CRAG

\usepackage{cmbright}
\usepackage{caption}

Tags: #projects/my-talks #homotopy #arithmetic-geometry/Weil-conjectures #projects/my-talks/slides

Refs: Weil Conjectures Weil Conjectures Talks

A Quick Note

  • A big thanks to Daniel Litt for organizing this reading seminar, recommending papers, helping with questions!!

  • Goals for this talk:

    • Understand the Weil conjectures,
    • Understand why the relevant objects should be interesting,
    • See elementary but concrete examples,
    • Count all of the things!

Generating Functions

Varieties

Fix \(q\) a prime and \({ \mathbf{F} }_q\) the (unique) finite field with \(q\) elements, along with its (unique) degree \(n\) extensions \begin{align*} { \mathbf{F} }_{q^n} = \left\{{x\in \mkern 1.5mu\overline{\mkern-1.5mu{ \mathbf{F} }\mkern-1.5mu}\mkern 1.5mu_q {~\mathrel{\Big\vert}~}x^{q^n} - x = 0}\right\} \quad \forall~ n\in {\mathbf{Z}}^{\geq 1} \end{align*}

Definition (Projective Algebraic Varieties) Let \(J = \left\langle{f_1, \cdots, f_M}\right\rangle {~\trianglelefteq~}k[x_0, \cdots, x_n]\) be an ideal, then a projective algebraic variety \(X\subset {\mathbf{P}}^n_{ \mathbf{F} }\) can be described as \begin{align*} X = V(J) = \left\{{\mathbf{x} \in {\mathbf{P}}^n_{{ \mathbf{F} }_q} {~\mathrel{\Big\vert}~}f_1(\mathbf{x}) = \cdots = f_M(\mathbf{x}) = \mathbf{0}}\right\} \end{align*} where \(J\) is generated by homogeneous polynomials in \(n+1\) variables, i.e. there is a fixed \(d = \deg f_i \in {\mathbf{Z}}^{\geq 1}\) such that

\begin{align*} f(\mathbf{x}) = \sum_{\substack{\mathbf{I} = (i_1, \cdots, i_n) \\ \sum_j i_j = d}} \alpha_{\mathbf{I}} \cdot x_0^{i_1}\cdots x_n^{i_n} {\quad \operatorname{ and } \quad} f(\lambda \cdot \mathbf{x}) = \lambda^d f(\mathbf{x}), \lambda \in { \mathbf{F} }^{\times} .\end{align*}

Point Counts

  • For a fixed variety \(X\), we can consider its \({ \mathbf{F} }_q{\hbox{-}}\)points \(X({ \mathbf{F} }_q)\).
    • Note that \({\sharp}X({ \mathbf{F} }_q) < \infty\) is an integer
  • For any \(L/{ \mathbf{F} }_q\), we can also consider \(X(L)\)
    • For \([L: { \mathbf{F} }_q]\) finite, \({\sharp}X(L) < \infty\) and are integers for every such \(n\).
    • In particular, we can consider the finite counts \({\sharp}X({ \mathbf{F} }_{q^n})\) for any \(n\geq 2\).
  • So we can consider the sequence
\small

\begin{align*} [N_1, N_2, \cdots , N_n, \cdots ] \coloneqq[{\sharp}X({ \mathbf{F} }_q), ~ \sizeX ({ \mathbf{F} }_{q^2}), \cdots, ~{\sharp}X({ \mathbf{F} }_{q^n}), \cdots] .\end{align*}

\normalsize

  • Idea: associate some generating function (a formal power series) encoding the sequence, e.g.  \begin{align*}F(z) = \sum_{n=1}^\infty N_n z^n = N_1z + N_2 z^2 + \cdots .\end{align*}

Why Generating Functions?

For an ordinary generating function, the coefficients are related to the real-analytic properties of \(F\): \begin{align*} [z^n] \cdot F(z) = [z^n]\cdot T_{F, z=0}(z) = \frac{1}{n!} \qty{{\frac{\partial }{\partial z}\,}}^{n} F(z) \Bigg\rvert_{z = 0} = N_n \end{align*}

and also to the complex analytic properties: \begin{align*} [z^n] \cdot F(z) \coloneqq\frac{1}{2\pi i} \oint_{{\mathbb{S}}^1} {F(z) \over z^{n+1}}~dz = \frac{1}{2\pi i} \oint_{{\mathbb{S}}^1} \frac{N_n}{z} ~dz = N_n .\end{align*}

Using the Residue theorem. The latter form is very amenable to computer calculation.

Why Generating Functions?

  • An OGF is an infinite series, which we can interpret as an analytic function \({\mathbf{C}}\to {\mathbf{C}}\)

  • In nice situations, we can hope for a closed-form representation.

  • A useful example: by integrating a geometric series we can derive

\begin{align*} {1 \over 1-z} &= \sum_{n=0}^\infty z^n \hspace{5em} \qty{ = 1 + z + z^2 + \cdots} \\ \implies \int {1 \over 1-z} &= \int \sum_{n = 0}^\infty z^n \\ &= \sum_{n=0}^\infty \int z^n \quad{for } {\left\lvert {z} \right\rvert} < 1 {\quad \operatorname{by uniform convergence} \quad} \\ &= \sum_{n=0}^\infty \frac{1}{n+1}z^{n+1} \\ \implies -\log\qty{1-z} &= \sum_{n=1}^\infty {z^n \over n} \hspace{5em} \qty{= z + {z^2 \over 2} + {z^3 \over 3} + \cdots} .\end{align*}

Exponential

  • For completeness, also recall that \begin{align*} \exp(z) \coloneqq\sum_{n=0}^\infty {z^n \over n!} \end{align*}

  • We can regard \(\exp, \log\) as elements in the ring of formal power series \({\mathbf{Q}}[z](z)\).

  • In particular, for any \(p(z) \in z\cdot {\mathbf{Q}}[[z]]\) we can consider \(\exp(p(z)), \log(1 + p(z))\)

  • Since \({\mathbf{Q}}\hookrightarrow{\mathbf{C}}\), we can consider these as a complex-analytic functions, ask where they converge, etc.

Zeta Functions

Definition: Local Zeta Function

Problem: count points of a (smooth?) projective variety \(X/{ \mathbf{F} }\) in all (finite) degree \(n\) extensions of \({ \mathbf{F} }\).

