all A i E-21 Z. And xi — x E (Zy D)it CE, so B(x,xi—x) E Z, 1 LT A i E Z. hence A i — Ai E Z. And B(x, y) E Z so that T Now the converse. If one Ai is in (1 1 Z) — Z, then so are they all since A i — Ai E Z for all j. In this event replace x by x + y. We may therefore assume that all A i are in Z. We still have EA E 2Z. Now x = E A is an element of D and it has the property B(x, y) E Z, hence xEDnEçEby Proposition 106:6. q. e. d. The last proposition can be used to show that E does not represent 1. Suppose we have E A = 1 with E A i xi in E. If one A i is in Z, then so are they all, and in this event Al = ± 1, A 2 = A 3 = • • • = A n = 0, say. But then E A i is not in 2 Z, and this contradicts the fact that EA; is in E. So E cannot represent 1 in the above way. On the other hand, if each A i is in (--ï l Z) — Z we have Ai+ • • • + 2,1!
2.
So E does not represent 1 in any way, as we asserted. Why is E indecomposable ? Consider the vectors x2 — xi, x3 — x2, . .
xn,
+ xn in E.
Call them yl, , y„. These vectors obviously span V. Now Q (yi) - 2, and E does not -represent 1, hence each yi must fall in exactly one component of the indecomposable splitting of E. But B(y i, yi+i) =f 0. Hence all yi fall in the same component of the indecomposable splitting. Hence this component has dimension n, so it is equal to E. Hence E is indecomposable.
Finally we must prove that any other indecomposable unimodular lattice E' adjacent to D is in the same proper class as E. It is enough to prove that E' is in the same class as E since the symmetry rxi_ z, is a unit of E. Let Y
, 2
(al
+
+ an xn)
(ai E Z)
Chapter X. Integral Theory of Quadratic Forms over Global Fields
333
be an element of E'— D. If one of the a i were even we would have xi E E' n D S E' by Proposition 106:6 and E' would then be split by Zxi. So all ai are odd. By making suitable choices of sign we can write a= 1 — 4 b i (b i EZ) bn xn) E2D S E' r D. Then n. Put z = 2 (bi + • for 1 ± • • • ± x.)
is also in E'—D. Define an isometry o. E O, (V) by the equations axi = ±xi for 1 .< I n. Then a D=D and ay = y' z. Hence o.E is a unimodular lattice adjacent to D —crD which contains y'd- z. Therefore o..E = E'. So E' is in the same class as E. 1 06:1 2. Suppose n = 8. If L is a unimodular lattice adjacent to E, then L is either in the class of D or in the class of E. Proof. Take a vector z in L — E. Then 2z is in E and so we can write z
1
(Ai xi ± • • • + A 8 x8)
with all
8 A .- AiEZ,
Ai
A i E2Z.
1) First suppose that all A. are even integers. Then E (T A.) is an odd integer since z is not in E. Put z' =x1. Then z z' is in E and B (z, z') E Z. Now L is a unimodular lattice adjacent to E with zEL—E, and D is a unimodular lattice adjacent to E with z' E D— E. Hence L = D by Corollary 106:8a. 2) Next suppose that all A i are odd integers. Then E (--A) is again an odd integer since z is not in E. Put , z =
(—x1+ x2 + • • • + x8) •
There is a unimodular lattice adjacent to E that contains z', and by Corollary 106:8a this lattice must be equal to L, hence z' is in L — E. Then z' = —
+ (x1 + • • • ± x8) .
So clsL = clsE by Proposition 106:9. 3) Let us complete the case of integral A i. Since 12(z) E Z we can suppose that exactly four A i are odd, say A 1, . . , A4 odd and A 8,...,A 8 even. By successively adding and subtracting suitable vectors of the form xi ± xi to z we can find a new z in L E which hag one of the four forms z
2(
+ X2 ± x3 ± x4),
Z
= 2(
+ X2 + X3 ± X4) + X5
334
Part Four. Arithmetic Theory of Quadratic Forms over Rings
(apply Proposition 106:8). In the first two cases L contains a vector z with Q (z).. 1, hence it is decomposable, hence L c4. 18 by Proposition 106:10, hence clsL = clsD. In the second two cases write 1
z = -j-- (T xl + x2 + x3 + x4) + ( ± xl + x5) , and apply Proposition 106:9 to find cIsL = clsE. 4) Finally assume that all A i are in (-- Z) — Z. All the vectors 4x1, ± x1 +
X2 , . • . , ± Xi ±
x8
are in E; using Proposition 106:8 we can successively add and subtract certain of the above vectors (with correct signs attached) to obtain a new zinL—E of the form 1 z .— 4 ( ± a xl ± x2 ± • • • ± x8)
where a is one of the numbers 1, 3, 5, 7. Now a cannot be 1 or 7 since Q (z) E Z. If a — 3 we have Q (z) . 1, hence L splits, hence L-.:-41. 1 8 by Proposition 106:10, hence cIsL = clsD. If a . 5 we write 1, „ z = -71- (T a xi. ± x2 ± • • - ± x2) ± 2x1
and apply Proposition 106:9 to find clsL = clsE.
