Sorry, your browser cannot access this site
This page requires browser support (enable) JavaScript
Learn more >

宝宝巴士第三集. 把之前的内容重新整合了一下, 配上了Weil对他的猜想在curve 情况下的证明懒得改之前的typo 了:(

Overview

In Weil4, Andre Weil made his remarkable conjecture about the Zeta function over finite field. For any projective variety XX over finite field Fq\mathbb{F}_{q}, we denote the set of Fqm\mathbb{F}_{q^{m}} points by X(Fqm)X(\mathbb{F}_{q^{m}}).

Theorem 1 (Weil-Grothendieck-Deligne). Let Nm=X(Fqm)N_{m}=\left| X(\mathbb{F}_{q^{m}}) \right| and ZX(t)=exp(m1Nmmtm)Z_{X}(t)=\exp(\sum\limits_{m\geq 1}\frac{N_{m}}{m}t^{m}) be the Zeta function of varitey XX. Assume dimX=n\dim X=n.

ZX(t)Z_{X}(t) is a rational function.

Let E=ΔΔE=\Delta\cdot \Delta be the self-intersection number of diagonal. Then ZX(t)Z_{X}(t) satisfies the functional equation

ZX(1qnt)=±qnE/2tEZX(t).Z_{X}(\frac{1}{q^{n}t})=\pm q^{nE /2}t^{E}Z_{X}(t).

We can write

ZX(t)=P1(t)P2n1(t)P0(t)P2n(t)Z_{X}(t)=\frac{P_{1}(t)\cdots P_{2n-1}(t)}{P_{0}(t)\cdots P_{2n}(t)}

where P0(t)=1tP_{0}(t)=1-t, P2n(t)=1qntP_{2n}(t)=1-q^{n}t and for 1i2n11\leq i\leq 2n-1, Pi(t)P_{i}(t) is a polynomial with integer coefficients, with roots αij\alpha_{ij} such that αij=qi/2\left| \alpha_{ij} \right| =q^{i /2}.

Define the ii-th betti number Bi(X)=degPi(t)B_{i}(X)=\deg P_{i}(t). Then E=i(1)iBiE=\sum\limits_{i}(-1)^{i}B_{i}.

The study on Weil conjecture advanced the integration of modern algebraic geometry and number theory. In the 1960s Grothendieck and Dwork used etale cohomology and p-adic analysis proved the rationality part of Weil conjecture. Grothendieck’s work also proved the functional equation part and pointed out the connection between Weil conjecture and cohomology theory. Deligne roved the Riemann Conjecture part in 1971 and was awarded the Fields Prize for his work. In this expository ariticle, we will follow Weil’s idea to prove Theorem 1 in smooth projective curve case.

Degree and Frobenius morphism

The idea of Deligne’s work is reduce Theorem 1 to some cohomology problems, and those NmN_{m} are number of fixed points under some automorphism of XX. First we recall some facts in intersection theory, a quick reference for this is Ha2 Appendix 3. From now on we assume XX is a nonsingular curve over some field kk and we first work on X×XX\times X. Denote the intersection product of divisor CC and DD by CDC\cdot D.

Definition 2. Let L\mathcal{L} be a line sheaf on XX. The Hilbert polynomial χ(L(n))\chi(\mathcal{L}(n)) is a quadratic function α2(L)2n2+α1(L)n+α0(L)\frac{\alpha_{2}(\mathcal{L})}{2}n^{2}+\alpha_{1}(\mathcal{L})n+\alpha_{0}(\mathcal{L}). The degree of L\mathcal{L} is defined to be α1(L)α1(OX)\alpha_{1}(\mathcal{L})-\alpha_{1}(\mathcal{O}_{X}).

If DD is a divisor of XX, by Hirzebruch-Riemann-Roch theorem, DH=deg(OX(D))D\cdot H= \deg(\mathcal{O}_{X}(D)), we define the number be the degree of the divisor DD. The degree of DD is a additive function on the divisor group Div(X)\mathrm{Div}(X) since the first chern class c1c_{1} is additive on tensor products.

Definition 3. Let f:XYf:X\to Y be a proper morphism of projective varieties. Define the degree of ff be the number [K(X),K(f(X))][K(X),K(f(X))]. If dimY<dimX\dim Y<\dim X, define the degree be 0.

Using the projection formula in Chow ring f(xfy)=fxyf_{*}(x\cdot f^{*}y)=f_{*}x\cdot y, we can show fCfD=deg(f)CDf^{*}C\cdot f^{*}D=\deg(f)C\cdot D. For any line bundle L\mathcal{L}, deg(fL)=deg(f)deg(L)\deg (f^{*}\mathcal{L})=\deg (f)\deg (\mathcal{L}). Note the degree is always a finite number.

We shall also recall some properties of geometric points on a variety XX. For any point xXx\in X, denote the residue field at xx by k(x)k(x). The set of X(Fqm)X(\mathbb{F}_{q^{m}}) points is exactly xXHom(k(x),Fqm)\coprod\limits_{x\in X}\mathop{\mathrm{Hom}}(k(x), \mathbb{F}_{q^{m}}).