Definition (Local Zeta Function) The local zeta function of an algebraic variety \(X\) is the following formal power series:

\begin{align*} Z_X(z) &= \exp\qty{ \sum_{n=1}^\infty N_n {z^n \over n} } \in {\mathbf{Q}}[[z]] {\quad \operatorname{where} \quad} N_n \coloneqq{\sharp}X({ \mathbf{F} }_n) .\end{align*}

Note that \begin{align*} z \qty{{\frac{\partial }{\partial z}\,}} \log Z_X(z) &= z {\frac{\partial }{\partial z}\,} \qty{N_1z + N_2 {z^2 \over 2} + N_3 {z^3 \over 3} + \cdots} \\ &= z \qty{N_1 + N_2 z + N_3 z^2 + \cdots } \hspace{5em} \text{(unif. conv.)}\\ &= N_1 z + N_2 z^2 + \cdots = \sum_{n=1}^\infty N_n z^n ,\end{align*} which is an ordinary generating function for the sequence \((N_n)\).

Examples

Example: A Point

Take \(X = {\operatorname{pt}}= V(\left\{{f(x) = x-c}\right\})/{ \mathbf{F} }\) for \(c\) a fixed element of \({ \mathbf{F} }\). This yields a single point over \({ \mathbf{F} }\), then \begin{align*} \sizeX({ \mathbf{F} }_q) &\coloneqq N_1 = 1 \\ \sizeX({ \mathbf{F} }_{q^2}) &\coloneqq N_2 = 1 \\ &\vdots \\ \sizeX({ \mathbf{F} }_{q^n}) &\coloneqq N_n = 1 \end{align*}

and so \begin{align*} Z_{\operatorname{pt}}(z) &= \exp\qty{ 1\cdot z + 1 \cdot {z^2 \over 2} + 1 \cdot {z^3 \over 3} + \cdots } = \exp\qty{\sum_{n=1}^\infty {{z}^n \over n} } \\ &= \exp\qty{ -\log\qty{1-z} } \\ &= {1 \over 1 - z} .\end{align*}

Notice: \(Z\) admits a closed form and is a rational function.

Example: The Affine Line

Take \(X = {\mathbf{A}}^1/{ \mathbf{F} }\) the affine line over \({ \mathbf{F} }\), then We can write \begin{align*} {\mathbf{A}}^1({ \mathbf{F} }_{q^n}) = \left\{{\mathbf{x} = [x_1] {~\mathrel{\Big\vert}~}x_1 \in { \mathbf{F} }_{q^n}}\right\} \end{align*} as the set of one-component vectors with entries in \({ \mathbf{F} }_n\), so

\begin{align*} X({ \mathbf{F} }_q) &= q \\ X({ \mathbf{F} }_{q^2}) &= q^2 \\ &\vdots \\ X({ \mathbf{F} }_{q^n}) &= q^n .\end{align*}

Then \begin{align*} Z_X(z) &= \exp\qty{\sum_{n=1}^\infty q^n {z^n \over n} } \\ &= \exp\qty{\sum_{n=1}^\infty {\qty{qz}^n \over n} } \\ &= \exp(-\log(1 - qz)) \\ &= \frac 1 {1 - qz} .\end{align*}

Example: Affine m-space

Take \(X = {\mathbf{A}}^m/{ \mathbf{F} }\) the affine line over \({ \mathbf{F} }\), then We can write \begin{align*} {\mathbf{A}}^m({ \mathbf{F} }_{q^n}) = \left\{{\mathbf{x} = [x_1, \cdots, x_m] {~\mathrel{\Big\vert}~}x_i \in { \mathbf{F} }_{q^n}}\right\} \end{align*}

as the set of one-component vectors with entries in \({ \mathbf{F} }_n\), so

\noindent

\begin{minipage}{.45\textwidth}
\small

\[
X({ \mathbf{F} }_q) &= q^m \\
X({ \mathbf{F} }_{q^2}) &= (q^2)^m \\
&\vdots \\
X({ \mathbf{F} }_{q^n}) &= q^{nm}
.\]

   \normalsize
\end{minipage}
\begin{minipage}{.45\textwidth}
   \centering
    \includegraphics[width = .5\linewidth]{figures/image_2020-04-23-00-21-46.png}
\captionof{figure}{${\mathbf{A}}^2 / { \mathbf{F} }_3 ~(q=3, m = 2, n = 1)$}
\end{minipage}

Then

\small

\begin{align*} Z_X(z) &= \exp\qty{\sum_{n=1}^\infty q^{nm} {z^n \over n} } = \exp\qty{\sum_{n=1}^\infty {\qty{q^m z}^n \over n} } \\ &= \exp(-\log(1 - q^mz)) \\ &= \frac 1 {1 - q^mz} .\end{align*}

\normalsize

Example: Projective Line

Take \(X = {\mathbf{P}}^1/{ \mathbf{F} }\), we can still count by enumerating coordinates:

\small
\begin{align*} {\mathbf{P}}^1({ \mathbf{F} }_{q^n}) &= \left\{{[x_1: x_2] {~\mathrel{\Big\vert}~}x_1, x_2 \neq 0 \in { \mathbf{F} }_{q^n}}\right\}/\sim ~= \left\{{[x_1: 1] {~\mathrel{\Big\vert}~}x_1 \in { \mathbf{F} }_{q^n}}\right\} {\textstyle\coprod}\left\{{[1: 0]}\right\} .\end{align*}
\normalsize

Thus

\noindent

\begin{minipage}{.45\textwidth}
\small
    \[
    X({ \mathbf{F} }_q) &= q + 1\\
    X({ \mathbf{F} }_{q^2}) &= q^2 + 1 \\
    &\vdots \\
    X({ \mathbf{F} }_{q^n}) &= q^n + 1
    .\]
    \normalsize
\end{minipage}
\begin{minipage}{.45\textwidth}
   \centering
    \includegraphics[width = .5\linewidth]{figures/image_2020-04-23-00-07-42.png}
\captionof{figure}{${\mathbf{P}}^1 / { \mathbf{F} }_3 ~(q=3, n = 1)$}
\end{minipage}

Thus

\small
\begin{align*} Z_X(z) &= \exp\qty{\sum_{n=1}^\infty \qty{q^n + 1} {z^n \over n} } \\ &= \exp\qty{\sum_{n=1}^\infty {q^n } {z^n \over n} + \sum_{n=1}^\infty 1 \cdot {z^n \over n}}\\ &= {1 \over (1-qz)(1-z)} .\end{align*}
\normalsize

The Weil Conjectures

Weil 1

(Weil 1949)

Let \(X\) be a smooth projective variety of dimension \(N\) over \({ \mathbf{F} }_{q}\) for \(q\) a prime, let \(Z_X(z)\) be its zeta function, and define \(\zeta_X(s) = Z_X(q^{-s})\).