q. e. d.
§ 106 F. Summary 106:13. Theorem. V is an n-ary quadratic space over the field of rational numbers Q with positive definite localization V.., and L is a unimodular lattice with respect to Z on V. Suppose 1 _< n __< 9. Then L has
exactly one of the forms /n , 08, Os I Ii.
Proof. The fact that L cannot have more than one the above forms is clear from Theorem 105:1 and the fact that any lattice with matrix 08 is indecomposable. If 1 s n s 7 we have L .'_". in by Proposition 106:10. Next let n = 8. Consider adjacent lattices I and K on V. If JL.-_-- /8, then K-...- 4 1f K splits and K .s...A 08 if K does not split. If J L--.._• 08, then K has matrix /8 or 08 by Proposition 106:12. Hence by Propositions 106:2 and 106:4 we have L a= 1.8 or L ae Os. Finally n = 9. Here it is enough to prove that L represents 1, since then L will split. This will be achieved if we can prove the following: if J and K are adjacent unimodular lattices on V such that J represents 1, then K also represents 1. Since J represents 1 we have a splitting J . Zxi i_ J'. Let y be any vector in K — J. If J' ae 08 we write 2y = tnxi + z with in E Z and z E r ; then J' is even and so the fact
Chapter X. Integral Theory of Quadratic Forms over Global Fields
335
that Q (y) E Z implies that m E 2Z; hence E (Zy j)* --JnKSK and K therefore represents 1 as required. Otherwise r /8 ; then we have a base for J in which = Z ± • • - Z xi, ; this time write 2y = m1 x, + m8 x8 (mi E Z) ; then at least one nii must be even since Q (y) E Z; for this xi we have xi E (Zy j) 4* = I n K _S K and K therefore represents I as required. q. e. d. With some perseverance it is possible to extend these results' to higher dimensions using the general principles of this paragraph. For 1 n S 13 the distinct proper classes on V are determined by lattices of the form , 08 I 1 2 112 One obtains new indecomposable lattices in 14, 15, and 16 dimensions. 9
45
See M. KNESER, Arch. Math. (1957), pp. 241-250. For an example of the classical approach using "reduction theory" we refer the reader to B. W. JONES, The arithmetic theory of quadratic forms (Buffalo, 1950).
Bibliography Our original intention was to provide a complete bibliography with full documentation, but we were soon discouraged by the complexity of such a task. Instead we have decided to give the following short list of books and articles' in order to enable the reader to trace individual results to their source and to lead him to other fields of interest in number theory and the theory of quadratic forms. E. Arlin, Algebraic numbers and algebraic functions (Princeton University, 1951). E. Artin, Geometric algebra (New York, 1957). C. Chevalley, The algebraic theory of spinors (New York, 1954). L. E. Dickson, Studies in the theory of numbers (Chicago, 1930). L. E. Dickson, History of Me theory of numbers vol. III (New York, 1952). J. Dieudonné, La géométrie des groupes classiques (Berlin, 1955). M. Eichler, Quadratische Formen und orthogonale GrupPen (Berlin, 1952). H. Hasse, Zahlentheorie (Berlin, 1949). B. W. Jones, The arithmetic theory of quadratic forms (Buffalo, 1950). M. Kneser, Klassenzahlen indefiniter quadratischer Formen in drei oder mehr VerAnderlichen, Arch. Math. (1956), pp. 323-332. H. Minkowski, Gesammelte Abhandlungen (Berlin, 1911). • O. T. O'Meara, Integral equivalence of quadratic forms in ramified local fields, Am. J. Math. (1957), pp. 157-186. C. L. Siegel, -Ober die analytische Theorie der quadratischen Formen, Ann. Math. 36 (1935), pp. 527-606; 37 (1936), pp. 230-263; 38 (1937), pp. 212-291. B. L. van der Waenlen, Die Reduktionstheorie der positiven quadratischen Formen, Acta Math. (1956), pp. 265-309. G. L. Watson, Integral quadratic forms (Cambridge, 1960). A. Weil, Adeles and algebraic groups (Institute for Advanced Study, 1961). E. Witt, Theorie der quadratischen Formen in beliebigen Kiirpern, Crelle's J. 176 (1937), pp. 31-44. 1 Additional references to specific points of interest are given in footnotes in the text.