Proposition 4. Let XX, mm defined as above, then

X(Fqm)=dmd{x closed point [k(x):Fq]=d}\left| X(\mathbb{F}_{q^{m}}) \right|=\sum_{d\mid m} d\cdot \left| \{x \text{ closed point }| [k(x):\mathbb{F}_{q}]=d\} \right|

Proof

Proof. First any image of geometric point Spec(Fqm)X\mathop{\mathrm{Spec}}(\mathbb{F}_{q^{m}})\to X is closed since the extension is algebraic. We act Gal(Fqm/Fq)\mathrm{Gal}(\mathbb{F}_{q^{m}}/\mathbb{F_{q}}) on Hom(k(x),Fqm)\mathop{\mathrm{Hom}}(k(x),\mathbb{F}_{q^{m}}) by composition on the left. Clearly the action is transitive and has stablizer Gal(k(x)/Fq)\mathrm{Gal}(k(x)/\mathbb{F}_{q}). So Hom(k(x),Fqm)=[k(x):Fq]\left| \mathop{\mathrm{Hom}}(k(x),\mathbb{F}_{q^{m}})\right|=[k(x):\mathbb{F}_{q}]. ◻

Define the number {x closed point [k(x):Fq]=d}\left| \{x \text{ closed point }| [k(x):\mathbb{F}_{q}]=d\} \right| to be the degree of closed point xx. For curve XX, deg(x)\deg(x) is the number of points in Weil divisor σGal(Fqm/Fq)σ(x)\sum\limits_{\sigma\in \mathrm{Gal}(\mathbb{F}_{q^{m}}/\mathbb{F_{q}})}\sigma(x), i.e. the number of points in the divisor which is stable under the Galois action.

Consider the absolute Frobenius φ:XX\varphi:X\to X which is identity map on the underlying topological space and ssqs\mapsto s^{q} on the structure sheaf. φ\varphi is a morphism of Fq\mathbb{F}_{q} schemes. It can be represented as an action of Galois group Gal(Fqˉ/Fq)\mathrm{Gal}(\bar{\mathbb{F}_{q}}/\mathbb{F}_{q}):

Definition 5. The Galois group Gal(Fqˉ/Fq)\mathrm{Gal}(\bar{\mathbb{F}_{q}} /\mathbb{F}_{q}) acts on XX as follows: For each σGal(Fqˉ/Fq)\sigma\in \mathrm{Gal}(\bar{\mathbb{F}_{q}} /\mathbb{F}_{q}), φˉ:XFqˉXFqˉ\bar{\varphi}:X_{\bar{\mathbb{F}_{q}}}\to X_{\bar{\mathbb{F}_{q}}} is the fibre product σ×idX\sigma\times id_{X}.

X¹FqX¹FqSpec¹FqSpec¹Fq'¾

The discussion above also shows that for nmn\mid m, the stablizer of X(Fqm)X(\mathbb{F}_{q^{m}}) under the Galois group action Gal(Fqm/Fqn)\mathrm{Gal}(\mathbb{F}_{q^{m}}/\mathbb{F}_{q^{n}}) is X(Fqn)X(\mathbb{F}_{q^{n}}).

Example 6. If we consider an affine variety YY which admits a closed immersion YAFqnY\to \mathbb{A}^{n}_{\mathbb{F}_{q}}. Then the Frobenius action on YFqˉY_{\bar{\mathbb{F}_{q}}} can be derived from the Frobenius action on AFqˉn\mathbb{A}^{n}_{\bar{\mathbb{F}_{q}}}. In particular, the Frobenius action on AFqˉn\mathbb{A}^{n}_{\bar{\mathbb{F}_{q}}} has the corresponding Fqˉ\bar{\mathbb{F}_{q}} algebra morphism Fqˉ[x1,,xn]Fqˉ[x1,,xn]\bar{\mathbb{F}_{q}}[x_{1},\dots, x_{n}]\to \bar{\mathbb{F}_{q}}[x_{1},\dots, x_{n}] given by aapa\to a^{p} on the coefficients. Therefore, the Fqˉ\bar{\mathbb{F}_{q}} points on YY has the morphism given by (a1,,an)(a1q,,anq)(a_{1},\dots, a_{n})\mapsto (a_{1}^{q},\dots, a_{n}^{q}).

Lemma 7. Let XX is reduced kk-scheme and YY is separated kk-scheme. f:XYf:X\to Y is a morphism. Let Γf(X)\Gamma_{f}(X) be the closed subscheme with reduced subscheme structure. Then Γf(X)X\Gamma_{f}(X)\cong X.

Proof

Proof. Note that Γf\Gamma_{f} is obtained by pulling back the diagonal:

XX£YYY£Y¡fff£id¢

Δ\Delta is closed immersion so Γf\Gamma_{f} is a closed immersion, hence Γf\Gamma_{f} is an isomorphism to the schematic image Γf(X)\overline{\Gamma_{f}(X)}. Since XX is reduced, we may factor Γf\Gamma_{f} through Γf(X)\Gamma_{f}(X), which induces an morphism from Γf(X)\overline{\Gamma_{f}(X)} to Γf(X)\Gamma_{f}(X). From the definition of reduced subscheme, Γf(X)Γf(X)\overline{\Gamma_{f}(X)}\cong \Gamma_{f}(X). ◻

Theorem 8. Let Γφ\Gamma_{\varphi} and δ\delta be the graph of φˉ\bar{\varphi} and idid in XFqˉ×XFqˉX_{\bar{\mathbb{F}_{q}}}\times X_{\bar{\mathbb{F}_{q}}}. The intersection multiplicity ΓφΔ=Nm\Gamma_{\varphi}\cdot \Delta= N_{m}.