  • (Rationality)

    \(Z_X(z)\) is a rational function:

    \begin{align*} Z_X(z) &= {p_1(z) \cdot p_3(z) \cdots p_{2N-1}(z) \over p_0(z) \cdot p_2(z) \cdots p_{2N}(z)} \in {\mathbf{Q}}(z),\quad\text{i.e. }\quad p_i(z) \in {\mathbf{Z}}[z] \\ \\ P_0(z) &= 1-z \\ P_{2N}(z) &= 1 - q^N z \\ P_j(z) &= \prod_{j=1}^{\beta_i} \qty{1 - a_{j, k} z} {\quad \operatorname{for some reciprocal roots} \quad} a_{j, k} \in {\mathbf{C}} \end{align*} where we’ve factored each \(P_i\) using its reciprocal roots \(a_{ij}\).

    \small
    In particular, this implies the existence of a meromorphic continuation of the associated function \(\zeta_X(s)\), which a priori only converges for \(\Re(s)\gg 0\). This also implies that for \(n\) large enough, \(N_n\) satisfies a linear recurrence relation.
    \normalsize

Weil 2

  • (Functional Equation and Poincare Duality)

    Let \(\chi(X)\) be the Euler characteristic of \(X\), i.e. the self-intersection number of the diagonal embedding \(\Delta \hookrightarrow X\times X\); then \(Z_X(z)\) satisfies the following functional equation:

    \begin{align*} Z_X\qty{1 \over q^N z} = \pm \qty{q^{N \over 2} z}^{\chi(X)} ~~Z_X(z) .\end{align*}

    Equivalently, \begin{align*} \zeta_X(N-s) = \pm \qty{q^{\frac N 2 - s}}^{\chi(X)} ~\zeta_X(s) .\end{align*}

    \small 
    Note that when \(N=1\), e.g. for a curve, this relates \(\zeta_X(s)\) to \(\zeta_X(1-s)\).
    \normalsize

    Equivalently, there is an involutive map on the (reciprocal) roots \begin{align*} z &\iff {q^N \over z} \\ \alpha_{j, k} &\iff \alpha_{2N-j, k} \end{align*} which sends interchanges the coefficients in \(p_j\) and \(p_{2N-j}\).

Weil 3

  • (Riemann Hypothesis)

    The reciprocal roots \(\alpha_{j,k}\) are algebraic integers (roots of some monic \(p\in {\mathbf{Z}}[x]\)) which satisfy \begin{align*} {\left\lvert {\alpha_{j,k}} \right\rvert}_{\mathbf{C}}= q^{j \over 2}, \quad\quad 1 \leq j \leq 2N-1,~ \forall k .\end{align*}

  • (Betti Numbers)

    If \(X\) is a “good reduction mod \(q\)” of a nonsingular projective variety \(\tilde X\) in characteristic zero, then the \(\beta_i = \deg p_i(z)\) are the Betti numbers of the topological space \(\tilde X({\mathbf{C}})\).

Moral:

  • The Diophantine properties of a variety’s zeta function are governed by its (algebraic) topology.
  • Conversely, the analytic properties of encode a lot of geometric/topological/algebraic information.

Why is (3) called the “Riemann Hypothesis”?

Recall the Riemann zeta function is given by \begin{align*} \zeta(s) = \sum_{n=1}^\infty \frac{1}{n^s} = \prod_{p\text{ prime}} {1 \over 1 - p^{-s}} .\end{align*}

After modifying \(\zeta\) to make it symmetric about \(\Re(s) = \frac 1 2\) and eliminate the trivial zeros to obtain \(\widehat{\zeta}(s)\), there are three relevant properties

  • “Rationality”: \(\widehat{\zeta}(s)\) has a meromorphic continuation to \({\mathbf{C}}\) with simple poles at \(s=0, 1\).
  • “Functional equation”: \(\widehat{\zeta}(1-s) = \widehat{\zeta}(s)\)
  • “Riemann Hypothesis”: The only zeros of \(\widehat{\zeta}\) have \(\Re(s) = \frac 1 2\).

Why is (3) called the “Riemann Hypothesis”?

Suppose it holds for some \(X\). Use the facts:

  • \({\left\lvert {\exp\qty{z}} \right\rvert} = \exp\qty{\Re(z)}\) and
  • \(a^z \coloneqq\exp\qty{z ~\operatorname{Log}(a)}\),

and to replace the polynomials \(P_j\) with \begin{align*} L_j(s) \coloneqq P_j(q^{-s}) = \prod_{k=1}^{\beta_j} \qty{1 - \alpha_{j, k} q^{-s}} .\end{align*}

Analogy to Riemann Hypothesis

Now consider the roots of \(L_j(s)\): we have

\small
\begin{align*} L_j(s_0) = 0 \iff (1 - \alpha_{j, k} q^{-s}) &= 0 {\quad \operatorname{for some} \quad} k\\ \iff q^{-s_0} &= {1 \over \alpha_{j, k}} \\ \iff {\left\lvert {q^{-s_0}} \right\rvert} &= {\left\lvert {1 \over \alpha_{j, k}} \right\rvert} \quad\quad \stackrel{\text{\tiny by assumption}}{=} q^{ -{j \over 2}} \\ \\ \iff q^{-\frac j 2} \stackrel{(a)}= \exp\qty{- \frac j 2 \cdot \operatorname{Log}(q)} &= {\left\lvert { \exp\qty{-s_0 \cdot \operatorname{Log}(q)} } \right\rvert} \\ &\stackrel{(b)}= {\left\lvert { \exp\qty{-\qty{\Re(s_0) + i\cdot \Im(s_0)} \cdot \operatorname{Log}(q)} } \right\rvert} \\ &\stackrel{(a)}= \exp\qty{-\qty{\Re(s_0)} \cdot \operatorname{Log}(q)} \\ \\ \iff - \frac j 2 \cdot \operatorname{Log}(q) &= -\Re(s_0) \cdot \operatorname{Log}(q) {\quad \operatorname{by injectivity} \quad} \\ \iff \Re(s_0) = \frac j 2 .\end{align*}
\normalsize

Analogy with Riemann Hypothesis

Roughly speaking, we can apply \(\operatorname{Log}\) (a conformal map) to send the \(\alpha_{j, k}\) to zeros of the \(L_j\), this says that the zeros all must lie on the “critical lines” \(\frac{j}{2}\).

\begin{figure}[h]
\centering
\includegraphics[width = 1.0\textwidth]{figures/image_2020-04-23-02-26-27.png} 
\end{figure}

In particular, the zeros of \(L_1\) have real part \(\frac 1 2\) (analogy: classical ../RH).

