ebook img

A generalization of the Hasse-Witt matrix of a hypersurface PDF

0.19 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview A generalization of the Hasse-Witt matrix of a hypersurface

A GENERALIZATION OF THE HASSE-WITT MATRIX OF A HYPERSURFACE 7 1 ALAN ADOLPHSON AND STEVEN SPERBER 0 2 n Abstract. The Hasse-Witt matrix of a hypersurface in Pn over a a finite field of characteristic p gives essentially complete mod p J information about the zeta function of the hypersurface. But if 7 the degreed of the hypersurfaceis ≤n, the zeta function is trivial 1 mod p and the Hasse-Witt matrix is zero-by-zero. We generalize ] a classical formula for the Hasse-Witt matrix to obtain a matrix G that gives a nontrivial congruence for the zeta function for all d. A We also describe the differential equations satisfied by this matrix . and prove that it is generically invertible. h t a m [ 1. Introduction 1 Let p be a prime number, let q = pa, and let F be the field of q v q 9 elements. Let 0 N 5 a (1.1) f (x) = λ x j ∈ F [x ,...,x ] 4 λ j q 0 n 0 j=1 . X 1 be a homogeneous polynomial of degree d. We write a = (a ,...,a ) 0 j 0j nj 7 with ni=0aij = d. Let Xλ ⊆ Pn be the hypersurface defined by the 1 equation f (x) = 0 and let Z(X /F ,t) be its zeta function. We write λ λ q : P v P (t)(−1)n Xi (1.2) Z(Xλ/Fq,t) = λ (1−t)(1−qt)···(1−qn−1t) r a for some rational function P (t) ∈ 1+tZ[[t]]. If d = 1 then P (t) = 1, λ λ so we shall always assume that d ≥ 2. Define a nonnegative integer µ by the equation n+1 = µ+1, d (cid:24) (cid:25) where ⌈r⌉ denotes the least integer greater than or equal to the real number r. By a result of Ax[4] (see also Katz[5, Proposition 2.4]) we have P (q−µt) ∈ 1+tZ[[t]]. λ Date: January 18, 2017. 1 2 ALAN ADOLPHSONANDSTEVEN SPERBER Our goal in this paper is to give a mod p congruence for P (q−µt). We λ do this by defining a generalization of the classical Hasse-Witt matrix, which gives such a congruence for µ = 0. Presumably our matrix is the matrix of a “higher Hasse-Witt” operation as defined by Katz[6, Section 2.3.4], but so far we have not been able to prove this. It will be convenient to define an augmentation of the vectors a . j Set a+ = (a ,...,a ,1) ∈ Nn+2, j = 1,...,N, j 0j nj where N denotes the nonnegative integers. Note that the vectors a+ all j lie on the hyperplane n u = du in Rn+2. We shall be interested i=0 i n+1 in the lattice points on this hyperplane that lie in (R )n+2: set >0 P n U = u = (u ,...,u ) ∈ Nn+2 | u = du and u > 0 for all i . 0 n+1 i n+1 i (cid:26) i=0 (cid:27) X Note that u ∈ U implies that u ≥ µ+1. Let n+1 U = {u = (u ,...,u ) ∈ U | u = µ+1}, min 0 n+1 n+1 a nonempty set by the definition of µ. We define a matrix of polynomi- alswithrowsandcolumnsindexedbyU : letA(Λ) = [A (Λ)] , min uv u,v∈Umin where Λν1···ΛνN A (Λ) = (−1)µ+1 1 N ∈ Q[Λ ,...,Λ ]. uv 1 N ν !···ν ! 