Proof

Proof. The fixed point of XX under Galois group action Gal(Fqˉ/Fqm)\mathrm{Gal}(\bar{\mathbb{F}_{q}} /\mathbb{F}_{q^{m}}) is exactly the Fqm\mathbb{F}_{q^{m}}-points X(Fqm)X(\mathbb{F}_{q^{m}}). So xx is Fqm\mathbb{F}_{q^{m}}-point if and only if it’s in the intersection of ΓφΔ\Gamma_{\varphi}\cdot \Delta. It suffices to show each point has multiplicity 1. From the lemma, Γφ\Gamma_{\varphi} and Δ\Delta are all isomorphic to XX, so they are nonsingular. For every point (φ(x),x)ΓφΔ(\varphi(x),x)\in \Gamma_{\varphi}\cap \Delta, the tangent space is the product of corresponding tangent space XX. Note (dφ)x=0(d\varphi)_{x}=0, Γφ\Gamma_{\varphi} and Δ\Delta intersect transversally, so each point has single multiplicity. ◻

A natural question is what is the degree of Frobenius for the projective curve XX. The degree is a generic condition so we may focus on an affine open set. The case K(X)=Fq(t)K(X)=\mathbb{F}_{q}(t) is clear, the morphism is given by σ:Fq(t)Fq(tq)\sigma: \mathbb{F}_{q}(t)\to \mathbb{F}_{q}(t^{q}). For the general case, K(X)K(X) is finite algebraic over Fq(t)\mathbb{F}_{q}(t). Since the action on K(X)K(X) restrict on Fq(t)\mathbb{F}_{q}(t) is the action on Fq(t)\mathbb{F}_{q}(t), [K(X):Fq(t)]=[σK(X):σFq(t)][K(X):\mathbb{F}_{q}(t)]=[\sigma K(X):\sigma \mathbb{F}_{q}(t)], and the results follow from the multiplicity of extension degree.

Hodge Index Theorem and Hasse Weil Bound

Theorem 9 (Hasse-Weil).

Nm1qm2gqm\left| N_{m}-1-q^{m} \right| \leq 2g\sqrt{q^{m}}

To get the estimation of ΓφΔ\Gamma_{\varphi}\cdot \Delta, we need the Hodge index theorem. Recall some facts in the intersection theory: Fix an very ample line sheaf OX×X(1)\mathcal{O}_{X\times X}(1), and denote the corresponding hyperplane section by HH.

Lemma 10. Let L1\mathcal{L}_{1}, L2\mathcal{L}_{2} be two line sheaves on a variety XX. Then Hom(L1,L2)=0\mathrm{Hom}(\mathcal{L}_{1},\mathcal{L}_{2})=0 provided degL1>degL2\deg\mathcal{L}_{1}>\deg\mathcal{L}_{2}.

Proof

Proof. Since rk(L1)=α2(L1)α2(OX)\mathrm{rk}(\mathcal{L}_{1})=\frac{\alpha_{2}(\mathcal{L}_{1})}{\alpha_{2}(\mathcal{O}_{X})} and rk(L2)=α2(L2)α2(OX)\mathrm{rk}(\mathcal{L}_{2})=\frac{\alpha_{2}(\mathcal{L}_{2})}{\alpha_{2}(\mathcal{O}_{X})}, we have α2(L1)=α2(L2)\alpha_{2}(\mathcal{L}_{1})=\alpha_{2}(\mathcal{L}_{2}). The Hilbert polynomial P(L1)(n)<P(L2)(n)P(\mathcal{L}_{1})(n)<P(\mathcal{L}_{2})(n). Suppose there is a morphism ψ:L1L2\psi:\mathcal{L}_{1}\to \mathcal{L}_{2}. Let K\mathcal{K}, F\mathcal{F} be the kernel and the image of ψ\psi, respectively.

We first show the support of F\mathcal{F} is XX. Assume there is a point xXx\in X that is not in the support of F\mathcal{F}, since SuppF\mathrm{Supp}\mathcal{F} is closed, theres’s an open neighborhood UU that has empty intersection with SuppF\mathrm{Supp}\mathcal{F}. We can choose an affine open set W=SpecAW=\mathrm{Spec}A that has nonempty intersection with XUX-U and UU. Pick an nonzero global section ss supported on XUX-U and ff is an element in ideal which define WSuppsW\cap \mathrm{Supp}s. Then ss is zero in the localization of D(f)WD(f)\subset W, so sfn=0sf^{n}=0 for some nn. However, ss is also in L1(W)\mathcal{L}_{1}(W), which is torsion-free, contradict to the above argument.