Precise Relation

  • Difficult to find in the literature!
  • Idea: make a similar definition for schemes, then take \(X = \operatorname{Spec}{\mathbf{Z}}\).
  • Define the “reductions mod \(p\)\(X_p\) for closed points \(p\).
  • Define the local zeta functions \(\zeta_{X_p}(s) = Z_{X_p}(p^{-s})\).
  • (Potentially incorrect) Evaluate to find \(Z_{X_p}(z) = {1 \over 1 - z}\).
  • Take a product over all closed points to define

\begin{align*} L_X(s) &= \prod_{p\text{ prime}} \zeta_{X_p}(p^{-s}) \\ &= \prod_{p\text{ prime}} \qty{ 1 \over 1 - p^{-s}} \\ &= \zeta(s) ,\end{align*}

which is the Euler product expansion of the classical Riemann Zeta function.

If anyone knows a reference for this, let me know!

Weil for Curves

Weil for Curves

The Weil conjectures take on a particularly nice form for curves. Let \(X/{ \mathbf{F} }_q\) be a smooth projective curve of genus \(g\), then

  • (Rationality) \begin{align*}Z_X(z) = {p_1(z) \over p_0(z) p_2(z)} = {p_1(z) \over (1-z)(1-qz)}\end{align*}
  • (Functional Equation) \begin{align*}Z_X\qty{1 \over qz} = \qty{z\sqrt q}^{2-2g} Z_X(z)\end{align*}
  • (Riemann Hypothesis) \begin{align*}p_1(z) = \prod_{i=1}^{\beta_1} (1 - a_iz ) {\quad \operatorname{where} \quad} {\left\lvert {a_i} \right\rvert} = \sqrt{q}\end{align*}
  • (Betti Numbers) \({\mathcal{P}}_{\Sigma_g}(x) = 1 + 2g\cdot x + x^2 \implies \deg p_1 = \beta_1 = 2g\).
\tiny

\({\mathcal{P}}\) here is the Poincaré polynomial, the generating function for the Betti numbers. \(\Sigma_g\) is the surface (real 2-dimensional smooth manifold) of genus \(g\).

\normalsize

The Projective Line

Recall \(Z_{{\mathbf{P}}^1/{ \mathbf{F} }_q}(z) = {1 \over (1-z)(1-qz)}\).

  • Rationality: Clear!

  • Functional Equation: \(g=0 \implies 2g-2 = 2\) \begin{align*} Z_{{\mathbf{P}}^1}\qty{{1 \over qz}} = {1 \over (1-{1 \over qz})(1 - {q \over qz}) } = {qz^2 \over (1-z)(1-qz)} = {\qty{\sqrt q z}^2 \over (1-z)(1-qz) } .\end{align*}

  • Riemann Hypothesis: Nothing to check (no \(p_1(z)\))

  • Betti numbers: Use the fact that \({\mathcal{P}}_{{\mathbf{CP}}^1} = 1 + 0\cdot x + x^2\), and indeed \(\deg p_0 = \deg p_2 = 1,~ \deg p_1 = 0\).

Note that even Betti numbers show up as degrees in the denominator, odd in the numerator. Allows us to immediately guess the zeta function for \({\mathbf{P}}^n/{ \mathbf{F} }_q\) by knowing \(H^* {\mathbf{CP}}^\infty\)!

Elliptic Curves

\begin{figure}[h]
\caption{Some Elliptic Curves}
\centering
\includegraphics[width = 0.9\textwidth]{figures/image_2020-04-23-02-40-31.png} 
\end{figure}

Consider \(E/{ \mathbf{F} }_q\).

  • (Nontrivial!) The number of points is given by \begin{align*} N_n \coloneqq E({ \mathbf{F} }_{q^n}) = (q^n + 1) - ( \alpha^n + {\mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu}^n ) {\quad \operatorname{where} \quad} {\left\lvert {\alpha} \right\rvert} = {\left\lvert {\mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu} \right\rvert} = \sqrt{q} \end{align*}
    • Proof: Involves trace (or eigenvalues?) of Frobenius, (could use references)
  • \(\dim_{\mathbf{C}}E/{\mathbf{C}}= N = 1\) and \(g=1\).

The Weil Conjectures say we should be able to write \begin{align*} Z_E(z) = {p_1(z) \over p_0(z) p_2(z)} = {p_1(z) \over (1-z) (1 - qz)} = { (1 - a_{1}z)(1 - a_{2}z) \over (1-z)(1- qz)} .\end{align*}

Elliptic Curves: Weil 1

  • Rationality: using the point count, we can compute
    \small
    \begin{align*} Z_E(z) &= \exp \sum_{n=1}^\infty {\sharp}E({ \mathbf{F} }_{q^n}) {z^n \over n} \\ &= \exp \sum_{n=1}^\infty \qty{q^n + 1 - \qty{\alpha^n + \mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu^n}} {z^n \over n} \\ &= \exp \qty{ \sum_{n=1}^\infty q^n\cdot {z^n \over n} } \exp \qty{ \sum_{n=1}^\infty 1\cdot {z^n \over n} } \\ & \hspace{5em} \exp \qty{ \sum_{n=1}^\infty -\alpha^n \cdot {z^n \over n} } \exp \qty{ \sum_{n=1}^\infty -\mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu^n\cdot{z^n \over n} } \\ \\ &= \exp\qty{-\log\qty{1-qz} } \cdot \exp\qty{-\log\qty{1-z} } \\ &\hspace{5em} \exp\qty{\log\qty{1- \alpha z} } \cdot \exp\qty{\log\qty{1 - \mkern 1.5mu\overline{\mkern-1.5mu\alpha \mkern-1.5mu}\mkern 1.5muz} } \\ \\ &= {(1-\alpha z)(1-\mkern 1.5mu\overline{\mkern-1.5mu\alpha \mkern-1.5mu}\mkern 1.5muz) \over (1-z)(1-qz)} \in {\mathbf{Q}}(z) ,\end{align*}
    \normalsize
    which is a rational function of the expected form (Weil 1).
\tiny

Note that the “expected” point counts show up in the denominator, along with the even Betti numbers, while the “correction” factor appears in the denominator and odd Betti numbers.