1 N νX∈NN PNj=1νja+j =pu−v Note that since the (n + 1)-st coordinate of each a+ equals 1, the j condition on the summation implies that N ν = (p−1)(µ+1). j j=1 X When µ = 0, it follows that ν ≤ p−1 for all j, hence the matrix A(Λ) j can be reduced modulo p. We denote by A¯(Λ) ∈ F [Λ] its reduction p modulo p. Using thealgorithmofKatz[6, Algorithm2.3.7.14], onethen checks that A¯(λ) is the Hasse-Witt matrix of the hypersurface f = 0. λ It is somewhat surprising that even when µ > 0 we still have ν ≤ p−1 j for all j. Lemma 1.3. If u,v ∈ U , ν ∈ NN, and N ν a+ = pu −v, then min j=1 j j ν ≤ p − 1 for all j. In particular, A (Λ) ∈ (Q ∩ Z )[Λ], so A (Λ) j uv p uv P can be reduced modulo p. A GENERALIZATION OF THE HASSE-WITT MATRIX 3 The proof of Lemma 1.3 will be given in Section 2. By the results of [1, Theorem 2.7 or Theorem 3.1], which will be recalled in Section 2, Lemma 1.3 implies immediately that each A (Λ) is a mod p solution uv of an A-hypergeometric system of differential equations. Write the rational function P (t) of (1.2) as λ Q (t) λ P (t) = , λ R (t) λ where Q (t) and R (t) are relatively prime polynomials with λ λ Q (q−µt), R (q−µt) ∈ 1+tZ[t]. λ λ If X is smooth, it is known that P (t) is a polynomial, i. e., R (t) = 1. λ λ λ Ourmainresultisthefollowing, whichdoesnotrequireanysmoothness assumption. Theorem 1.4. If n is not divisible by d, then R (q−µt) ≡ 1 (mod q) λ and Q (q−µt) ≡ det I −tA¯(λpa−1)A¯(λpa−2)···A¯(λ) (mod p). λ Notethateven inth(cid:0)eclassical caseoftheHasse-Wi(cid:1)tt matrix(µ = 0), this result contains something new, as we do not assume that X is a λ smooth hypersurface. The proof of Theorem 1.4 will occupy Sections 3–5. To describe the zeta function, we apply the p-adic cohomology theory of Dwork, as in Katz[7, Sections 4–6]. Indeed, Equation (3.5) below is a refined version of [7, Equation (4.5.33)]. We discuss the case d|n in Section 6. If d|n, the conclusion of Theorem 1.4 need not hold, and the rational function P (q−µt) (mod p) is instead described by Theorem 6.2. We prove the λ generic invertibility of the matrix A¯(Λ) in Section 7. 2. Proof of Lemma 1.3 It will be convenient for later applications to prove a more general version of Lemma 1.3. Put S = {0,1,...,n} and let I ⊆ S. Define an integer µ by the equation I |I| = µ +1. I d (cid:24) (cid:25) Note that µ ≥ 0 if I 6= ∅, µ = −1, and, in the notation of the I ∅ Introduction, µ = µ. Set S n UI = u = (u ,...,u ) ∈ Nn+2 | u = du and u > 0 for all i ∈ I . 0 n+1 i n+1 i (cid:26) i=0 (cid:27) X 4 ALAN ADOLPHSONANDSTEVEN SPERBER Note that u ∈ UI implies that u ≥ µ +1. Let n+1 I UI = {u = (u ,...,u ) ∈ UI | u = µ +1}, min 0 n+1 n+1 I a nonempty set by the definition of µ . Lemma 1.3 is the special case I I = S of the following result. Lemma 2.1. If u,v ∈ UI , ν ∈ NN, and N ν a+ = pu −v, then min j=1 j j ν ≤ p−1 for all j. j P Proof. The result is trivial when I = ∅ since U∅ = {(0,...,0)}, so as- min sume I 6= ∅. Let u = (u ,...,u ,µ +1),v = (v ,...,v ,µ +1) ∈ UI . 