The Hilbert polynomial of F\mathcal{F} is also a quadratic function. Since XX is integral, the rank of F\mathcal{F} is integer, and it is 1. Thus rk(K)=0\mathrm{rk}(\mathcal{K})=0, which means α2(K)=0\alpha_{2}(\mathcal{K})=0, and dimSuppK1\dim \mathrm{Supp}\mathcal{K}\leq 1. Using a similar argument, we can prove that K=0\mathcal{K}=0. Therefore, P(L1)(n)=P(F)(n)>P(L2)(n)P(\mathcal{L}_{1})(n)=P(\mathcal{F})(n)>P(\mathcal{L}_{2})(n). This contradicts the fact that F\mathcal{F} is a subsheaf of L2\mathcal{L}_{2}. ◻

Lemma 11. Let {Di}\{D_{i}\} be a set of divisor on X×XX\times X with bounded degree. Then {h0(OX×X(Di))}\{h^{0}(\mathcal{O}_{X\times X}(D_{i}))\} is bounded.

Proof

Proof. Assume degDi<m\deg D_{i}< m. We first reduce to the case degDi\deg D_{i} is a negative fixed number. For any divisor DD in the family and the corresponding line sheaf OX(D)=L\mathcal{O}_{X}(D)=\mathcal{L}, consider the exact sequence

0L(m)sLLC00\to \mathcal{L}(-m)\xrightarrow{\cdot s} \mathcal{L}\to \mathcal{L}|_{C}\to 0

where ss is a regular section of L\mathcal{L} and CC is the support of ss, by Bertini theorem, such ss exists. Then h0(L)h0(L(m))+h0(LC)h^{0}(\mathcal{L})\leq h^{0}(\mathcal{L}(-m))+h^{0}(\mathcal{L}|_{C}). Note that degC(LC)=degC(DC)=degX×X(D)\deg_{C}(\mathcal{L}|_{C})=\deg_{C}(D\cap C)=\deg_{X\times X}(D). The genus gCg_{C} is independent on the choice of CC, and is given by the adjunction formula gC=1+12(KEm+EmEm)g_{C}=1+\frac{1}{2}(K\cdot E_{m} +E_{m}\cdot E_{m}) where EmE_{m} is the divisor corresponding to O(m)\mathcal{O}(m) and therefore bounded.

If degC(LC)>2g2\deg_{C}(\mathcal{L}|_{C})>2g-2, then h0(LC)=degC(LC)+1gm+1gh^{0}(\mathcal{L}|_{C})=\deg_{C}(\mathcal{L}|_{C})+1-g\leq m+1-g.

If degC(LC)2g2\deg_{C}(\mathcal{L}|_{C})\leq 2g-2, then h0(LC)h0(LC(2g1m))=gh^{0}(\mathcal{L}|_{C})\leq h^{0}(\mathcal{L}|_{C}(2g-1-m))=g.

Then we have h0(LC)h^{0}(\mathcal{L}|_{C}) bounded and we reduce to the case deg(Di)\deg(D_{i}) is negative constant by twisting a small number. Since a global section in h0(OX(D))h^{0}(\mathcal{O}_{X}(D)) will induce a morphism from OX\mathcal{O}_{X} to OX(D)\mathcal{O}_{X}(D), by above lemma, OX(Di)\mathcal{O}_{X}(D_{i}) has no global sections. This finishes the proof. ◻

Theorem 12 (Hodge Index Theorem). Let DD be a divisor on X×XX\times X. If deg(D)=0\deg (D)=0, then DD<0D\cdot D<0.

Proof

Proof. By above lemma, {h0(nD)}\{h^{0}(nD)\} and {h0(KnD)}\{h^{0}(K-nD)\} are all bounded. Using Riemann-Roch theorem on surface,

h0(nD)h1(nD)+h0(KnD)=12nD(nDK)+1+pah^{0}(nD)-h^{1}(nD)+h^{0}(K-nD)=\frac{1}{2}nD(nD-K)+1+p_{a}

For large nn, left hand side of the equation is negative, so n2DD<0n^{2}D\cdot D<0. ◻

For any automorphism f:XXf:X\to X, we compute the self intersection number of Γf\Gamma_{f}. Let deg(f)=d\deg(f)=d. Using the adjunction formula, Γf(Γf+K)=2g2\Gamma_{f}\cdot (\Gamma_{f}+K)=2g-2. Since ωX×X=p1ωXp2ωX\omega_{X\times X}=p_{1}^{*}\omega_{X}\otimes p_{2}^{*}\omega_{X}, KX×X=p1KX+p2KXK_{X\times X}=p_{1}^{*}K_{X}+p_{2}^{*}K_{X}. Γfp1KX\Gamma_{f}\cdot p_{1}^{*}K_{X} is the same as the degΓf(p1ωXOΓf)=degΓf(Γfp1ωX)\deg_{\Gamma_{f}}(p_{1}^{*}\omega_{X}\otimes \mathcal{O}_{\Gamma_{f}})=\deg_{\Gamma_{f}}(\Gamma_{f}^{*}p_{1}^{*}\omega_{X}), since p1Γf=fp_{1}\circ \Gamma_{f}=f, it is also the same as degΓf(fωX)=deg(f)deg(ωX)=d(2g2)\deg_{\Gamma_{f}}(f^{*}\omega_{X})=\deg (f)\deg(\omega_{X})=d(2g-2). Similarly, Γfp2KX=2g2\Gamma_{f}\cdot p_{2}^{*}K_{X}=2g-2. So ΓfΓf=d(2g2)\Gamma_{f}\cdot \Gamma_{f}=-d(2g-2).