\normalsize

Elliptic Curves: Weil 2 and 3

  • Functional Equation: we use the equivalent formulation of “Poincaré duality”: \begin{align*} {(1-\alpha z)(1-\mkern 1.5mu\overline{\mkern-1.5mu\alpha \mkern-1.5mu}\mkern 1.5muz) \over (1-z)(1-qz)} = {p_1(z) \over p_0(z) p_2(z)} \implies \begin{cases} z &\iff {q \over z} \\ \alpha_{j, k} &\iff \alpha_{2-j, k} \end{cases} \end{align*}

This amounts to checking that the coefficients of \(p_0, p_2\) are interchanged, and that the two coefficients of \(p_1\) are interchanged:

\begin{align*} \mathrm{Coefs}(p_0) = \left\{{1}\right\} &\xrightarrow{z\to {q\over z}} \left\{{1 \over q}\right\} = \mathrm{Coefs}(p_2) \\ \mathrm{Coefs}(p_1) = \left\{{\alpha, \mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu}\right\} &\xrightarrow{z\to {q\over z}} \left\{{{q \over \alpha}, {q \over \mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu} }\right\} = \left\{{\mkern 1.5mu\overline{\mkern-1.5mu\alpha\mkern-1.5mu}\mkern 1.5mu, \alpha}\right\} {\quad \operatorname{using} \quad} \alpha\mkern 1.5mu\overline{\mkern-1.5mu\alpha \mkern-1.5mu}\mkern 1.5mu= q .\end{align*}

  • RH: Assumed as part of the point count (\({\left\lvert {\alpha } \right\rvert}= q^{1\over 2}\))

  • Betti Numbers: \({\mathcal{P}}_{\Sigma_1}(x) = 1 + 2x + x^2\), and indeed \(\deg p_0 = \deg p_2 = 1,~ \deg p_1 = 2\).

History

  • 1801, Gauss: Point count and RH showed for specific elliptic curves

  • 1924, Artin: Conjectured for algebraic curves ,

  • 1934, Hasse: proved for elliptic curves.

  • 1949, Weil: Proved for smooth projective curves over finite fields, proposed generalization to projective varieties

  • 1960, Dwork: Rationality via \(p{\hbox{-}}\)adic analysis

  • 1965, Grothendieck et al.: Rationality, functional equation, Betti numbers using étale cohomology

    • Trace of Frobenius on \(\ell{\hbox{-}}\)adic cohomology
    • Expected proof via the standard conjectures. Wide open!
  • 1974, Deligne: Riemann Hypothesis using étale cohomology, circumvented the standard conjectures

  • Recent: Hasse-Weil conjecture for arbitrary algebraic varieties over number fields

    • Similar requirements on \(L{\hbox{-}}\)functions: functional equation, meromorphic continuation
    • 2001: Full modularity theorem proved, extending Wiles, implies Hasse-Weil for elliptic curves
    • Inroad to The Langlands Program: show every \(L\) function coming from an algebraic variety also comes from an automorphic representation.

Weil for Projective m-space

Setup

Take \(X = {\mathbf{P}}^m/{ \mathbf{F} }\) We can write \begin{align*} {\mathbf{P}}^m({ \mathbf{F} }_{q^n}) = {\mathbf{A}}^{m+1}({ \mathbf{F} }_{q^n}) \setminus \left\{{\mathbf{0}}\right\}/\sim ~= \left\{{\mathbf{x} = [x_0, \cdots, x_m] {~\mathrel{\Big\vert}~}x_i \in { \mathbf{F} }_{q^n}}\right\}/\sim \end{align*}

But how many points are actually in this space?

\begin{figure}[h]
\caption{Points and Lines in ${\mathbf{P}}^2/{{ \mathbf{F} }_3}$}
\centering
\includegraphics[width = 0.4\textwidth]{figures/image_2020-04-19-22-32-07.png} 
\end{figure}

A nontrivial combinatorial problem!

q-Analogs and Grassmannians

To illustrate, this can be done combinatorially: identify \({\mathbf{P}}^m_{ \mathbf{F} }= {\operatorname{Gr}}_{{ \mathbf{F} }}(1, m+1)\) as the space of lines in \({\mathbf{A}}^{m+1}_{ \mathbf{F} }\).

Theorem
The number of \(k{\hbox{-}}\)dimensional subspaces of \({\mathbf{A}}^N_{{ \mathbf{F} }_q}\) is the \(q{\hbox{-}}\)analog of the binomial coefficient: \begin{align*} \genfrac{[}{]}{0pt}{}{N}{k}_q \coloneqq\frac{(q^N - 1)(q^{N-1}-1) \cdots (q^{N - (k-1)} - 1)}{(q^k-1)(q^{k-1} - 1) \cdots (q-1)} .\end{align*}

Remark: Note \(\lim_{q\to 1} \genfrac{[}{]}{0pt}{}{N}{k}_q = {N \choose k}\), the usual binomial coefficient.

Proof: To choose a \(k{\hbox{-}}\)dimensional subspace,

  • Choose a nonzero vector \(\mathbf{v}_1 \in {\mathbf{A}}^n_{ \mathbf{F} }\) in \(q^N - 1\) ways.
    • For next step, note that \({\sharp}\mathop{\mathrm{span}}\left\{{\mathbf{v}_1}\right\} = {\sharp}\left\{{\lambda \mathbf{v}_1 {~\mathrel{\Big\vert}~}\lambda \in { \mathbf{F} }_q}\right\} = {\sharp}{ \mathbf{F} }_q = q\).
  • Choose a nonzero vector \(\mathbf{v}_2\) not in the span of \(\mathbf{v}_1\) in \(q^N - q\) ways.
    • Now note \({\sharp}\mathop{\mathrm{span}}\left\{{\mathbf{v}_1, \mathbf{v}_2}\right\} = {\sharp}\left\{{\lambda_1 \mathbf{v}_1 + \lambda_2 \mathbf{v}_2 {~\mathrel{\Big\vert}~}\lambda_i \in { \mathbf{F} }}\right\} = q\cdot q = q^2\).

Proof continued

  • Choose a nonzero vector \(\mathbf{v}_3\) not in the span of \(\mathbf{v}_1, \mathbf{v}_2\) in \(q^N -q^2\) ways.

  • \(\cdots\) until \(\mathbf{v}_k\) is chosen in \begin{align*}(q^N-1)(q^N-q) \cdots (q^N - q^{k-1}) {\quad \operatorname{ways} \quad}.\end{align*}

    • This yields a \(k{\hbox{-}}\)tuple of linearly independent vectors spanning a \(k{\hbox{-}}\)dimensional subspace \(V_k\)
  • This overcounts because many linearly independent sets span \(V_k\), we need to divide out by the number of ways to choose a basis inside of \(V_k\).