0 n I 0 n I min Fixk ∈ {1,...,N}. Weclaimthereexistsanindexi ∈ {0,...,n}such 0 that u if i ∈ I, (2.2) a ≥ i0 0 i0k u +1 if i 6∈ I. ( i0 0 For if (2.2) fails for all i ∈ {0,...,n}, then 0 u−a+ = (u −a ,...,u −a ,µ ) ∈ UI, k 0 0k n nk I contradicting the definition of µ . I If ν ≥ p, then k pu if i ∈ I, ν a ≥ pa ≥ i0 0 k i0k i0k pu +p if i 6∈ I, ( i0 0 hence in both cases we have ν a > pu −v . k i0k i0 i0 But our hypothesis N ν a+ = pu−v implies that j=1 j j P νkai0k ≤ pui0 −vi0. This contradiction shows that ν ≤ p−1. And since k was arbitrary, k (cid:3) the lemma is established. We recall the definition of the A-hypergeometric system of differen- tial equations associated to the set A = {a+}N . Let L ⊆ ZN be the j j=1 lattice of relations on A, N L = l = (l ,...,l ) ∈ ZN | l a+ = 0 , 1 N j j (cid:26) j=1 (cid:27) X and let β = (β ,...,β ) ∈ Cn+2. The A-hypergeometric system with 0 n+1 parameter β is the system of partial differential operators in variables A GENERALIZATION OF THE HASSE-WITT MATRIX 5 Λ ,...,Λ consisting of the box operators 1 N ∂ lj ∂ −lj (cid:3) = − for l ∈ L l ∂Λ ∂Λ lYj>0(cid:18) j(cid:19) lYj<0(cid:18) j(cid:19) and the Euler (or homogeneity) operators N ∂ Z = a Λ −β for i = 0,...,n i ij j i ∂Λ j j=1 X and N ∂ Z = Λ −β . n+1 j n+1 ∂Λ j j=1 X Let AI(Λ) = [AI (Λ)] , where uv u,v∈UI min Λν1···ΛνN (2.3) AI (Λ) = (−1)µI+1 1 N ∈ Q[Λ ,...,Λ ]. uv ν !···ν ! 1 N 1 N νX∈NN PNj=1νja+j=pu−v NotethatinthenotationoftheIntroductionwehaveAS (Λ) = A (Λ). uv uv By Lemma 2.1, the polynomials AI (Λ) have p-integral coefficients. uv Lemma 2.1 also says that pu−v is very good in the sense of [1, Sec- tion 2]. We may therefore apply [1, Theorem 2.7] (or [1, Theorem 3.1] since this system is nonconfluent) to conclude that A¯I (Λ) is a mod p uv solution of the A-hypergeometric system with parameter β = pu − v (or, equivalently, β = −v since we have reduced modulo p). 3. The zeta function To make a connection between the matrix A(Λ) and the zeta func- tion (1.2), we apply a consequence of the Dwork trace formula devel- oped in [3] (see Equation (3.5) below). Let γ be a zero of the series 0 ∞ tpi/pi having ord γ = 1/(p−1), where ord is the p-adic valua- i=0 p 0 p tion normalized by ord p = 1. Let L be the space of series p 0 P n L = c γpun+1xu | u −du = 0, c ∈ C , and {c } is bounded . 0 u 0 i n+1 u p u (cid:26)u∈Nn+2 i=0 (cid:27) X X For I ⊆ {0,...,n}, let LI be the subset of L defined by 0 0 LI = c γpun+1xu ∈ L | u > 0 for i ∈ I . 0 u 0 0 i (cid:26)u∈Nn+2 (cid:27) X 6 ALAN ADOLPHSONANDSTEVEN SPERBER Let AH(t) = exp( ∞ tpi/pi) be the Artin-Hasse series, a power series i=0 in t that has p-integral coefficients, and set P ∞ θ(t) = AH(γ t) = θ ti. 0 i i=0 X We then have i (3.1) ord θ ≥ . p i p−1 We define the Frobenius operator on L . Put 0 N ˆ ˆ a+ (3.2) θ(λ,x) = θ(λjx j ), j=1 Y ˆ where λ denotes the Teichmu¨ller lifting of λ. We shall also need to ˆ consider the series θ (λ,x) defined by 0 a−1 N a−1 (3.3) θ0(λˆ,x) = θ (λˆjxa+j )pi = θ(λˆpi,xpi). i=0 j=1 i=0 YY (cid:0) (cid:1) Y Define an operator ψ on formal power series by (3.4) ψ c xu = c xu. u pu (cid:18)u∈Nn+2 (cid:19) u∈Nn+2 X X Denote by α the composition λˆ α := ψa ◦“multiplication by θ (λˆ,x).” λˆ 0 The map α operates on L and is stable on each LI. The proof of λˆ 0 0 Theorem 1.4 will be based on the following