• Nie Znaleziono Wyników

Primitive minima of positive definite quadratic forms

N/A
N/A
Protected

Academic year: 2021

Share "Primitive minima of positive definite quadratic forms"

Copied!
6
0
0

Pełen tekst

(1)

LXIII.1 (1993)

Primitive minima of positive definite quadratic forms

by

Aloys Krieg (M¨unster)

1. Introduction. The main purpose of the reduction theory is to construct a fundamental domain of the unimodular group acting discontin- uously on the space of positive definite quadratic forms. This fundamental domain is for example used in the theory of automorphic forms for GLn

(cf. [11]) or in the theory of Siegel modular forms (cf. [1], [4]). There are several ways of reduction, which are usually based on various minima of the quadratic form, e.g. the Korkin–Zolotarev method (cf. [10], [3]), Venkov’s method (cf. [12]) or Vorono¨ı’s approach (cf. [13]), which also works in the general setting of positivity domains (cf. [5]). The most popular method is Minkowski’s reduction theory [6] and its generalizations (cf. [9], [15]).

Minkowski’s reduction theory is based on attaining certain minima, which can be characterized as the successive primitive minima of the quadra- tic form. Besides these we have successive minima, but a reduction accord- ing to successive minima only works for n ≤ 4 (cf. [14]). In this paper we introduce so-called primitive minima, which lie between successive and suc- cessive primitive minima (cf. Theorem 2). Using primitive minima we obtain a straightforward generalization of Hermite’s inequality in Theorem 1. As an application we get a simple proof for the finiteness of the class num- ber. Finally we describe relations with Rankin’s minima (cf. [8]) and with Venkov’s reduction (cf. [12]).

2. Various minima. Let Pn denote the set of all real positive definite n×n matrices. GLn(Z) stands for the unimodular group of degree n, i.e. the group of units in the ring Mn(Z). An integral n × k matrix P ∈ Mn×k(Z), n ≥ k, is called primitive, if the g.c.d. of all the k-rowed minors of P is 1.

This is equivalent to the fact that there exists a matrix (P, ∗) ∈ GLn(Z) (cf. [7]). Moreover, set

A[B] := (Bt)AB for matrices A, B of appropriate size.

(2)

A matrix S = (sjk) ∈ Pn is called Minkowski-reduced whenever (M.1) S[g] ≥ skk for all g = (γ1, . . . , γn)t ∈ Zn

with g.c.d. (γk, . . . , γn) = 1 , 1 ≤ k ≤ n , (M.2) sk,k+1≥ 0 for 0 < k < n .

The set of Minkowski-reduced matrices is a fundamental domain of Pnwith respect to the discontinuous group of mappings

Pn→ Pn, S 7→ S[U ] , U ∈ GLn(Z) .

In order to determine a unimodular matrix U such that S[U ] is Minkowski- reduced proceed as follows (cf. [4]): Given S ∈ Pn define its minimum by (1) µ(S) := inf{S[h] | 0 6= h ∈ Zn} .

Determine g1 ∈ Zn with µ(S) = S[g1]. As soon as g1, . . . , gk, 0 < k < n, are given, choose gk+1∈ Zn such that

(2) S[gk+1] = inf{S[h] | (g1, . . . , gk, h) ∈ Mn×(k+1)(Z) is primitive} . If necessary replace gk+1by −gk+1in order to get gkSgk+1 ≥ 0. In this way we construct a unimodular matrix U = (g1, . . . , gn) such that T = S[U ] is Minkowski-reduced. The diagonal entries of T are given by (1) and (2) and may therefore be called the successive primitive minima of S.

Besides these the successive minima µ1(S), . . . , µn(S) of S ∈ Pn were introduced (cf. [14]). Determine g1∈ Zn as in (1), i.e.

µ1(S) = µ(S) = S[g1] .

As soon as g1, . . . , gk, 0 < k < n, are given, choose gk+1∈ Zn such that (3) µk+1(S) = S[gk+1] = inf{S[h] | h ∈ Zn, rank(g1, . . . , gk, h) = k + 1} . Using Steinitz’ theorem we have the alternative definition

(4) µk(S) = inf

 t ∈ R

there is H = (h1, . . . , hk) ∈ Mn×k(Z) , rank H = k , S[hj] ≤ t , 1 ≤ j ≤ k

 , 1 ≤ k ≤ n . Comparing (3) and (4) it is interesting to investigate the analogue for prim- itive matrices in place of maximal rank matrices. We define

(5) νk(S) = inf

 t ∈ R

there is a primitive H = (h1, . . . , hk) in Mn×k(Z) , S[hj] ≤ t , 1 ≤ j ≤ k

 , 1 ≤ k ≤ n . We call νk(S) the k-th primitive minimum of S. Obviously one has (6) µk(S) ≤ νk(S) , 1 ≤ k ≤ n , ν1(S) = µ1(S) = µ(S) .

(3)

3. A generalization of Hermite’s inequality. For S ∈ Pn we have (7) µ(S) = ν1(S) ≤ ν2(S) ≤ . . . ≤ νn(S) .

Since U P, U ∈ GLn(Z), is primitive if and only if P is, we conclude (8) νk(S[U ]) = νk(S) for U ∈ GLn(Z) , 1 ≤ k ≤ n .