  • By the same argument, this is given by \begin{align*}(q^k-1)(q^k-q) \cdots (q^k - q^{k-1})\end{align*}

Thus

\small
\begin{align*} {\sharp}\text{subspaces} &= \frac{ (q^N-1)(q^N-q)(q^N - q^2) \cdots (q^N - q^{k-1}) }{ (q^k-1)(q^k-q)(q^k-q^2) \cdots (q^k - q^{k-1})}\\ \\ &= {q^N - 1 \over q^k - 1} \cdot \qty{q \over q} {q^{N-1} - 1 \over q^{k-1} - 1} \cdot \qty{q^2 \over q^2}{q^{N-2} - 1 \over q^{k-2} - 1} \cdots \qty{q^{k-1} \over q^{k-1}}{q^{N - (k-1)} - 1 \over q^{k - (k-1) - 1}} \\ &= \frac{(q^N - 1)(q^{N-1}-1) \cdots (q^{N - (k-1)} - 1)}{(q^k-1)(q^{k-1} - 1) \cdots (q-1)} .\end{align*}
\normalsize

\(\hfill\blacksquare\)

Counting Points

Note that we’ve actually computed the number of points in any Grassmannian.

Identify \({\mathbf{P}}^m_{ \mathbf{F} }= {\operatorname{Gr}}_{{ \mathbf{F} }}(1, m+1)\) as the space of lines in \({\mathbf{A}}^{m+1}_{ \mathbf{F} }\).

We obtain a simplification (importantly, a sum formula) when setting \(k=1\): \begin{align*} \genfrac{[}{]}{0pt}{}{m+1 }{1}_q = {q^{m+1}-1 \over q - 1} = q^{m} + q^{m-1} + \cdots + q + 1 = \sum_{j=0}^{m} q^j .\end{align*}

Thus \begin{align*} X({ \mathbf{F} }_q) &= \sum_{j=0}^{m} q^j \\ X({ \mathbf{F} }_{q^2}) &= \sum_{j=0}^{m} \qty{q^2}^j \\ &\vdots \\ X({ \mathbf{F} }_{q^n}) &= \sum_{j=0}^{m} \qty{q^n}^j .\end{align*}

Computing the Zeta Function

So \begin{align*} Z_X(z) &= \exp\qty{ \sum_{n=1}^\infty \sum_{j=0}^{m} \qty{q^n}^j {z^n \over n} } \\ &= \exp\qty{ \sum_{n=1}^\infty \sum_{j=0}^{m} {\qty{q^jz}^n \over n} } \\ &= \exp\qty{ \sum_{j=0}^{m} \sum_{n=1}^\infty {\qty{q^jz}^n \over n} } \\ &= \exp\qty{ \sum_{j=0}^{m-1} -\log(1 - q^j z) } \\ &= \prod_{j=0}^{m} \qty{1-q^jz}^{-1}\\ &= \qty{1 \over 1 - z} \qty{1 \over 1 - qz} \qty{1 \over 1 - q^2 z} \cdots \qty{1 \over 1- q^m z} ,\end{align*}

Miraculously, still a rational function! Consequence of sum formula, works in general.

Checking the Weil Conjectures

\begin{align*} Z_X(z) = \prod_{j=0}^m \qty{1 \over 1 - q^j z} .\end{align*}

  • Rationality: Clear!

  • Functional Equation: Less clear, but true:

\begin{align*} Z_X\left(\frac{1}{q^{m} z}\right) &=\frac{1}{\left(1-1 / q^{m} t\right)\left(1-q / q^{m} t\right) \cdots\left(1-q^{m} / q^{m} z\right)} \\ &=\frac{q^{m} z \cdot q^{m-1} z \ldots q z \cdot z}{(1-z)(1-q z) \ldots\left(1-q^{m} z\right)} \\ &=q^{m(m+1) \over 2} z^{m+1} \cdot Z_X(z)\\ &= \qty{q^{m\over 2}z}^{\chi(X)} \cdot Z_X(z) \\ .\end{align*}

Checking

\begin{align*} Z_X(z) = \prod_{j=0}^m \qty{1 \over 1 - q^j z} .\end{align*}

  • Riemann Hypothesis: Reduces to the statement \(\left\{{ \alpha_i }\right\} = \left\{{{q^m \over \alpha_j }}\right\}\).
  • Clear since \(\alpha_j = q^j\) and every \(\alpha_i\) is a lower power of \(q\).
  • Betti Numbers: Use the fact that \({\mathcal{P}}_{{\mathbf{CP}}^m}(x) = 1 + x^2 + x^4 + \cdots + x^{2m}\)
  • Only even dimensions, and correspondingly no numerator.

An Easier Proof: “Paving by Affines”

Quick recap: \begin{align*} Z_{{\operatorname{pt}}} = {1 \over 1 - z} && Z_{{\mathbf{P}}^1}(z) = {1 \over 1 - qz} && Z_{{\mathbf{A}}^1}(z) = {1\over (1-z)(1-qz)} .\end{align*}

Note that \({\mathbf{P}}^1 = {\mathbf{A}}^1 {\textstyle\coprod}\left\{{\infty}\right\}\) and correspondingly \(Z_{{\mathbf{P}}^1}(z) = Z_{{\mathbf{A}}^1}(z) \cdot Z_{{\operatorname{pt}}}(z)\). This works in general:

Lemma (Excision)
If \(Y/{ \mathbf{F} }_q \subset X/{ \mathbf{F} }_q\) is a closed subvariety, for \(U = X\setminus Y\), \(Z_X(z) = Z_Y(z) \cdot Z_{U}(z)\).

Proof: Let \(N_n = {\sharp}Y({ \mathbf{F} }_{q^n})\) and \(M_n = {\sharp}U({ \mathbf{F} }_{q^n})\), then \begin{align*} \zeta_X(z) &= \exp\qty{\sum_{n=1}^\infty \qty{N_n + M_n} {z^n \over n} } \\ &= \exp\qty{\sum_{n=1}^\infty N_n \cdot {z^n \over n} + \sum_{n=1}^\infty M_n \cdot {z^n \over n}}\\ &= \exp\qty{\sum_{n=1}^\infty N_n \cdot {z^n \over n}} \cdot \exp\qty{\sum_{n=1}^\infty M_n \cdot {z^n \over n}} = \zeta_Y(z) \cdot \zeta_U(z) .\end{align*}

An Easier Proof: “Paving”

Note that geometry can help us here: we have a decomposition \({\mathbf{P}}^n = {\mathbf{P}}^{n-1} {\textstyle\coprod}{\mathbf{A}}^n\), and thus inductively a stratification \begin{align*} {\mathbf{P}}^m = {\textstyle\coprod}_{j=0}^m {\mathbf{A}}^j = {\mathbf{A}}^0 {\textstyle\coprod}{\mathbf{A}}^1 {\textstyle\coprod}\cdots {\textstyle\coprod}{\mathbf{A}}^m .\end{align*}