Let f=idf=id, the self intersection number ΔΔ=22g\Delta\cdot \Delta=2-2g.

XX£XXXSpeck¡fidfp2p1

Let d1(D)=D(P×X)d_{1}(D)=D\cdot (P\times X) and d2(D)=D(X×P)d_{2}(D)=D\cdot (X\times P') for some PP, PP' closed points in XX. We can choose the degree of PP and PP' to be 1.

Theorem 13. $$D\cdot D\leq 2d_{1}(D)d_{2}(D)$$

Proof

Proof. Let VV be the vector space generated by basis P×XP\times X, X×PX\times P' and DD. The intersection product defines a quadratic form

M=(01d1(D)10d2(D)d1(D)d2(D)DD)M=\begin{pmatrix} 0&1&d_{1}(D)\\ 1&0&d_{2}(D)\\ d_{1}(D)&d_{2}(D)&D\cdot D\end{pmatrix}

det(M)=2d1(D)d2(D)DD\det(M)=2d_{1}(D)d_{2}(D)-D\cdot D. If det(M)<0\det (M)<0, since MM is indefinite, we may choose A1,A2,A3A_{1},A_{2},A_{3} such that AiAj=0A_{i}\cdot A_{j}=0, and AiTMAi=aiA_{i}^{T}MA_{i}=a_{i} such that a1,a2>0a_{1}, a_{2}>0, a3<0a_{3}<0. Then appropriate linear combination can give a degree 0 divisor with positive self intersection. Contradict to Hodge index theorem. ◻

Back to our original theorem:

Proof

Proof of Theorem 9. By the above theorem, if we define DE=d1(E)d2(D)+d1(D)d2(E)DED*E=d_{1}(E)d_{2}(D)+d_{1}(D)d_{2}(E)-D\cdot E, then DD0D*D\geq 0. Apply above argument to XFqˉ×XFqˉX_{\bar{\mathbb{F}_{q}}}\times X_{\bar{\mathbb{F}_{q}}} and Frobenius automorphism φ\varphi. Using the Cauchy inequality for bilinear forms, we have ΓφΔ(ΔΔ)(ΓφΓφ)\left| \Gamma_{\varphi}*\Delta \right|\leq \sqrt{(\Delta*\Delta)(\Gamma_{\varphi}*\Gamma_{\varphi})}, that is

ΓφΔ1deg(φ)2gdeg(φ)\left| \Gamma_{\varphi}\cdot \Delta-1-\deg(\varphi) \right|\leq 2g\sqrt{\deg (\varphi)}

Main Theorem

We first give some equivalent expression of Zeta function.

Proposition 14. Let XX be a smooth projective varitey. Then

ZX(t)=x closed11tdeg(x).Z_{X}(t)=\prod_{x \text{ closed}}\frac{1}{1-t^{\deg(x)}}.

Proof

Proof. Set ndn_{d} be the number of closed points with degree dd in XX. By Proposition 4,

log(ZX(t))=m1Nmmtm=m1dmnddmtm=d1ndetdee=d1ndlog(1td).\begin{aligned} \log(Z_{X}(t))&=\sum_{m\geq 1}\frac{N_{m}}{m}t^{m}=\sum_{m\geq 1}\sum_{d\mid m}\frac{n_{d}\cdot d}{m}t^{m}\\ &= \sum_{d\geq 1}n_{d}\sum_{e} \frac{t^{de}}{e}\\ &=\sum_{d\geq 1}-n_{d}\log(1-t^{d}).\end{aligned}

We therefore have the desired formula. ◻

For curve XX, the degree of divisor xi closedaixi\sum\limits_{x_{i} \text{ closed}} a_{i} x_{i} can be realized as iaideg(xi)\sum_{i}a_{i}\deg(x_{i}). So once we expand ZX(t)=x closedi=0tideg(x)Z_{X}(t)=\prod\limits_{x\text{ closed}}\sum_{i=0}^{\infty}t^{i\cdot \deg(x)}, it immediately follows that ZX(t)=D effectivetdeg(D)Z_{X}(t)=\sum\limits_{D\text{ effective}}t^{\deg(D)} for Weil divisors.

The degree map deg:Pic(X)Z\deg:\mathop{\mathrm{Pic}}(X)\to \mathbb{Z} is a group homomorphism, so we may write its image as eZe\mathbb{Z}. Define the set of degree mm line bundles by Picm(X)\mathop{\mathrm{Pic}}^{m}(X). They are coset of Pic0(X)\mathop{\mathrm{Pic}}^{0}(X), so they are either empty or has size Pic0(X)\left| \mathop{\mathrm{Pic}}^{0}(X) \right|.

Theorem 15. If L=O(D)\mathcal{L}=\mathcal{O}(D) and DD is an effective divisor, then the number of effective divisors linearly equivalent to DD is qh0(L)1q1\frac{q^{h^{0}(\mathcal{L})}-1}{q-1}.