Note that a primitive matrix can be completed to a unimodular matrix.

Hence given 1 ≤ k ≤ n there exists Uk∈ GLn(Z) such that

(9) S[Uk] = T = (tij) , t11 ≤ t22≤ . . . ≤ tnn, tkk = νk(S) . Theorem 1. Given S ∈ Pn one has

ν1(S) . . . νn(S) ≤ (43)n(n−1)/2det S .

P r o o f. We use induction on n; the case n = 1 is obvious. According to (8) and (9) we may assume s11 = µ(S) = ν1(S) =: µ without restriction.

By the method of completing squares we obtain a decomposition S = µ 0

0 T

  1 at 0 I



= µ µat µa T + µaat



, T ∈ Pn−1, a ∈ Rn−1, where I is the (n − 1) × (n − 1) identity matrix. Given 0 < k < n there exists a primitive matrix G = (g1, . . . , gk) ∈ M(n−1)×k(Z) such that

T [gj] ≤ νk(T ) , 1 ≤ j ≤ k .

Next choose g = (γ1, . . . , γk)t ∈ Zk such that the entries of g + Gta belong to the interval [−12;12]. Now

H = 1 gt

0 G



∈ Mn×(k+1)(Z) and H0=gt G



∈ Mn×k(Z) are primitive. One has

S γj

gj



= µ(γj + atgj)2+ T [gj] ≤ 14ν1(S) + νk(T ) , 1 ≤ j ≤ k . Since H0 is primitive we conclude

νk(S) ≤ 14ν1(S) + νk(T ) . Now (7) leads to

S 1 0



= ν1(S) ≤ νk(S) ≤ 14ν1(S) + νk(T ) . Since H is primitive, we now have

νk+1(S) ≤ 14ν1(S) + νk(T ) and νk+1(S) ≤ 43νk(T ) . According to ν1(S) det T = det S the induction hypothesis yields

ν1(S) . . . νn(S) ≤ (43)n−1ν1(S)ν1(T ) . . . νn−1(T )

≤ (43)n(n−1)/2ν1(S) det T = (43)n(n−1)/2det S .

(4)

In view of (7) we obtain Hermite’s inequality (cf. [7]) as Corollary 1. Given S ∈ Pn one has

µ(S)n≤ (43)n(n−1)/2det S .

Denote the class number by hn(N ) , N ≥ 1, i.e. hn(N ) is the number of GLn(Z)-equivalence classes of integral S ∈ Pn with det S = N .

Corollary 2. The class numbers hn(N ) , N ≥ 1, are finite. One has hn(N ) = O(Nn(n+1)/2) as N → ∞ .

P r o o f. By (9) it suffices to count the number of integral S ∈ Pn with det S = N and skk ≤ νn(S) , 1 ≤ k ≤ n. In view of νk(S) ≥ 1 Theorem 1 implies

0 < skk ≤ νn(S) ≤ ν1(S) . . . νn(S) ≤ (43)n(n−1)/2N .

Next S ∈ Pn yields sjjskk − s2jk > 0, hence |sjk| < (43)n(n−1)/2N for 1 ≤ j < k ≤ n. Thus the number of these S is O(Nn(n+1)/2) as N → ∞.

For other proofs of Corollary 2 we refer to [7].

4. Relations with other types of minima. The first relation is derived in

Theorem 2. Let S = (sij) ∈ Pn be Minkowski-reduced. Given 1 ≤ k ≤ n one has

µk(S) ≤ νk(S) ≤ skk ≤ αkµk(S) ≤ αkνk(S) , where

αk = 1 if k ≤ 4 , (54)k−4 if k ≥ 4 .

P r o o f. νk(S) ≤ skk follows from s11 ≤ . . . ≤ snn. The remaining parts are consequences of (6) and [14], Satz 7 and (45).

If k ≥ 5 there are quadratic forms S with νk(S) > µk(S). Just as in [14]

consider the matrix S attached to the quadratic form

x21+ x22+ x23+ x24+ (x1+ x2+ x3+ x4)x5+54x25. One easily checks

µk(S) = νj(S) = 1 , 1 ≤ k ≤ 5 , 1 ≤ j ≤ 4 , ν5(S) = 54. Next consider the minima

δk(S) := inf{det(S[P ]) | P ∈ Mn×k(Z) primitive}

= inf{det(S[G]) | G ∈ Mn×k, rank G = k} , 1 ≤ k ≤ n , which were introduced by Rankin [8].

(5)

Proposition 1. Given S ∈ Pn and 1 ≤ k ≤ n one has ν1(S) . . . νk(S) ≤ (43)k(k−1)/2δk(S) .

P r o o f. Choose a primitive P ∈ Mn×k(Z) with δk(S) = det(S[P ]).

Apply Theorem 1 to S[P ]. In view of the obvious inequalities νj(S[P ]) ≥ νj(S) for 1 ≤ j ≤ k, the claim follows.