formula for the rational function P (t) defined in (1.2). By [3, Equation 7.12] we have λ (3.5) P (qt) = det(I −qn+1−|I|tα | LI)(−1)n+1+|I|. λ λˆ 0 I⊆{0,1,...,n} Y To exploit (3.5) we shall need p-adic estimates for the action of α λˆ on LI. Expand (3.3) as a series in x, say, 0 (3.6) θ (λˆ,x) = θ (λˆ)xw. 0 0,w w∈NA X Note that from the definitions we have θ (λˆ) ∈ Q (ζ ,γ ). A direct 0,w p q−1 0 calculation shows that for v ∈ UI, (3.7) α (xv) = θ (λˆ)xu, λˆ 0,qu−v u∈UI X A GENERALIZATION OF THE HASSE-WITT MATRIX 7 thus we need p-adic estimates for the θ (λˆ) with u,v ∈ UI. 0,qu−v Expand (3.2) as a series in x: (3.8) θ(λˆ,x) = θ (λˆ)xw, w w∈NA X where (3.9) θ (λˆ) = θ(w)λˆν w ν ν∈NN X with N θ if N ν a+ = w, (3.10) θ(w) = j=1 νj j=1 j j ν 0 if N ν a+ 6= w. (Q Pj=1 j j From (3.1) we have the estimate P N (3.11) ord θ(w) ≥ j=1νj = wn+1. p ν p−1 p−1 P In particular, this implies the estimate w ˆ n+1 (3.12) ord θ (λ) ≥ . p w p−1 By (3.3) and (3.8) we have a−1 (3.13) θ (λˆ) = θ (λˆpi). 0,w u(i) u(0),...,Xu(a−1)∈NAYi=0 Pa−1piu(i)=w i=0 In particular, we get the formula a−1 (3.14) θ (λˆ) = θ (λˆpi). 0,qu−v w(i) w(0),...,Xw(a−1)∈NAYi=0 Pa−1piw(i)=qu−v i=0 Applying (3.12) to the products on the right-hand side of (3.14) gives a−1 a−1 (i) w (3.15) ord θ (λˆpi) ≥ n+1. p w(i) p−1 (cid:18)i=1 (cid:19) i=0 Y X ˆ This estimate is not directly helpful for estimating θ (λ) because 0,qu−v we lack information about the w(i). Instead we proceed as follows. Fix w(0),...,w(a−1) ∈ NA with a−1 (3.16) piw(i) = qu−v. i=0 X 8 ALAN ADOLPHSONANDSTEVEN SPERBER We construct inductively from {w(i)}a−1 a related sequence {w˜(i)}a ⊆ i=0 i=0 UI such that (3.17) w(i) = pw˜(i+1) −w˜(i) for i = 0,...,a−1. First of all, take w˜(0) = v. Eq. (3.16) shows that w(0) + w˜(0) = pw˜(1) for some w˜(1) ∈ Zn+2; since w(0) ∈ NA and w˜(0) ∈ UI we conclude that w˜(1) ∈ UI. Suppose that for some 0 < k ≤ a−1 we have defined w˜(0),...,w˜(k) ∈ UI satisfying (3.17) for i = 0,...,k −1. Substituting pw˜(i+1) −w˜(i) for w(i) for i = 0,...,k −1 in (3.16) gives a−1 (3.18) −w˜(0) +pkw˜(k) + piw(i) = pau−v. i=k X Since w˜(0) = v, we can divide this equation by pk to get w˜(k) +w(k) = pw˜(k+1) for some w˜(k+1) ∈ Zn+2. Since w(k) ∈ NA and (by induction) w˜(k) ∈ UI, we conclude that w˜(k+1) ∈ UI. This completes the inductive construction. Note that in the special case k = a−1, this computation gives w˜(a) = u. Summing Eq. (3.17) over i = 0,...,a−1 and using w˜(0) = v, w˜(a) = u, gives a−1 a−1 (3.19) w(i) = pu−v +(p−1) w˜(i), i=0 i=1 X X hence a−1 (i) a−1 w pu −v (3.20) n+1 = n+1 n+1 + w˜(i) . p−1 p−1 n+1 i=0 i=1 X X For w(0),...,w(a−1) as in (3.16), we thus get from (3.15) a−1 a−1 pu −v (3.21) ord θ (λˆpi) ≥ n+1 n+1 + w˜(i) . p w(i) p−1 n+1 (cid:18)i=0 (cid:19) i=1 Y X Since w˜(i) ∈ UI, we have (3.22) w˜(i) = µ +1 if w˜(i) ∈ UI and w˜(i) ≥ µ +2 if w˜(i) 6∈ UI . n+1 I min n+1 I min From (3.21) and (3.22) we get the following result. Lemma 3.23. For u,v ∈ UI and w(0),...,w(a−1) as in (3.16), we have a−1 pu −v (3.24) ord θ (λˆpi) ≥ n+1 n+1 +(a−1)(µ +1). p w(i) p−1 I (cid:18)i=0 (cid:19) Y A GENERALIZATION OF THE HASSE-WITT MATRIX 9 Furthermore, if any of the terms w˜(1),...,w˜(a−1) of the associated se- quence satisfying (3.17) is not contained in UI , then min a−1 pu −v (3.25) ord θ (λˆpi) ≥ n+1 n+1 +(a−1)(µ +1)+1. p w(i) p−1 I (cid:18)i=0 (cid:19) Y ˆ Our desired estimate for θ (λ) now follows from (3.14). 