Index Abelian extension, 35 Absolute units, 73 Adapted base, 212 Adjacent lattices, 326 Adjunction, 225 hyperbolic, 257 Algebraic function field, 58 Algebraic number field, 58 Algebras, 114 central, 119 Clifford, 132 commutative, 115 compatible with quadratic spaces, 131 division, 115 E-ification of, 130 Hasse, 150 homomorphism of, 115 ideals in, 119, 121 quaternion, 142 reciprocal, 128 similar, 128 simple, 119 splitting of, 128 tensor product of, 116 Almost all elements of a set, xi Analytic map, 3 Anisotropic quadratic space, 94 Anisotropic vector, 94 Approximation theorems, strong, 42 strong for rotations, 315 very strong, 77 weak, 8 weak for rotations, 290 ARTIN, E., viii, 52, 54, 78, 79, 141, 190, 336 Automorphs, 101 intepal, 223 proper, 101 proper integral, 223 Axis of a ternary rotation, 105 Bilinear form, alternating, 82 hermitian, 82 symmetric, 82
Bilinear mapping, 113 multiplicative, 116 Brauer group, 128 CASSELS, J. W. S., 187 Cauchy sequence, 9 CHEVALLEY, C., 112, 158, 173, 336 Class number, 48 finiteness of, 79 for quadratic forms, 309 Class of a lattice, 222 of a matrix, 221 Classical ideal theory, 81 Classification of quadratic forms, complex, 156 for finite fields, 158 general, 152 global, 189 global integral, 320, 334 local, 170 local integral, 247, 248, 267, 277 real, 156 Clifford algebra, 132 derived base for a, 135 even element of a, 135 Coefficient of a vector in a lattice, 210 Compatibility of an algebra with a quadratic space, 131 Completeness, 9 Completion, 9 notation for, 11 Components of a splitting, 89, 224 Constant field, 55 Cyelotomic extension, 191 absolute, 191 Decomposable lattice, 321 completely, 324 Decomposition field, 36 Decomposition group, 35 Dedekind domain, 52 Dedekind set of spots, 42 Defining base of a quaternion algebra, 142 Defining matrices, 122
338
Index
Definite quadratic space, 154 Definite set of spots, 311 Degree of inertia, 24, 37 Derived base for a Clifford algebra, 135 DICKSON, L. E., 336 DIEUDONNi, 3., 82, 336 Dimension of a lattice, 213 Direct sum of lattices, 224 Dirichlet unit theorem, 77 Discriminant of a free lattice, 222 of a quadratic space, 87 of a set of vectors, 87 Dual base, 92 Dual of a lattice, 230 Dyadic local field, 158
M., vii, 336 E-ification, 129 EICHLER,
of a quadratic space, 131 of an algebra, 130 Eisenstein polynomial, 64 Elementary transformation, 100 Equivalence of matrices, 85 integral, 221 Even element of a Clifford algebra, 135 Even lattice, 324 Finiteness of class number, 79 for quadratic forms, 309 Fixed space of an isometry, 103 FLANDERS, FL, 190 Forms of degree d, 87 Free lattice, 213 Frobenius automorphism, 64, 193 Function field, 58 Fundamental invariants of a lattice, 263 same set of, 264 Fundamental type of a lattice, 263 Galois extension, 35 General linear group, 84 Genus of a lattice, 297 Geometry, orthogonal, 82 symplectic, 82 unitary, 82 Global field, 58 rational, 55 separable generation of a, 58 Global square theorem, 182 Group of units of a lattice, 223