Given T ∈ Pk and S ∈ Pn, 1 ≤ k ≤ n, we define

νT(S) := inf{tr(S[P ]T ) | P ∈ Mn×k(Z) primitive} , where tr is the trace. Clearly the minimum is attained and one has

νI(S) ≥ ν1(S) + . . . + νk(S) , I =

1 0

. ..

0 1

∈ Pk,

where equality holds at least for k ≤ 4. If k = n and T ∈ Pn has no non-trivial automorphs, then Venkov [12] showed that

{S ∈ Pn | tr(ST ) = νT(S)}

is a fundamental domain of Pn with respect to the action of the unimodular group.

Proposition 2. Let S ∈ Pn, T ∈ Pk, 1 ≤ k ≤ n. Then one has νT(S) ≥ kδk(S)1/k(det T )1/k ≥ k(34)(k−1)/2µ(S)µ(T ) .

P r o o f. Choose a primitive P ∈ Mn×k(Z) with νT(S) = tr(S[P ]T ).

Then apply the result of Barnes and Cohn [2] to S[P ] and T : νT(S) = tr(S[P ]T ) ≥ k(det(S[P ]))1/k(det T )1/k.

One has det(S[P ]) ≥ δk(S). Now the claim follows by virtue of Proposi- tion 1, Corollary 1 and (7).

References

[1] A. N. A n d r i a n o v, Quadratic Forms and Hecke Operators, Grundlehren Math.

Wiss. 286, Springer, Berlin 1987.

[2] E. S. B a r n e s and M. J. C o h n, On the inner product of positive quadratic forms, J. London Math. Soc. (2) 12 (1975), 32–36.

[3] D. G r e n i e r, Fundamental domains for the general linear group, Pacific J. Math.

132 (1988), 293–317.

[4] H. K l i n g e n, Introductory Lectures on Siegel Modular Forms, Cambridge University Press, Cambridge 1990.

[5] M. K o e c h e r, Beitr¨age zu einer Reduktionstheorie in Positivit¨atsbereichen I , Math.

Ann. 141 (1960), 384–432.

[6] H. M i n k o w s k i, Diskontinuit¨atsbereich f¨ur arithmetische ¨Aquivalenz , J. Reine Angew. Math. 129 (1905), 220–274.

(6)

[7] M. N e w m a n, Integral Matrices, Academic Press, New York 1972.

[8] R. A. R a n k i n, On positive definite quadratic forms, J. London Math. Soc. 28 (1953), 309–319.

[9] S. S. R y s h k o v, On the Hermite–Minkowski reduction theory for positive quadratic forms, J. Soviet Math. 6 (1976), 651–671.

[10] S. S. R y s h k o v and E. P. B a r a n o v s k i˘ı, Classical methods in the theory of lattice packings, Russian Math. Surveys 34 (4) (1979), 1–68.

[11] A. T e r r a s, Harmonic Analysis on Symmetric Spaces and Applications II , Springer, New York 1988.

[12] A. B. V e n k o v, On the reduction of positive quadratic forms, Izv. Akad. Nauk SSSR Ser. Mat. 4 (1940), 37–52 (in Russian).

[13] G. V o r o n o¨ı, Sur quelques propri´et´es des formes quadratiques positives parfaites, J. Reine Angew. Math. 133 (1907), 97–178.

[14] B. L. van der W a e r d e n, Die Reduktionstheorie der positiven quadratischen For- men, Acta Math. 96 (1956), 263–309.

[15] H. W e y l, Theory of reduction for arithmetical equivalence, Trans. Amer. Math.

Soc. 48 (1940), 126–164.

MATHEMATISCHES INSTITUT

WESTF ¨ALISCHE WILHELMS–UNIVERSIT ¨AT EINSTEINSTR. 62

W-4400 M ¨UNSTER

FEDERAL REPUBLIC OF GERMANY E-mail: ALOYS@MATH.UNI-MUENSTER.DE

Received on 30.4.1992 (2257)

Cytaty

Powiązane dokumenty

These bases allow us to prove Theorem 6.13 that states that the linear action on the space of algebraic restrictions of closed 2-forms to the germ of a quasi-homogeneous

It will be shown that for positive 2D linear systems a linear form of the state vec- tor can be chosen as a Lyapunov function and there exists a strictly positive diagonal matrix P

Kr¨ uskemper gave a local global principle for scaled trace forms of odd dimension over an algebraic number field (see [16], Theorem 1).. We give a stronger version of

However, it is necessary that P(x) and Q(x) assume simultaneously arbitrarily large positive values prime to any fixed natural number.. This condition will be

The aim of the present paper is to prove the Nullstellensatz (Theorem 1) and the Darstellungssatz (Theorem 4) in the remaining open case where K is an algebraic function field in

- understand the general purpose, role and benefits of insurance, - demonstrate a knowledge of how the insurance market works, - understand how the general insurance market

© 2013 Académie des sciences. Published by Elsevier Masson SAS. Nous prouvons également que ceci n’est plus vrai pour les formes de degré supérieur... © 2013 Académie des

This abstract result provides an elementary proof of the existence of bifurcation intervals for some eigenvalue problems with nondifferentiable nonlinearities1. All the results