Proof

Proof. The number is exactly the size of O(D)\left| \mathcal{O}(D) \right|. See Ha2 Theorem II 7.7. ◻

From now on we assume XX is a smooth projective curve over Fq\mathbb{F}_{q} with genus gg. If deg(L)>2g2\deg(\mathcal{L})>2g-2, then h1(L)=0h^{1}(\mathcal{L})=0, so by Riemann-Roch h0(L)=deg(L)g+1h^{0}(\mathcal{L})=\deg(\mathcal{L})-g+1. We will show

Theorem 16 (Weil conjecture, curve case).

ZX(t)Z_{X}(t) is a rational function.

Let gg be the genus of XX. ZX(t)Z_{X}(t) satisfies the functional equation

ZX(1qt)=q1gt22g.Z_{X}(\frac{1}{qt})=q^{1-g}t^{2-2g}.

We can write $$Z_{X}(t)=\frac{P_{1}(t)}{P_{0}(t)P_{2}(t)},$$ where P0(t)=1tP_{0}(t)=1-t and P2(t)=1qtP_{2}(t)=1-qt. The betti numbers satisfies

22g=B0B1+B2.2-2g=B_{0}-B_{1}+B_{2}.

Pi(t)P_{i}(t) is a polynomial with integer coefficients, with roots αij\alpha_{ij} such that αij=qi/2\left| \alpha_{ij} \right| =q^{i /2}.

Proof

Proof of Theorem 16 part 1. As above, assume the image of deg:Pic(X)Z\deg:\mathop{\mathrm{Pic}}(X)\to \mathbb{Z} is eZe\mathbb{Z}. Let d0d_{0} be the smallest number such that d0e2g1d_{0}e\geq 2g-1. Let h=Pic0(X)h=\left| \mathop{\mathrm{Pic}}^{0}(X) \right|. Then

ZX(t)=deg(D)2g2tdeg(D)+dd0hqdeg+11q1tde.(1)Z_{X}(t)=\sum_{\deg(D)\leq 2g-2}t^{\deg(D)}+\sum_{d\geq d_{0}}h\frac{q^{de-g+1}-1}{q-1}t^{de}.\qquad (1)

deg(D)2g2tdeg(D)\sum\limits_{\deg(D)\leq 2g-2}t^{\deg(D)} is a polynomial in tet^{e} of degree less than or equal to 2g2e\frac{2g-2}{e}.

dd0hqdeg+11q1tde=hq1(q1g(qt)d0e1(qt)etd0e1te)\sum_{d\geq d_{0}}h\frac{q^{de-g+1}-1}{q-1}t^{de}=\frac{h}{q-1}(q^{1-g}\frac{(qt)^{d_{0}e}}{1-(qt)^{e}}-\frac{t^{d_{0}e}}{1-t^{e}})

is also a rational function in tet^{e}. So we may write ZX(t)=f(te)(1te)(1qete)Z_{X}(t)=\frac{f(t^{e})}{(1-t^{e})(1-q^{e}t^{e})}.

limt1(t1)ZX(t)=limt1h(t1)q1(q1g(qt)d0e1(qt)etd0e1te)=limt1htd0e(t1)(1te)(q1)=he(q1).\begin{aligned} \lim\limits_{t\to 1}(t-1)Z_{X}(t)&=\lim\limits_{t\to 1}\frac{h(t-1)}{q-1}(q^{1-g}\frac{(qt)^{d_{0}e}}{1-(qt)^{e}}-\frac{t^{d_{0}e}}{1-t^{e}})\\&=\lim_{t\to 1}-\frac{ht^{d_{0}e}(t-1)}{(1-t^{e})(q-1)}=\frac{h}{e(q-1)}. \end{aligned}

ZX(t)Z_{X}(t) has single pole at t=1t=1. Consider Zeta function for XFqeX_{\mathbb{F}_{q^{e}}}. By the universal property of base change XF(qe)m=X(Fqem)\left| X_{\mathbb{F}_({q}^{e})^{m}} \right| =\left| X(\mathbb{F}_{q^{em}}) \right|, thus

logZXFqe(te)=m1XFqe(Fqem)mtem=m1X(Fqem)mtem\log Z_{X_{\mathbb{F}_{q^{e}}}}(t^{e})=\sum_{m\geq 1}\frac{\left| X_{\mathbb{F}_{q^{e}}}(\mathbb{F}_{q^{em}}) \right| }{m}t^{em}=\sum_{m\geq 1}\frac{\left| X(\mathbb{F}_{q^{em}}) \right| }{m}t^{em}

Let ξ\xi be the primitive root of order ee. We have i=1eξil={0 if ele if el\sum\limits_{i=1}^{e}\xi^{il}=\left\{\begin{aligned} 0& \text{ if } e\nmid l\\ e& \text{ if } e\mid l \end{aligned}\right.. So

logZXFqe(te)=i=1el1X(Fql)lξiltl.\log Z_{X_{\mathbb{F}_{q^{e}}}}(t^{e})=\sum_{i=1}^{e}\sum_{l\geq 1}\frac{\left| X(\mathbb{F}_{q^{l}}) \right| }{l}\xi^{il}t^{l}.