B ASSE, H., 336 Hasse algebra, 150 Hasse norm theorem, 186 Hasse symbol, 167 Hasse-Minkowski theorem, 189 Ecxil, E., 52 Hensel's lemma, 26 Herbrand's lemma, 179 Hilbert symbol, 164 Hilbert's reciprocity law, 190, 201 Hyperbolic plane, 94 Hyperbolic space, 99 Ideals, 44 class group, 48 class number, 48 decomposable, 48 divisibility of, 44 finiteness of class number, 79 fractional, 44 g.c.d. of, 45 indecomposable, 48 integral, 44 1.c.m. of, 45 localization of, 217 multiples of, 44 order of an ideal, 50 p-ification of, 217 prime, 48 principal, 44 product of, 45 relatively prime, 45 value of, 45 Mee notation, 68-69, 80, 172-174, 176 Identification process, 130 Indecomposable splitting of a lattice, 321 Indefinite quadratic space, 154 Indefinite set of spots, 311 Index of a quadratic space, 99 Integers, algebraic, 81 at a set of spots, 42 at a spot, 20, 174 p-adic, 57 rational, 54-55 Integral equivalence of matrices, 221 Integral matrix, 214 Integral polynomial, 25 Integral representation of matrices, 221 Invariant factor theorem, 214 Invariant factors, 216
Index Involution, 96 Isometry, 84 at a spot, 186 fixed space of an, 103 of lattices, 220 spinor norm of an, 137 Isotropy, 94 at a spot, 186 Isotropy theorems, for finite fields, 158 for function fields, 188 general, 153 global, 187 local, 170 IWASAWA, K., 78
viii, 239 N., 186 JONES, B. W., 335, 336 Jordan chain, 246 Jordan invariants, 276 Jordan splitting, 243 saturated, 264 Jordan type, 245 JACOBOWITZ, R., JACOBSON,
KNESER, M., 293, 296, 335, 336 LANG, S., 170 Lattice, 209 adapted base for a, 212 adjacent, 326 adjunction of a, 225 almost free, 213 base for a, 212 class of a, 222 coefficient of a vector in a, 210 completely decomposable, 324 components of a, 224 decomposable, 321 dimension of a, 213 direct sum of lattices, 224 discriminant of a, 222 dual, 230 even, 324 free, 213 fundamental invariants of a, 263 fundamental type of a, 263 genus of a, 297 group of units of a, 223 in a space, 209 indecomposable splitting of a, 321 isometry of a, 220 localization of a, 217
339
matrix of a, 221 maximal, 234 maximal vector of a, 210 modular, 231 n-ary, 220 norm generator of a, 253 norm group of a, 251 norm of a, 227 odd, 324 on a space, 209 orthogonal base for a, 225 orthogonal splitting of a, 224 orthogonal sum of lattices, 224 p-ification of a, 217 radical of a, 225 rank of a, 212 regular, 225 representations of a, 220 saturated splitting of a, 264 scale of a, 227 scaling a, 238 spinor genus of a, 298 unimodular, 231 volume of a, 229 weight generator of a, 253 weight of a, 253 Legendre symbol, 205 Local degree, 31, 37 Local field, 59 dyadic, 158 non-dyadic, 158 2-adic, 276 Local norm, 32 at a spot, 176 Local square theorem, 159 Local trace, 32 Localization, 186 convention, 290 of a lattice, 217 of a linear transformation, 289 of an ideal, 217 Matrices, class of, 221 equivalence of, 85 integral, 214 integral equivalence of, 221 integral representation of, 221 of a lattice, 221 of a quadratic space, 85 representation of, 85 unimodular, 214 Maximal ideal at a spot, 21
22*
340
Index
Maximal lattice, 234 Maximal vector of a lattice, 210 MrismowsKI, H., 336 Modular lattice, 231 Multiplicative bilinear mapping, 116 Negative definite quadratic space, 154 Negative element at a real spot, 20 Negative index of a quadratic space, 155 Non-dyadic local field, 158 Norm (algebraic), xi of a quaternion, 143 Norm generator, 253 Norm group, 251 Norm of a genus, etc., 299 Norm of a lattice, 227 Norm of a linear transformation, 286 of a matrix, 287 of a vector, 12, 285 Number field, 58 Odd lattice, 324 O'MEARA, O. T., 247, 267, 277, 336 Order of an element, 38 of an ideal, 50 Ordinary absolute value, 14, 19 Orthogonal base, 90, 225 Orthogonal complement, 90 Orthogonal group, 84 center of, 108 commutator subgroup of, 106 generation by symmetries, 102 special subgroups of, 100, 141, 156, 158, 280, 293 Orthogonal sets, 83 Orthogonal splittings, 88, 224 p-adic numbers, 57 integers, 57 spot, 55 p-adic topology, 7 p-ification, 186 convention, 290 of a lattice, 217