0,qu−v Corollary 3.26. For u,v ∈ UI we have pu −v ˆ n+1 n+1 (3.27) ord θ (λ) ≥ +(a−1)(µ +1). p 0,qu−v I p−1 4. The action of α on LI λˆ 0 In this section, we use Corollary 3.26 to study the action of α on LI. λˆ 0 From (3.7) and the formula of Serre[8, Proposition 7] we have ∞ (4.1) det(I −tα | LI) = aI tm, λˆ 0 m m=0 X where (4.2) aI = (−1)m sgn(σ) θ (λˆ) ∈ Q (ζ ,γ ), m 0,qu−σ(u) p q−1 0 UmX⊆UIσX∈Sm uY∈Um the outer sum is over all subsets U ⊆ UI of cardinality m, and S is m m the group of permutations on m objects. Proposition 4.3. The coefficient aI is divisible by qm(µI+1) and sat- m isfies the congruence (4.4) aI ≡ (−1)m sgn(σ) θ (λˆ) (mod pqm(µI+1)). m 0,qu−σ(u) UmX⊆UmIinσX∈Sm uY∈Um In particular, aI ≡ 0 (mod pqm(µI+1)) if m > |UI |. m min Proof. If U ⊆ UI is a subset of cardinality m and σ is a permutation m of U , then by (3.27) m pu −σ(u) ˆ n+1 n+1 ord θ (λ) ≥ m(a−1)(µ +1)+ p 0,qu−σ(u) I p−1 (cid:18)uY∈Um (cid:19) uX∈Um = m(a−1)(µ +1)+ u I n+1 uX∈Um ≥ ma(µ +1) I 10 ALAN ADOLPHSONANDSTEVEN SPERBER since u ∈ U implies u ≥ µ + 1. It follows from (4.2) that aI is m n+1 I m divisible by qm(µI+1). Furthermore, u ≥ µ +2 if u 6∈ UI , so n+1 I min ord θ (λˆ) ≥ ma(µ +1)+1 if U 6⊆ UI . p 0,qu−σ(u) I m min (cid:18)uY∈Um (cid:19) (cid:3) The congruence (4.4) now follows from (4.2). As an immediate corollary of Proposition 4.3, we have the following result. Corollary 4.5. The reciprocal roots of det(I−tα | LI) are all divisible λˆ 0 by qµI+1. Corollary 4.5 allows us to analyze the terms on the right-hand side of (3.5). Proposition 4.6. The reciprocal roots of det(I −qn+1−|I|tα | LI) are λˆ 0 divisible by qµ+2 unless either |I| = n+1 or |I| = n and n is divisible by d, in which case they are divisible by qµ+1. Proof. Corollary 4.5 and the definition of µ imply that the reciprocal I roots of det(I −qn+1−|I|tα | LI) are divisible by q to the power λˆ 0 |I| (4.7) n+1−|I|+ . d (cid:24) (cid:25) If |I| = n + 1, this reduces to µ + 1. Suppose |I| = n. From the definition of µ we have n = µd+r with 0 ≤ r ≤ d−1. The expression (4.7) then reduces to 1+⌈(µd+r)/d⌉, which equals µ+2 if r > 0 and equals µ+1 if r = 0. If |I| = n−1, expression (4.7) reduces to 2+⌈(µd+r−1)/d⌉, which equals µ + 3 if r > 1 and equals µ + 2 if r = 0,1. Finally, note that expression (4.7) cannot decrease when |I| decreases, so expression (4.7) will be ≥ µ+2 for |I| < n−1. (cid:3) From (3.5) and Proposition 4.6 we get the following result. Proposition 4.8. If n is not divisible by d, then (4.9) P (q−µt) ≡ det(I −q−µ−1tα | LS) (mod q). λ λˆ 0 If n is divisible by d, then det(I −q−µ−1tα | LS) (4.10) P (q−µt) ≡ λˆ 0 (mod q). λ n det(I −q−µtα | LS\{i}) i=0 λˆ 0 Q

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.