ZXFqe(te)=i=1eZX(ξit)=i=1ef(ξiete)(1ξiete)(1qeξiete)=ZX(t)e.Z_{X_{\mathbb{F}_{q^{e}}}}(t^{e})=\prod_{i=1}^{e}Z_{X}(\xi^{i}t)=\prod_{i=1}^{e}\frac{f(\xi^{ie}t^{e})}{(1-\xi^{ie}t^{e})(1-q^{e}\xi^{ie}t^{e})}=Z_{X}(t)^{e}.

By a similar argument, ZXFqeZ_{X_{\mathbb{F}_{q^{e}}}} also has single pole at t=1t=1, so e=1e=1. So ZX(t)=f(t)(1t)(1qt)Z_{X}(t)=\frac{f(t)}{(1-t)(1-qt)} where ff is an integer coefficient polynomial. Back to equation (1), we have degf2g\deg f\leq 2g. ◻

Remark 17. The result is also true for singular curves. For a singular curve XX, denote the set of singular points by SS and U=X\SU=X\backslash S. Clearly X(Fqm)=S(Fqm)+U(Fqm)\left| X(\mathbb{F}_{q^{m}}) \right|=\left| S(\mathbb{F}_{q^{m}}) \right| +\left| U(\mathbb{F}_{q^{m}}) \right| and ZX(t)=ZS(t)ZU(t)Z_{X}(t)=Z_{S}(t)Z_{U}(t). One can easily check ZS(t)Z_{S}(t) is rational function since dim(S)=0\dim(S)=0.

Proof

Proof of Theorem 16 part 2. If g=0g=0, the only curve is P1\mathbb{P}^{1}. So Nm=q2m1qm1N_{m}=\frac{q^{2m}-1}{q^{m}-1} and ZP1(t)=exp(m1qm+1mtm)=1(1t)(1qt)Z_{\mathbb{P}^{1}}(t)=\exp(\sum\limits_{m\geq 1}\frac{q^{m}+1}{m}t^{m})=\frac{1}{(1-t)(1-qt)}. ZP1(t)Z_{\mathbb{P}^{1}}(t) obviously satisfies the functional equation.

If g1g\geq 1. Rewrite equation (1)

ZX(t)=d=02g2(LPicd(X)qh0(L)1q1)td+d2g1hqdg+11q1td=d=02g2(LPicd(X)qh0(L)1q1)td+d2g1hqdg+1q1tdd=0htdq1=d=02g2(LPicd(X)qh0(L)1q1)td+hqgt2g1(1qt)(q1)h(1t)(q1).\begin{aligned} Z_{X}(t)&=\sum_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\mathcal{L})}-1}{q-1})t^{d}+\sum_{d\geq 2g-1}h\frac{q^{d-g+1}-1}{q-1}t^{d}\\&=\sum_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\mathcal{L})}-1}{q-1})t^{d}+\sum_{d\geq 2g-1}\frac{hq^{d-g+1}}{q-1}t^{d}-\sum_{d=0}\frac{ht^{d}}{q-1}\\&=\sum_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\mathcal{L})}-1}{q-1})t^{d}+\frac{hq^{g}t^{2g-1}}{(1-qt)(q-1)}-\frac{h}{(1-t)(q-1)}. \end{aligned}

Let S1(t)=d=02g2(LPicd(X)qh0(L)1q1)tdS_{1}(t)=\sum\limits_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\mathcal{L})}-1}{q-1})t^{d} and S2(t)=hqgt2g1(1qt)(q1)h(1t)(q1)S_{2}(t)=\frac{hq^{g}t^{2g-1}}{(1-qt)(q-1)}-\frac{h}{(1-t)(q-1)}.

S2(1qt)=hqg(1qt)2g1(11t)(q1)h(11qt)(q1)=q1gt22g(hqgt2g1(1qt)(q1)h(1t)(q1))=q1gt22gS2(t).\begin{aligned} S_{2}(\frac{1}{qt})&=\frac{hq^{g}(\frac{1}{qt})^{2g-1}}{(1-\frac{1}{t})(q-1)}-\frac{h}{(1-\frac{1}{qt})(q-1)}\\&=q^{1-g}t^{2-2g}(\frac{hq^{g}t^{2g-1}}{(1-qt)(q-1)}-\frac{h}{(1-t)(q-1)})=q^{1-g}t^{2-2g}S_{2}(t).\end{aligned}

Note that LωXL1\mathcal{L}\mapsto \omega_{X}\otimes \mathcal{L}^{-1} gives a bijection between line bundles with degree lies in the interval [0,2g2][0,2g-2]. By Serre duality, h0(ωXL1)=h1(L)=h0(L)+g1deg(L)h^{0}(\omega_{X}\otimes \mathcal{L}^{-1})=h^{1}(\mathcal{L})=h^{0}(\mathcal{L})+g-1-\deg(\mathcal{L}).