of a linear transformation, 289 of an ideal, 217 Pigeon holing principle, 69 Place, 29 POLLAK, B., viii Positive definite quadratic space, 154 Positive element at a real spot, 20
Positive index of a quadratic space, 155 Power series expansion, 39 Prime, 55 Prime element, 38 Prime function, 55 Prime ideal, 48 Prime number, 54 Prime polynomial, 55 Prime spot, see under Spots Principle of domination, 4 Product formula, 66 Quadratic defect, 160 Quadratic form, 87-88 of a quaternion algebra, 145 Quadratic map, 83 Quadratic reciprocity law, 207 Quadratic space, 83 anisotropie, 94 definite, 154 discriminant of a, 87 E-ification of a, 131 hyperbolic, 99 indefinite, 154 index of a, 99 isotropic, 94 matrix of a, 85 n-ary, 83
negative definite, 154 negative index of a, 155 orthogonal base for a, 90 positive definite, 154 positive index of a, 155 radical of a, 91 regular, 91 representations of a, 83, 84 scaling a, 83 totally isotropic, 94 universal, 83 Quaternion, 142 conjugate, 143 norm of a, 143 pure, 142 scalar, 142 trace of a, 143 Quaternion algebra, 142 defining base of a, 142 quadratic form of a, 145
Radical, 91, 225 Radical splitting, 93, 226 Ramification index, 22, 37
Index Ramified extension, 60 fully, 60 Rank of a lattice, 212 Rational function field, 55 constant field of a, 55 integers of a, 55 Rational global field, 55 integers of a, 55 Rational number field, 54 integers of a, 54 Reciprocal algebra, 128 Reducibility criterion, 28 Refle.xion, 100 Regular lattice, 225 Regular quadratic space, 91 Representation, 83 for lattices, 220 for matrices, 85 for quadratic spaces, 84 integral, for matrices, 221 Residue class field, 21 notation for, 21 obtained by natural restriction, 23 representative set of a, 39 RIEHM, C., viii, 277 Root of unity, 190 primitive, 190 Rotation, 100 axis of a ternary rotation, 105
SAM, C. H., viii, 250 SAMUEL, P., 52, 58 Saturated Jordan splitting, 264 Scale of a genus, etc., 299 Scale of a lattice, 227 Scaling a lattice, 238 a quadratic space, 83 a vector, 83 SCHILLING, O. E G., 60, 168 SCHMIDT, F. K., viii Sequence, limit of, 3 Series, sum of, 3 SIEGEL, C. L., 187, 336 Similarity of algebras, 128 Spinor genus of a lattice, 298 Spinor norm of an isometry, 137 notation, 138 Split rotation, 295 principal, 296 Splittings, of algebras, 128 of lattices, 224
341
of quadratic spaces, 88 radical splittings, 93, 226 Spots, 7 archimedean, 7 carrying a spot, 7 complete, 9 complex, 19 Dedekind set of, 42 definite set of, 311 discrete, 37 dividing, 7, 8 fully dividing, 8 indefinite set of, 311 induced, 7 infinite, 55 integer at a spot, 20, 174 isometry at a spot, 186 isotropy at a spot, 186 local norm at a spot, 176 non-archimedean, 7 p-adic, 55 real, 19 square at a spot, 174 trivial, 7 unit at a spot, 20, 174 unramified at a spot, 60, 176, 191 Square at a spot, 174 Strong approximation theorem, 42 for rotations, 315 Symmetric bilinear form, 82 Symmetry, 96 TATE, J., 190 Ten-field construction, 196 Tensor products, 113 notation, 114, 118 of algebras, 116 Topological field, 2 ring, 287 vector space, 286 Triangle law, 1 Unimodular lattice, 231 Unimodular matrix, 214 Unique factorization theorem, 49 Units, absolute, 73 at a set of spots, 42, 73 at a spot, 20, 174 of a lattice, 223 Universal quadratic space, 83 Unramified at a spot, 60, 176, 191
342
Index