Recalling that \begin{align*} Z_{X{\textstyle\coprod}Y}(z) = Z_X(z) \cdot Z_Y(z) \end{align*}

and \(Z_{{\mathbf{A}}^j}(z) = {1 \over 1 - q^j z}\), we have \begin{align*} Z_{{\mathbf{P}}^m}(z) = \prod_{j=0}^m Z_{{\mathbf{A}}^j}(z) = \prod_{j=0}^m {1 \over 1 - q^j z} .\end{align*}

Grassmannians

Motivation

Consider now \(X = {\operatorname{Gr}}(k, m) / { \mathbf{F} }\) – by the previous computation, we know \begin{align*} X({ \mathbf{F} }_{q^n}) = \genfrac{[}{]}{0pt}{}{m}{k}_{q^n} \coloneqq\frac{(q^{nm} - 1)(q^{{nm}-1}-1) \cdots (q^{{nm} - n(k-1)} - 1)}{(q^{nk}-1)(q^{n(k-1)} - 1) \cdots (q^n-1)} \end{align*}

but the corresponding Zeta function is much more complicated than the previous examples: \begin{align*} Z_X(z) &= \exp\qty{ \sum_{n=1}^\infty \genfrac{[}{]}{0pt}{}{m}{k}_{q^n} {z^n \over n} } = \cdots ? .\end{align*}

Since \(\dim_{\mathbf{C}}{\operatorname{Gr}}_{\mathbf{C}}(k, m) = 2k(m-k)\), by Weil we should expect \begin{align*} Z_X(z) = \prod_{j=0}^{2k(m-k)} \frac{p_{2(j+1)}(z)}{p_{2j}(z)} \end{align*} with \(\deg p_j = \beta_j\).

Grassmannian

It turns out that (proof omitted) one can show \begin{align*} \genfrac{[}{]}{0pt}{}{m}{k}_{q} = \sum_{j=0}^{k(m-k)} \lambda_{m, k}(j) q^j \implies Z_X(z) = \prod_{j=0}^{k(m-k)} \qty{ \frac{1}{ 1 - q^j x} }^{\lambda_{m, k}(j)} \end{align*}

where \(\lambda_{m, k}\) is the number of integer partitions of of \([i]\) into at most \(m-k\) parts, each of size at most \(k\).

  • One proof idea: use combinatorial identities to write \(q{\hbox{-}}\)analog \(\genfrac{[}{]}{0pt}{}{m}{k}_{q}\) as a sum
  • Second proof idea: “pave by affines” (need a reference!)

This lets us conclude that the Poincare polynomial of the complex Grassmannian is given by \begin{align*} {\mathcal{P}}_{{\operatorname{Gr}}_{\mathbf{C}}(m, k)}(x) = \sum_{n=1}^{k(m-k)} \lambda_{m, k}(n) x^{2n}, \end{align*}

In particular, the \(H^* {\operatorname{Gr}}_{\mathbf{C}}(m, k)\) vanishes in odd degree.

Weil’s Proof

Diagonal Hypersurfaces

Proof of rationality of \(Z_X(T)\) for \(X\) a diagonal hypersurface.

  • Set \(q\) to be a prime power and consider \(X/{ \mathbf{F} }_q\) defined by \begin{align*}X = V(a_0x_0^{n} + \cdots + a_r x_r^{n}) \subset { \mathbf{F} }_q^{r+1}.\end{align*}
  • We want to compute \(N = {\sharp}X\).
  • Set \(d_i = \gcd(n_i, q-1)\).
  • Define the character \begin{align*} \psi_q: { \mathbf{F} }_q & \to {\mathbf{C}}^{\times}\\ a &\mapsto \exp\qty{2\pi i ~\operatorname{Tr}_{{ \mathbf{F} }_q/{ \mathbf{F} }_p}(a) \over p} .\end{align*}
    • By Artin’s theorem for linear independence of characters, \(\psi_q \not \equiv 1\) and every additive character of \({ \mathbf{F} }_q\) is of the form \(a \mapsto \psi_q(ca)\) for some \(c\in { \mathbf{F} }_q\).
  • Shorthand notation: say \(a\sim 0 \iff a \equiv 0 \operatorname{mod}1\).

A Diagonal Hypersurface

  • Fix an injective multiplicative map \begin{align*} \phi: \mkern 1.5mu\overline{\mkern-1.5mu{ \mathbf{F} }\mkern-1.5mu}\mkern 1.5mu_q^{\times}\to {\mathbf{C}}^{\times} .\end{align*}

  • Define \begin{align*} \chi_{\alpha, n}: { \mathbf{F} }_{q^n}^{\times}&\to {\mathbf{C}}^{\times}\\ x & \mapsto \phi(x)^{\alpha\qty{q^n-1}} \\ \\ \quad {\quad \operatorname{for} \quad} \alpha \in {\mathbf{Q}}/{\mathbf{Z}}, n\in {\mathbf{Z}}, & \quad \alpha\qty{q^n-1} \equiv 0 \operatorname{mod}1 .\end{align*}

    • Extend this to \({ \mathbf{F} }_{q^n}\) by \begin{align*} \begin{cases} 1 & \alpha \equiv0 \operatorname{mod}1 \\ 0 & \text{else} \end{cases} .\end{align*}
    • Set \(\chi_\alpha = \chi_{\alpha, 1}\).
  • Proposition: \begin{align*}\alpha(q-1) \equiv 0 \operatorname{mod}1 \implies \chi_{\alpha, n}(x) = \chi_\alpha(\mathrm{Nm}_{{ \mathbf{F} }_{q^n} / { \mathbf{F} }_q }(x) )\end{align*}

  • Proposition: \begin{align*}d \coloneqq\gcd(n, q-1), u \in { \mathbf{F} }_q \implies {\sharp}\left\{{x\in { \mathbf{F} }_q {~\mathrel{\Big\vert}~}x^n = u}\right\} = \sum_{d\alpha \sim 0} \chi_\alpha(u)\end{align*}

A Diagonal Hypersurface

  • This implies \begin{align*} N &= \sum_{\substack{\alpha = [\alpha_0, \cdots, \alpha_r] \\ d_i \alpha_i \sim 0}} \quad \sum_{\substack{\mathbf{u} = [u_0, \cdots ,u_r] \\ \Sigma~ a_i u_i = 0}} \quad \prod_{j=0}^r \chi_{\alpha_j}(u_j) \\ \\ &= q^r + \sum_{\substack{\alpha,~ \alpha_i \in (0, 1) \\ d_i \alpha_i \sim 0}} \qty{ \prod_{j=0}^r \chi_{\alpha_j}(a_j ^{-1}) \sum_{\Sigma~ u_i=0} \quad \prod_{j=0}^r \chi_{\alpha_j}(u_j) } .\end{align*} since the inner sum is zero if some but not all of the \(\alpha_i \sim 0\).