S1(1qt)=d=02g2(LPicd(X)qh0(ωXL1)q1)(1qt)2g2d=d=02g2(LPicd(X)qh0(L)+g1dq1)(1qt)2g2d=q1gt22gS1(t).\begin{aligned} S_{1}(\frac{1}{qt})&=\sum_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\omega_{X}\otimes \mathcal{L}^{-1})}}{q-1})(\frac{1}{qt})^{2g-2-d}\\&=\sum_{d=0}^{2g-2}(\sum_{\mathcal{L}\in \mathop{\mathrm{Pic}}^{d}(X)}\frac{q^{h^{0}(\mathcal{L})+g-1-d}}{q-1})(\frac{1}{qt})^{2g-2-d}\\&=q^{1-g}t^{2-2g}S_{1}(t). \end{aligned}

Sum together, we have the functional equation ZX(1qt)=q1gt22gZX(t)Z_{X}(\frac{1}{qt})=q^{1-g}t^{2-2g}Z_{X}(t). ◻

Proof

Proof of Theorem 16 part 3. Write f(t)=i=12g(1ωit)f(t)=\prod\limits_{i=1}^{2g}(1-\omega_{i}t) (ωi\omega_{i} can be 0).

ZX(1qt)=i=12g(1ωit)(11qt)(11t)=q12gt22gi=12g(tqωi)(1t)(1qt).Z_{X}(\frac{1}{qt})=\frac{\prod\limits_{i=1}^{2g}(1-\omega_{i}t)}{(1-\frac{1}{qt})(1-\frac{1}{t})}=\frac{q^{1-2g}t^{2-2g}\prod\limits_{i=1}^{2g}(tq-\omega_{i})}{(1-t)(1-qt)}.

Therefore

qgi=12g(tωiq)=i=12g(1ωit)(2)q^{g}\prod\limits_{i=1}^{2g}(t-\frac{\omega_{i}}{q})=\prod\limits_{i=1}^{2g}(1-\omega_{i}t)\qquad (2)

Observing the degree on each side, an immediate consequence is ωi0\omega_{i}\neq 0 for all ii. Therefore deg(f)=2g\deg(f)=2g and B0B1+B2=22gB_{0}-B_{1}+B_{2}=2-2g ◻

Remark 18. Equation (2) also implies i=12gωi=qg\prod\limits_{i=1}^{2g}\omega_{i}=q^{g}. If we let ωiqωi\omega_{i}\mapsto \frac{q}{\omega_{i}} in equation (2), we have

qgi=12g(t1ωi)=i=12g(1ωit),q^{g}\prod_{i=1}^{2g}(t-\frac{1}{\omega_{i}})=\prod_{i=1}^{2g}(1-\omega_{i}t),

and therefore ωiqωi\omega_{i}\mapsto \frac{q}{\omega_{i}} won’t change the set {ωi}i=12g\{\omega_{i}\}_{i=1}^{2g}.

Proof

Proof of Theorem 16 part 4. By above analysis,

ZX(t)=i=12g(1ωit)(1t)(1qt)=exp(m1Nmmtm).Z_{X}(t)=\frac{\prod\limits_{i=1}^{2g}(1-\omega_{i}t)}{(1-t)(1-qt)}=\exp(\sum_{m\geq 1}\frac{N_{m}}{m}t^{m}).

m1Nmmtm=i=12glog(1ωit)log(1t)log(1qt)=m1m1(1+qmi=12gωim)tm.\begin{aligned} \sum_{m\geq 1}\frac{N_{m}}{m}t^{m}&=\sum_{i=1}^{2g}\log(1-\omega_{i}t)-\log(1-t)-\log(1-qt)\\&=\sum_{m}\frac{1}{m\geq 1}(1+q^{m}-\sum_{i=1}^{2g}\omega_{i}^{m})t^{m}. \end{aligned}

So Nm=1+qmi=12gωimN_{m}=1+q^{m}-\sum\limits_{i=1}^{2g}\omega_{i}^{m} Use Theorem 9, we have i=12gωim2gqm\sum\limits_{i=1}^{2g}\omega_{i}^{m}\leq 2g\sqrt{q^{m}}. Assume ω1\omega_{1} has the largest absolute value, then limmi=12gωimω1m\lim\limits_{m\to \infty}\left| \frac{\sum\limits_{i=1}^{2g}\omega_{i}^{m}}{\omega_{1}^{m}} \right| converges to a positive value. So we have i=12gωimkω1m\left| \sum\limits_{i=1}^{2g}\omega_{i}^{m} \right| \geq k\left| \omega_{1}^{m} \right| for m0m\gg 0. Thus ω1(2gk)1/mq\left| \omega_{1} \right|\leq (\frac{2g}{k})^{1 /m} \sqrt{q}. Take mm\to \infty, we have ω1q\left| \omega_{1} \right|\leq \sqrt{q}. By above remark, we know that qωi=ωjq\left| \frac{q}{\omega_{i}} \right| =\left| \omega_{j} \right| \leq \sqrt{q}. Therefore all ωi=q\left| \omega_{i} \right| =\sqrt{q}. ◻

Reference

[1]: The Stacks Project Authors. Stacks project, 2024.

[2]: Robin Hartshorne. Algebraic geometry, volume 52. Springer Science & Business Media, 2013.

[3]: Paul Monsky. p-adic Analysis and Zeta Functions, volume 4. Kinokuniya, 1970.

[4]: André Weil. Numbers of solutions of equations in finite fields. 1949.

评论