Valuated field, 1 Valuations, 1 archimedean, 1 carrying a valuation, 3 complete, 9
discrete, 37 equivalent, 5 general, 29 non-archimedean, 1 normalized, 19, 59, 65 trivial, 2 valuation ring, 20, 29 Value group, 2 Very strong approximation theorem, 77 Volume of a genus, etc., 299
Volume of a lattice, 229 Volume of an idèle, 68 WAERDEN, B. L. VAN DER, 125, 336 WATSON, G. L., 305, 336 Weak approximation theorem, 8 for rotations, 290 Wedderbum's theorem, 125 Weight generator, 253 Weight of a lattice, 253 WEIL, A., 336 WHAPLES, G., viii, 29, 54, 78 WrIT, E., 151, 336 Witt's theorem, 97, 98 ZARISKI, 0., 52, 58 ZASSENHAUS, H., 137
Die Grundlehren der mathematischen Wissenschaften in Einzeldarstellungen mit besonderer Beriicksichtigung der Anwendungsgebiete Eine Auswahl 127. Hermes: Enumerability, Decidability, Computability. DM 39,-; US $14.50 128. Braun/Koecher: Jordan-Algebren. DM 58,-; US $21.50 129. Niko4m: The Mathematical Apparatus for Quantum Theories. DM 144,-; US $53.30 130. Morrey jr.: Multiple Integrals in the Calculus of Variations. DM 78,-; US $28.90 131. Hirzebruch: Topological Methods in Algebraic Geometry. DM 38,-; US $14.10 132. Kato: Perturbation Theory for Linear Operators. DM 79,20; US $29.30 133. Haupt/Kiinneth: Geometrische Ordnungen. DM 76,-; US $28.20 134. Huppert: Endliche Gruppen I. DM 156,-; US $57.80 135. Handbook for Automatic Computation. Vol. 1/Part a: Rutishauser: Description of ALGOL 60. DM 58,-; US $21.50 136. Greub: Multilinear Algebra. DM 32,-; US $11.90 137. Handbook for Automatic Computation. Vol. 1/Part b: Grau/Hill/Langmaack: Translation of Aiwa, 60. DM 64,-; US $23.70 138. Hahn: Stability of Motion. DM 72,-; US $26.70 139. Mathematische Hilfsmittel des Ingenieurs. 1. Teil. DM 88,-; US $32.60 140. Mathematische Hilfsmittel des Ingenieurs. 2. Tell. DM 136,-; US $50.40 141. Mathematische Hilfsmittel des Ingenieurs. 3. Tel DM 98,-; US $36.30 142. Mathematische Hilfsmittel des Ingenieurs. 4. Teil. DM 124, -; US $45.90 143. Schur/Grunsky: Vorlesungen iiber Invariantentheorie. DM 40,-; US $14.80 144. Weil: Basic Number Theory. DM 48,-; US $17.80 145. Butzer/Berens: Semi-Gronps of Operators and Approximation. DM 56,-; US $20.80 146. Treves: Locally Convex Spaces and Linear Partial Differential Equations. DM 36,-; US $13.40 147. Lamotke: Semisimpliziale algebraische Topologie. DM 56,-; US $20.80 148. Chandrasekharan: Introduction to Analytic Nnmber Theory. DM 28,-; US $10.40 149. Sario/Oikawa: Capacity Functions. DM 96,-; US $35.60 150. Iosifescn/Theodorescu: Random Processes and Learning. DM 68,-; US $25.20 151. Mandl: Analytical Treatment of One-dimensional Markov Processes. DM 36,-; US $13.40 152. Hewitt/Ross: Abstract Harmonic Analysis. Vol. 2: Structure and Analysis for Compact Groups. Analysis on Locally Compact Abelian Groups. DM 140,-; US $51.80 153. Federer: Geometric Measure Theory. DM 118,-; US $43.70 154. Singer: Bases in Banach Spaces I. DM 112,-; US $41.50 155. Willer: Foundations of the Mathematical Theory of Electromagnetic Waves. DM 58,-; US $21.50 156. van der Waerden: Mathematical Statistics. DM 68,-; US $25.20 157. Prohorov/Rozanov: Probability Theory. DM 68,- -; US $25.20 159. Ktithe: Topological Vector Spaces I. DM 78,-; US $28.90 160. Agrest/Maksimov: Theory of Incomplete Cylindrical Functions and Their Applications. DM 84,60; US $31.30 161, Bhatia/Szeg6: Stability Theory of Dynamical Systems. DM 58,-; US $21.50 162. Nevanlinna: Analytic Functions. DM 76,-; US $28.20