  • Evaluate the innermost sum by restricting to \(u_0 \neq 0\) and setting \(u_i = u_0 v_i\) and \(v_0 \coloneqq 1\): \begin{align*} \sum_{\Sigma~ u_i=0} \quad \prod_{j=0}^r \chi_{\alpha_j}(u_j) &= \sum_{u_0 \neq 0} \chi_{_{\Sigma ~ \alpha_i}}(u_0) \sum_{\Sigma ~v_i = 0} ~\prod_{j=0}^r \chi_{\alpha_j} (v_j) \\ &= \begin{cases} \qty{q-1} \sum_{\Sigma~ v_i = 0} ~\prod_{j=0}^r \chi_{\alpha_j}(v_j) & {\quad \operatorname{if} \quad} \sum \alpha_i \sim 0 \\ 0 & {\quad \operatorname{else} \quad} \end{cases} .\end{align*}

A Diagonal Hypersurface

  • Define the Jacobi sum for \(\alpha\) where \(\sum \alpha_i \sim 0\): \begin{align*} J(\alpha) \coloneqq\qty {1 \over q-1} \sum_{\Sigma~ u_i = 0} ~\prod_{j=0}^r \chi_{\alpha_j}(u_j) = \sum_{\Sigma~ v_i = 0} ~\prod_{j=1}^r \chi_{\alpha_j}(v_j) \end{align*}

  • Express \(N\) in terms of Jacobi sums as \begin{align*} N = q^r + \qty{q-1} \sum_{\substack{\Sigma \alpha_i \sim 0 \\ d_i \alpha_i \sim 0 \\ \alpha\in (0, 1)}} \prod_{j=0}^r \chi_{\alpha_j}(a_j^{-1}) J(\alpha) .\end{align*}

  • Evaluate \(J(\alpha)\) using Gauss sums: for \(\chi: { \mathbf{F} }_q \to {\mathbf{C}}\) a multiplicative character, define \begin{align*} G(\chi) &\coloneqq\sum_{x\in { \mathbf{F} }_q} \chi(x) \psi_q(x) .\end{align*}

  • Proposition: for any \(\chi \neq \chi_0\),

    • \({\left\lvert {G(\chi)} \right\rvert} = q^{1 \over 2}\)
    • \(G(\chi) G(\mkern 1.5mu\overline{\mkern-1.5mu\chi\mkern-1.5mu}\mkern 1.5mu) = q \chi(-1)\)
    • \(G(\chi_0) = 0\) \begin{align*} \chi(t) = {G(\chi) \over q} \sum_{x\in { \mathbf{F} }_q} \mkern 1.5mu\overline{\mkern-1.5mu\chi\mkern-1.5mu}\mkern 1.5mu(x) \psi_q(tx) .\end{align*}

A Diagonal Hypersurface

  • Proposition: if \begin{align*} \sum \alpha_i \sim 0 \implies J(\alpha) = {1 \over q} \prod_{k=1}^r G(\chi_{\alpha_k}) {\quad \operatorname{and} \quad} {\left\lvert {J(\alpha)} \right\rvert} = q^{r - 1\over 2} .\end{align*}

  • We thus obtain \begin{align*} N = q^r + \qty{q-1 \over q} \sum_{\substack{\Sigma \alpha_i \sim 0 \\ d_i \alpha_i \sim 0 \\ \alpha\in (0, 1)}} ~\prod_{j=0}^r \chi_{\alpha_j}(a_j^{-1}) G(\chi_{\alpha_j}) .\end{align*}

  • We now ask for number of points in \({ \mathbf{F} }_{q^\nu}\) and consider a point count \begin{align*} \mkern 1.5mu\overline{\mkern-1.5muN\mkern-1.5mu}\mkern 1.5mu_\nu = {\sharp}\left\{{[x_0: \cdots : x_r] \in {\mathbf{P}}^r_{{ \mathbf{F} }_q^\nu} {~\mathrel{\Big\vert}~}\sum_{i=0}^r a_i x_i^n = 0}\right\} .\end{align*}

  • Theorem (Davenport, Hasse) \begin{align*}\qty{q-1}\alpha \sim 0 \implies -G(\chi_{\alpha, \nu}) = \qty{-G(\chi_\alpha)}^\nu.\end{align*}

A Diagonal Hypersurface

  • We have a relation \(\qty{q^\nu - 1} \mkern 1.5mu\overline{\mkern-1.5muN\mkern-1.5mu}\mkern 1.5mu_\nu = N_\nu\).
  • This lets us write \begin{align*} \mkern 1.5mu\overline{\mkern-1.5muN\mkern-1.5mu}\mkern 1.5mu_\nu = \sum_{j=0}^{r-1} q^{j\nu} + \sum_{\substack{\sum \alpha_i ~\sim 0 \\ \gcd(n, q^\nu - 1)\alpha_i \sim 0 \\ \alpha_i \in (0, 1) }} \prod_{j=0}^r \mkern 1.5mu\overline{\mkern-1.5mu\chi\mkern-1.5mu}\mkern 1.5mu_{\alpha_{j, \nu}}(a_i) J_\nu(\alpha) .\end{align*}
  • Set \begin{align*} \mu(\alpha) &= \min\left\{{\mu {~\mathrel{\Big\vert}~}\qty{q^\mu - 1} \alpha \sim 0}\right\} \\ C(\alpha) &= (-1)^{r+1} \prod_{j=1}^r \mkern 1.5mu\overline{\mkern-1.5mu\chi\mkern-1.5mu}\mkern 1.5mu_{\alpha_0, \mu(\alpha)}(a_j) \cdot J_{\mu(\alpha)}(\alpha) .\end{align*}
  • Plugging into the zeta function \(Z\) yields \begin{align*} \exp\qty{\sum_{\nu = 1}^\infty \mkern 1.5mu\overline{\mkern-1.5muN\mkern-1.5mu}\mkern 1.5mu_\nu {T^\nu \over \nu} } = \prod_{j=0}^{r-1} \qty{1 \over 1 - q^j T} \prod_{\substack{\sum \alpha_i \sim 0 \\ \gcd(n, q^\nu - 1)\alpha_i \sim 0 \\ \alpha_i \in (0, 1) }} \qty{1 - C(\alpha) T^{\mu(\alpha)}}^{(-1)^r \over \mu(\alpha)} ,\end{align*} which is evidently a rational function! \(\hfill\blacksquare\)
#projects/my-talks #homotopy #arithmetic-geometry/Weil-conjectures #projects/my-talks/slides